Photo credits
p.132 sciencephotos/Alamy Stock Photo; p.146 Natrij, used under a
CC BY-SA 3.0 license; p.149 Martyn F. Chillmaid/Science Photo
Library.
Every effort has been made to trace all copyright holders, but if any
have been inadvertently overlooked the Publishers will be pleased to
make the necessary arrangements at the first opportunity.
Although every effort has been made to ensure that website
addresses are correct at time of going to press, Hodder Gibson
cannot be held responsible for the content of any website mentioned
in this book. It is sometimes possible to find a relocated web page by
typing in the address of the home page for a website in the URL
window of your browser.
Hachette UK’s policy is to use papers that are natural, renewable
and recyclable products and made from wood grown in well-
managed forests and other controlled sources. The logging and
manufacturing processes are expected to conform to the
environmental regulations of the country of origin.
Orders: please contact Hachette UK Distribution, Hely Hutchinson
Centre, Milton Road, Didcot, Oxfordshire, OX11 7HH. Telephone:
+44 (0)1235 827827. Email
[email protected]. Lines are
open from 9 a.m. to 5 p.m., Monday to Friday. You can also order
through our website: www.hoddereducation.co.uk
© Dr Sian Simmonds 2021
First published in 2021 by
Hodder Gibson, an imprint of Hodder Education,
An Hachette UK Company
211 St Vincent Street
Glasgow, G2 5QY
Impression number 54321
Year 2025 2024 2023 2022 2021
All rights reserved. Apart from any use permitted under UK copyright
law, no part of this publication may be reproduced or transmitted in
any form or by any means, electronic or mechanical, including
photocopying and recording, or held within any information storage
and retrieval system, without permission in writing from the publisher
or under licence from the Copyright Licensing Agency Limited.
Further details of such licences (for reprographic reproduction) may
be obtained from the Copyright Licensing Agency Limited, Saffron
House, 6-10 Kirby Street, London EC1N 8TS.
Cover photo © jonnysek - stock.adobe.com
Illustrations by Aptara, Inc.
Typeset in CronosPro 13/15 pts. by Aptara, Inc.
Printed in Spain
A catalogue record for this title is available from the British Library
ISBN: 978 1 3983 1220 3
eISBN: 978 1 3983 1199 2
Contents
Introduction
Unit 1 Inorganic chemistry
Chapter 1 Electromagnetic spectrum
Chapter 2 Shapes of atomic orbitals and quantum numbers
Chapter 3 Orbital notations
Chapter 4 Ionisation energy
Chapter 5 Shapes of molecules and polyatomic ions
Chapter 6 Transition metals, electronic configuration and oxidation
numbers
Chapter 7 Ligands and transition metal complexes
Chapter 8 Colours of transition metals and catalysts
Unit 1 Glossary
Unit 1 Exam-style questions
Unit 2 Physical chemistry
Chapter 9 Chemical equilibrium and factors affecting equilibrium
Chapter 10 Chemical equilibrium and the ionic product of water
Chapter 11 Strong and weak acids and bases
Chapter 12 Acid dissociation constant and salts
Chapter 13 Indicators and buffers
Chapter 14 Reaction feasibility, standard enthalpy of formation and
entropy
Chapter 15 Gibbs free energy
Chapter 16 Kinetics
Unit 2 Glossary
Unit 2 Exam-style questions
Unit 3 Organic chemistry and instrumental analysis
Chapter 17 Molecular orbitals and hybridisation
Chapter 18 Molecular orbitals and colour
Chapter 19 Homolytic and heterolytic fission
Chapter 20 Electrophiles, nucleophiles and reaction types
Chapter 21 Alkenes
Chapter 22 Haloalkanes including SN1 and SN2 mechanisms
Chapter 23 Alcohols
Chapter 24 Carboxylic acids
Chapter 25 Ethers
Chapter 26 Amines
Chapter 27 Benzene
Chapter 28 Stereochemistry (structural, geometric and optical
isomers)
Chapter 29 Experimental determination of structure using empirical
microanalysis and mass spectrometry
Chapter 30 Experimental determination of structure using NMR and
infrared
Chapter 31 Pharmaceutical chemistry
Unit 3 Glossary
Unit 3 Exam-style questions
Unit 4 Researching chemistry
Chapter 32 Errors and significant figures
Chapter 33 Practical skills and techniques
Chapter 34 Standard solutions and dilutions
Chapter 35 Volumetric analysis
Chapter 36 Gravimetric analysis
Chapter 37 Chromatography
Chapter 38 Colorimetry
Unit 4 Glossary
Unit 4 Exam-style questions
Appendix: Additional features of the Advanced Higher Chemistry
exam
Numeracy
Open-ended questions
Answers to Study questions and Exam-style questions
Introduction
Welcome to How to Pass Advanced Higher Chemistry. This book
follows the updated 2019 SQA arrangements for the Advanced
Higher Chemistry course.
Using this book
The content of this guide is designed to help you prepare for the
SQA Advanced Higher Chemistry exam. It provides a complete
summary of the Advanced Higher course and offers the following
features to aid understanding and improve recall of important points.
Key points
Key points
• Key point summaries at the start of each chapter detail the
knowledge and skills required by the SQA.
• Once you have read the chapter and completed the questions, go
back and check the key points to ensure that you have gained the
required knowledge from this chapter.
• Using the key points and the study questions, you should be able
to test your knowledge of the topic.
Hints & tips
Hints & tips
This feature gives you extra hints and tips and draws your attention
to particular things to think about and remember for each topic.
Examples
Examples
Throughout the book, you will find sets of worked examples to test
your knowledge of the concepts in each chapter and show how to
answer the questions at Advanced Higher level.
Before looking at the exemplar answer provided, try to recollect
how to answer such questions from the information you have read
in the chapter. It is essential for your learning to make every effort
to recall information; this effort of mentally searching for an answer
helps to strengthen your memory and will boost your ability to learn.
Review your answer against the exemplar provided.
Study questions
Study questions
Study questions are found at the end of each chapter. They should
be used to test your knowledge and understanding of the chemistry
covered. The questions included have been selected carefully and
some are modelled on those found in previous SQA Advanced
Higher examinations. Like the worked examples, you should always
try to answer the question using your own recall of the topic before
checking the answers given in the book.
Glossary
Electronegativity: Electronegativity is a measure of the attraction
an atom involved in a bond has for the electrons of the bond.
Frequency: This is the number of wavelengths that pass a fixed
point in one unit of time.
Exam-style questions
At the end of each unit you will find a number of exam-style
questions to work through. Answers to these can be found, along
with answers to all the Study questions, at the end of the book.
Additional features of the Advanced Higher
Chemistry exam
This section appears at the end of the book and offers advice on
how to tackle two important features of the Advanced Higher
Chemistry exam: numeracy and open-ended questions. Both are
explored in detail and are supported with worked examples and
study questions. Again, you should practise the exemplars and study
questions before comparing them against the worked examples.
The Advanced Higher
Chemistry course
The Advanced Higher Chemistry course comprises four areas which
are covered in the four units of this How to Pass guide.
Key area Topics covered
Inorganic Electromagnetic radiation and atomic spectra,
chemistry atomic orbitals; electronic configurations and the
Periodic Table; transition metals
Physical Chemical equilibrium; reaction feasibility; kinetics
chemistry
Organic Molecular orbitals; synthesis; stereochemistry;
chemistry experimental determination of a structure;
and pharmaceutical chemistry
instrumental
analysis
Researching Common chemical apparatus; skills involved in
chemistry experimental work; stochiometric calculations;
gravimetric analysis; volumetric analysis; practical
skills and techniques
Advanced Higher Chemistry
assessment
There are two parts to the Advanced Higher Chemistry assessment:
• project
• examination paper.
Project
The project is worth 25 marks. This is scaled by SQA to represent
25% of the overall marks for the course assessment. The project
allows you to carry out an in-depth investigation of a chemistry topic
of your choosing and produce a project report. You will be required
to plan and carry out a chemistry investigation on your own, after
discussing potential topics with your teacher or lecturer to ensure
you do not waste time researching unsuitable topics. You will need to
work autonomously, making independent and rational decisions
based on evidence and interpretation of scientific information, by
analysing and evaluating your results. This will help develop and
enhance your scientific literacy skills, and offers challenge by
requiring you to apply skills, knowledge and understanding in a new
context.
Examination question paper
The question paper has 110 marks. This is scaled by SQA to
represent 75% of the overall marks for the course assessment.
The question paper has two sections. Section 1 contains multiple-
choice questions and is worth 25 marks. Section 2 contains
restricted-response and extended-response questions and is worth a
total of 85 marks. This written exam lasts for 3 hours.
Advice for all Advanced Higher
Chemistry students
Advanced Higher Chemistry requires you to have a great deal of
knowledge and understanding of the topics studied, as well as the
confidence to answer standard questions that may be asked in the
exam. This book presents you with all the chemical knowledge you
will need in summarised form and includes likely question types you
will be asked.
The difficult part is learning and understanding this knowledge and
remembering how to do the standard questions. Try to get into the
habit of reviewing the relevant chapters at the same time as you are
studying the concepts in class. To ‘review’ means to read and test
your understanding and recall. Using the worked examples and
study questions in this book will help to reinforce the information you
have read. The more you retrieve the information, the more easily
knowledge will come to you. In other words, you need to review and
attempt the questions in each chapter several times.
Make sure to space out revision over a long time period to ensure
maximum retrieval from long-term memory. Research has shown
that it is a more effective strategy than simply doing the same
questions over and over. Once you have mastered one chapter and
checked your understanding, you should move on to another chapter
and repeat the process. You should then return to the original
chapter a week or two later and test your understanding again. You
may well have forgotten some knowledge and so will have to review
the questions again. This will work to reinforce the information in
your memory.
Specific advice is given at the back of the book for tackling numeracy
and open-ended questions. These are questions which most
students avoid tackling when revising or in the final exam. To
achieve greater success, it is necessary to practise answering these
types of questions when revising. This helps to build confidence and
the skills needed to face these questions in an examination setting.
I hope that this book will help you during your final chemistry studies
at school. Even if you do not choose to pursue chemistry at
university level, it is hoped that this course will have given you a
greater understanding of the use of chemistry in everyday life from
medicines to food production.
Unit 1 Inorganic chemistry
Chapter 1
Electromagnetic spectrum
Key points
• Electromagnetic radiation can be described as a wave (with a
wavelength and frequency) and as a particle; it is said to have a
dual nature.
• The relationship between wavelength (λ) and frequency (f) is
given by:
c=f×λ
where c is the speed of light in a vacuum, 3.00 × 108 m s−1.
• When electromagnetic radiation is absorbed or emitted by matter,
it behaves like a stream of particles; these particles are known as
photons. The equation
E = hf
where h is Planck’s constant, 6.63 × 10−34 J s, can be used to
calculate the energy associated with a single photon. The energy
associated with one mole of photons is given by
where L is Avogadro’sconstant, 6.02 × 1023 mol−1.
• The energy is often given in units of kJ mol–1.
• When a photon is absorbed or emitted, the electrons within the
substance gain or lose energy.
• Photons in high-frequency radiation transfer more energy than
photons in lower-frequency radiation.
• In emission spectroscopy, high temperatures are used to excite
the electrons within atoms in the sample and as the electrons
drop to lower energy levels, photons are emitted and measured.
An emission spectrum of a sample is produced by measuring the
intensity of light emitted at different wavelengths.
• When photons of light energy are absorbed by atoms, electrons
move from a lower energy level to a higher one. As this happens,
an absorption spectrum is produced. The difference between the
intensity of absorbed light and the transmitted light gives the
spectrum, and it varies with wavelength.
• Each element produces a unique pattern of frequencies of
radiation in both its emission and absorption spectra, and these
can be used to identify and quantify the elements present within a
sample.
• The concentration of an element within a sample is related to the
intensity of light emitted or absorbed and this can be used to
determine unknown concentrations of an element.
Electromagnetic radiation
Electromagnetic radiation is a form of energy. The electromagnetic
spectrum consists of many bands of electromagnetic radiation
which differ in terms of energy (E), wavelength (λ) and frequency (f).
Visible light is only a small part of the electromagnetic spectrum.
Visible light is split into seven different colours that have different
wavelengths and frequencies.
Electromagnetic radiation can be described as a wave and as a
particle and is said to have a dual nature.
Hints & tips
The colour order in the visible light electromagnetic spectrum can
be remembered from the mnemonic ROY G BIV.
Figure 1.1 The electromagnetic spectrum
Hints & tips
Frequency and wavelength are inversely proportional: as frequency
increases, wavelength decreases and vice versa.
When using the wave model to describe electromagnetic radiation,
the waves can be specified by their wavelength and frequency:
• The wavelength of a wave (given the symbol lambda, λ) is the
distance between adjacent crests or adjacent troughs and in
chemistry is typically measured in nanometres (1 nm = 1 × 10–9
m).
• The frequency (symbol, f) is determined by the number of
wavelengths which pass a fixed point in one second. This is
measured as the reciprocal of time (s–1) which is also called hertz
(Hz).
All electromagnetic radiation travels at the same velocity. This
constant is the speed of light (symbol, c) and, in a vacuum, it is
approximately equal to 3.00 × 108 m s–1. Frequency and wavelength
are related by the formula:
c = wavelength × frequency
or
c = f × λ.
Examples
1 Electromagnetic radiation is found to have a wavelength of 1100
nm. Calculate the frequency of this radiation. (Give your answer
to three significant figures.)
Answer
c=f×λ
Rearrange this formula to find f.
2 Electromagnetic radiation is found to have a frequency of 6 ×
1012 Hz. Calculate the wavelength of this radiation. (Give your
answer to three significant figures.)
Answer
c=f×λ
Rearrange this formula to find λ.
Electromagnetic radiation is useful in chemistry as it can be both
absorbed and emitted. When electromagnetic radiation is absorbed
or emitted it behaves like a stream of particles (so we have to use
the particle model rather than the wave model). These particles are
called photons. When a photon is emitted or absorbed, energy is
lost or gained by the electrons in the sample being studied. Photons
which are at high frequencies transfer larger amounts of energy than
photons in lower frequencies.
Hints & tips
That violet/UV is high energy is easily remembered by the fact that
UV radiation from the Sun causes sunburn and skin cancers.
So there is a relationship between the energy (E) carried by a photon
and its frequency. This relationship is given by the equation:
E = hf
where h is Planck’s constant and has the value 6.63 × 10–34 J s. In
chemistry, we use kJ not J so to convert J to kJ, the number value
must be divided by 1000.
To obtain the quantity of energy in kJ mol–1, the relationship of E has
to take into account the moles present, so uses Avogadro’s constant
(L), 6.02 × 1023 mol–1. So the energy in kJ mol–1 is given by the
equation:
When relating energy to wavelength, the equation to use is
Hints & tips
The values of Planck’s constant (h), the speed of light in a vacuum
(c) and Avogadro’s constant (L) are provided on the back page of
the data booklet along with SI prefixes and multiplication factors.
Example
3 Calculate the wavelength, in nm, of light required to break 1 mole
of H—Cl bonds. Use the bond enthalpy data from the data
booklet.
Answer
H–Cl bond enthalpy = 432 kJ mol–1
Rearrange this formula to find λ.
The question asks for the answer in nm so multiply by 109.
λ = 2.77 × 10–7 × 109 = 277 nm
Atomic emission spectra
When a beam of white light is passed through a prism or from a
diffraction grating onto a screen, a continuous spectrum is visible.
When white light from a source passes through a sample being
heated, the spectrum turns out not to be a continuous spectrum, but
a series of lines of different wavelengths and thus of different
colours. Each line corresponds to the energy given out when excited
electrons move down to a lower energy level, generating photons of
various frequencies.
These lines correspond to certain specific frequencies and
wavelengths found in the visible or ultraviolet spectrum. Each
element creates its own unique spectrum, something like a chemical
fingerprint, with its own specific frequencies and wavelengths.
Atomic absorption
spectroscopy
When a beam of continuous radiation like white light is directed
through a gaseous sample, it can cause an atom to make a
transition from its ground state to an excited state. If the frequency of
the light, and therefore the energy of the photon, corresponds to an
excitation energy of the atom, then the photon of light is absorbed.
The radiation that emerges will therefore have certain wavelengths
missing. These show up as dark lines on a continuous spectrum
called an atomic absorption spectrum (Figure 1.2).
Figure 1.2 Absorption and emission spectra for hydrogen and
helium
Both emission and absorption spectroscopy can be used to
determine whether a certain species is present in a sample and how
much of it is present, since the intensity of the transmitted or
absorbed radiation can be measured.
A calibration graph is first made using known concentrations of the
species being analysed. The radiation absorbed or emitted by the
species in these samples is then plotted against concentration.
When the unknown sample is analysed, the concentration of the
species can be found from the graph by reading off the concentration
of the known sample with the same absorbance.
Figure 1.3 A calibration graph of absorbance against
concentration
Study questions
1 Electromagnetic radiation is found to have a wavelength of 900
nm. Calculate the frequency of this radiation. (Give your answer
to three significant figures.)
[3 marks]
2 Electromagnetic radiation is found to have a frequency of 5 × 109
Hz. Calculate the wavelength of this radiation. (Give your answer
to three significant figures.)
[3 marks]
3 A line in the visible spectrum is caused by an electron transition
with an associated energy of 264 kJ mol–1. Calculate the
wavelength of this line in nanometres.
[3 marks]
4 Strontium is an element discovered on the west coast of
Scotland. When strontium is placed in a flame, a red colour is
observed which has a very prominent line in the spectrum.
a) Refer to your data booklet to determine the wavelength of this
line.
[2 marks]
b) Calculate the energy, in kJ mol–1, of the electron transition
related to this line.
[2 marks]
Chapter 2
Shapes of atomic orbitals and
quantum numbers
Key points
• The discrete lines observed in atomic spectra can be explained if
electrons, like photons, display the properties of both particles
and waves.
• Electrons behave as standing (stationary) waves in an atom.
These are waves that vibrate in time but do not move in space.
• There are different sizes and shapes of standing wave possible
around the nucleus, known as orbitals. Orbitals can hold a
maximum of two electrons.
• There are four types of orbitals (s, p, d and f), each with a
characteristic shape or set of shapes. Diagrams of the shapes of
s and p orbitals can be drawn and recognised. Diagrams of d
orbitals should be recognised.
• Electrons, like photons, display both wave- and particle-like
properties. Atomic orbitals are regions where there is a high
probability of finding the electrons.
• Quantum mechanics provides us with equations that predict the
shape and orientation of these orbitals based on four quantum
numbers that are unique for each electron.
• These four numbers are: principal quantum number n
(representing energy), angular momentum quantum number l
(representing shape), magnetic quantum number ml (representing
orientation) and spin magnetic quantum number ms (these are
opposite for paired electrons).
• An orbital holds a maximum of two electrons of opposite spin.
• In an isolated atom, the orbitals within each shell are said to be
degenerate.
Atomic orbital shapes
It is not possible to define a point in space where an electron is to be
found as they are constantly moving. However, the probability of
finding an electron within a certain volume of space can be
calculated. The regions where there is a high probability of finding an
electron are called atomic orbitals. In these areas, there is
approximately a 90% chance of detecting an electron. So, an orbital
provides a defined volume of space where there is a high probability
of finding an electron.
There are four orbital types: s, p, d and f. These letters were chosen
on the basis of observations of line spectra. Certain lines were
observed as a ‘sharp’, ‘principal’, ‘diffuse’ or ‘fundamental’ series,
hence s, p, d, f.
All s orbitals are spherical in shape. An s orbital in the first shell is
smaller in diameter than an s orbital in the second shell. Their
diameter increases as the shell number increases. They are the only
orbital present in the first shell. Each s orbital can only hold two
electrons of opposite spins.
The p orbitals are dumbbell shaped and they lie along the x-, y- and
z-axes as shown in Figure 2.1. p orbitals only occur from the second
shell onwards. These orbitals all have equal energy and are said to
be degenerate. The second shell can hold a maximum of eight
electrons. The s orbital can hold the first two of these electrons and
the p orbitals can each hold two electrons. Therefore, the second
shell has one s and three p orbitals.
Figure 2.1 Degenerate p orbitals aligned along three axes
The third shell can hold a maximum of 18 electrons. Two electrons
are held in the s orbital and another six electrons are held in the
three p orbitals. This means that ten electrons must be held in the d
orbitals. Each orbital can only hold two electrons so there must be
five d orbitals present in the third shell onwards. The shapes and
orientations of the d orbitals are shown in Figure 2.2.
Figure 2.2 Degenerate d orbitals and their orientations with
respective to the three axes
Elements with more than one electron produce more complex
emission spectra than hydrogen. When these are examined under
high resolution it shows that the lines are often not single lines but
are split into doublets or triplets, and so on. This provides evidence
that the electron shells are further subdivided into s, p, d and f
orbitals. Calculations using quantum mechanics show that all shells
have an s orbital and all the shells except the first have p orbitals. All
the shells except the first and second have d orbitals and so forth.
At Advanced Higher level, there is no need to learn about the shape
or orientation of f orbitals.
Hints & tips
Learn how to draw unlabelled s and p orbitals. Be able to recognise
the d orbitals and identify the correct axes.
Quantum numbers
Emission spectra of elements with more than one electron show
evidence of these atomic orbitals. The theory of quantum mechanics
was put forward to explain these atomic orbitals and explain why an
electron can show properties of both waves and particles.
An electron can only possess certain fixed amounts of energy known
as quanta. The basis of quantum theory states that the energy of an
electron can be defined in terms of four quantum numbers. Each
electron in any element will have its own individual set of quantum
numbers. This provides a type of ‘address’ for each electron in an
atom.
1 Each shell is described by a number, known as the principal
quantum number, n. This number indicates the electron’s main
energy level and is related to the orbital size.
The shells are numbered starting with the shell nearest to the
nucleus and working outwards. So, the first shell has a principal
quantum number n = 1, the second shell n = 2 and so forth. The
higher the value of the principal quantum number n, the higher the
potential energy associated with the shell and the further from the
nucleus the electron is likely to be found.
2 The shape of the subshell is given by the angular momentum
quantum number, l. l can have values from 0 to n – 1. If the
angular momentum quantum number l = 0, the electron is found in
an s-shaped orbital. If l = 1, then the electron is in a p-shaped
orbital; when l = 2, the electron is in a d-shaped orbital and when l
= 3, the electron is in a f-shaped orbital.
3 The orientation in space of the electron in the different degenerate
p, d and f orbitals is given by the magnetic quantum number, ml.
This number can have values between – l and +l. The angular
momentum quantum number has an effect on the values of the
magnetic quantum number. If n = 2 and l = 1, then ml will
correspond to the orientation of the p orbital and will have values
ml = –1, 0, +1 which correspond to the 2px, 2py and 2pz
orientations. If n = 3 and l = 2, then ml will correspond to the d
orbitals and will have values ml = –2, –1, 0, +1, +2 which
corresponds respectively to the 3dxy, 3dxz, 3dyz, 3dx2–y2 and 3dz2
orientations.
4 The final quantum number is the spin magnetic quantum
number, ms. This indicates the direction in which an electron is
spinning. This can be either clockwise or anticlockwise and has
one of two values: and No two electrons occupying the
same atomic orbital can
have the same spin; they must have opposite spins.
Using the four quantum numbers, n (shell), l (shape), ml (orientation)
and ms (spin), we can describe every electron in an atom.
Examples
1 Which line in the table could represent the quantum numbers for
one of the outer electrons in a sodium ion, Na+?
Answer
A: A sodium ion has the electronic configuration 2,8. The outer
electrons are in the second shell so the principal quantum
number must be n = 2. l = 1 suggests the p orbital and then ml =
0 suggests the 2py orientation. The spin is given by ms This
is the only line which would correspond to the quantum numbers
for the sodium ion.
2 State the values of the four quantum numbers (n, l, ml and ms)
for one of the electrons in the 3s orbital of sulfur in its ground
state.
Answer
n = 3 indicates the 3rd shell and l = 0 indicates the s orbital. ml =
0 as the s orbital has only one orientation and ms is either
So the answer can be either n = 3, l = 0, ml = 0 and ms
or n = 3, l = 0, ml = 0 and ms
Study questions
1 How many atomic orbitals make up the 4d subshell?
[1 mark]
2 The table shows four sets of quantum numbers.
a) Explain which set of numbers is not allowed according to
quantum theory.
[1 mark]
b) Which set of numbers represents an s subshell?
[1 mark]
c) Which set of numbers could belong to a hydrogen atom in its
ground state?
[1 mark]
Hints & tips
Remember that the maximum number of electrons in any orbital is
two electrons with opposite spins.
Chapter 3
Orbital notations
Key points
• The orbitals within each subshell of an isolated atom are
degenerate.
• Within atoms, electrons are arranged according to three rules:
the Pauli exclusion principle, the Aufbau principle and Hund’s
rule.
• The Pauli exclusion principle states that no two electrons in one
atom can have the same set of four quantum numbers. In other
words, no orbital can hold more than two electrons, and these two
electrons must have opposite spins.
• The Aufbau principle states that orbitals are filled in order of
increasing energy.
• Hund’s rule states that when degenerate orbitals are available,
electrons fill each singly, keeping their spins parallel before spin
pairing starts.
• The relative energies corresponding to each orbital can be
represented diagrammatically for the first four shells of a multi-
electron atom.
• Atomic numbers can be used to write the electronic configuration
of any element using spectroscopic notation and orbital box
notation. At Advanced Higher you are required to do this for
elements 1 to 36.
Rules for arranging electrons in
orbitals
Electrons occupy energy levels or shells when orbiting the nucleus.
In the previous chapter it was seen that some of these shells are
divided into subshells, in which the orbitals have the same energy.
Electrons are placed into an orbital according to certain rules:
• The first of these rules is called the Pauli exclusion principle.
This states that no two electrons in the same atom can have the
same four quantum numbers. So no orbital can hold more than two
electrons, and the two electrons must have opposite spins.
• The Aufbau principle states that electrons fill the subshells in
order of increasing energy. This means that the lowest energy
subshells are filled first. So an s subshell is filled before a p
subshell, which is filled before a d subshell. All the p and all the d
orbitals within a subshell are of equal energy. The 3d subshell is at
a slightly higher energy level than the 4s subshell. Spectroscopic
data gives the following arrangement of the energies of the
orbitals: 1s 2s 2p 3s 3p 4s 3d 4p 5s 4d 5p 6s 4f 5d 6p 7s 5f etc. So
the filling order is as shown in Figure 3.1.
• The final rule is Hund’s rule which states that when degenerate
orbitals are available, electrons fill each degenerate orbital singly
and with parallel spins, before pairing up to fill the orbitals.
Hints & tips
Degenerate orbitals have the same energy.
Hints & tips
The 4s orbital has lower energy than the 3d orbital.
Figure 3.1 The order in which the orbitals are filled. Follow the
arrow.
Hints & tips
Learn the definitions of the three rules: the Pauli exclusion principle,
the Aufbau principle and Hund’s rule. Understand how to apply
them.
Electronic configuration
When electrons are placed in an orbital, a number is written (as a
superscript) after the subshell notation. For example, one electron in
the 1s shell is written as 1s1 and six electrons in the p subshell is
written as 2p6. This is called the electronic configuration. The
electronic configurations are given on page 8 of the data booklet, but
this just gives the number of electrons in each principal shell; it does
not tell you which subshell they are found in. The electron
arrangement of oxygen is 2,6 and so the electronic configuration of
an oxygen atom is given as 1s2 2s2 2p4.
With the exception of the transition metals, atoms form ions which
have the electronic configuration of the nearest noble gas. An
oxygen atom gains two electrons to form an O2− ion. The electron
arrangement of an oxygen ion is 2,8 and so the electronic
configuration of an oxygen ion is 1s2 2s2 2p6. Particles which have
the same electronic configuration are said to be isoelectronic. An
atom of neon and a sodium ion (Na+) also have the electronic
configuration of 1s2 2s2 2p6 and so are isoelectronic with the oxygen
ion (O2−).
The atoms of chromium and copper have unusual electronic
configurations because of the stability of filled and half-filled orbitals.
Chromium has an electron arrangement of 2,8,13,1 and the
electronic configuration is 1s2 2s2 2p6 3s2 3p6 3d5 4s1 rather than
3d4 4s2. This is because having one electron in each of the 3d
orbitals makes it more stable. It also means they are symmetrical
around the nucleus.
Copper has an electron arrangement of 2,8,18,1 and so the
electronic configuration is 1s2 2s2 2p6 3s2 3p6 3d10 4s1 rather than
3d9 4s2. The ten electrons in the 3d orbitals make the atom more
stable due to the symmetry around the nucleus.
Hints & tips
Remember that chromium and copper both have unusual electronic
configurations so put a star beside them in your data-booklet
electron arrangement before you start your exam or test.
Note that when writing out the electronic configuration, you must
write it in order of principal quantum number with the accompanying
s, p, d, f notation: so the order of writing out is 1s 2s 2p 3s 3p 3d 4s
4p 4d 4f 5s 5p 5d but the order of orbital filling is 1s 2s 2p 3s 3p 4s
3d 4p 5s 4d 5p.
When the electrons are in their lowest possible energy levels, the
atom is said to be in the ground state.
Examples
1 Write the electronic configuration of a potassium atom.
Answer
The electron arrangement of a potassium atom is 2,8,8,1, so the
electronic configuration is 1s2 2s2 2p6 3s2 3p6 4s1.
The 3d and 4p orbitals are not occupied by electrons and so are
not written in the electron configuration.
2 Write the electronic configuration of a sulfide ion, S2−.
Answer
The electron arrangement of a sulfur atom is 2,8,6, so the
electronic configuration is 1s2 2s2 2p6 3s2 3p4.
When the sulfur atoms form sulfide ions, they gain two electrons
to give them the same electron arrangement as argon. So the
electronic configuration of a sulfide ion is 1s2 2s2 2p6 3s2 3p6.
3 Write the electron configuration of an iron(III) ion.
Answer
The electron arrangement of an iron atom is 2,8,14,2, so the
electronic configuration is 1s2 2s2 2p6 3s2 3p6 3d6 4s2.
When the iron atoms form iron(III) ions, they lose three electrons
to become Fe3+. Iron is a d-block element, and these elements
lose their 4s electrons first. So it loses two electrons from the 4s
and one electron from the 3d. The electronic configuration of an
iron(III) ion is 1s2 2s2 2p6 3s2 3p6 3d5.
Orbital box notation
From the previous chapter we know that there are three p orbitals in
the p subshell and five d orbitals in the d subshell. One p orbital lies
along the x-axis and is px; the py lies along the y-axis and the pz lies
along the z-axis. The different p orbitals are degenerate. Following
Hund’s rule, each degenerate orbital must be filled singly before the
electrons of a different spin are paired up. This allows the Pauli
exclusion principle to be followed. Electronic configuration illustrates
the total number of electrons in the subshell but does not show the
electrons in the degenerate orbitals; to show this, orbital box
notation is used. This notation shows the orbitals as boxes, with
arrows being used to represent the electrons. Electrons of opposite
spins are represented by arrows pointing in opposite directions
(Figures 3.2, 3.3 and 3.4).
Figure 3.2 The electronic configuration for an atom of nitrogen
is given as 1s2 2s2 2p3 and the orbital box notation is as shown
Figure 3.3 The electronic configuration for an atom of calcium
is given as 1s2 2s2 2p6 3s2 3p6 4s2 and the orbital box notation
is as shown
Figure 3.4 The electronic configuration for an atom of
vanadium is shown by 1s2 2s2 2p6 3s2 3p6 3d3 4s2 and the
orbital box notation is as shown
Examples
4 Draw the electronic configuration, using orbital box notation, for
the calcium ion.
Answer
The electronic configuration for an ion of calcium (Ca2+) is 1s2
2s2 2p6 3s2 3p6 and the orbital box notation is written as shown
in Figure 3.5.
Figure 3.5
5 Draw the electronic configuration, using orbital box notation, for
the nickel atom.
Answer
The electronic configuration for an atom of nickel is 1s2 2s2 2p6
3s2 3p6 3d8 4s2 and the orbital box notation is written as shown
in Figure 3.6.
Figure 3.6
Hints & tips
When writing out the electronic configuration or orbital box notation,
ensure that you write them out in order of principal quantum
number, not fill order.
Study questions
1 Write the electronic configuration of a potassium ion.
[1 mark]
2 Write the electronic configuration of a krypton atom.
[1 mark]
3 Draw the electronic configurations, using orbital box notation, for
the gallium and oxygen atoms.
[2 marks]
4 Draw the electronic configurations, using orbital box notation, for
the bromide and aluminium ions.
[2 marks]
5 Which of these is isoelectric with an argon atom?
A an oxygen atom
B a fluoride ion
C a sulfide ion
D a magnesium atom.
[1 mark]
Chapter 4
Ionisation energy
Key points
• First, second and subsequent ionisation energies with increasing
atomic number for the first 36 elements can be explained in terms
of the relative stability of different subshell electronic
configurations. This provides evidence for the electronic
configurations of subshells.
• Anomalies in the trends of ionisation energies can be understood
by considering the electronic configurations; there is a special
stability associated with half-filled and full subshells. The more
stable the electronic configuration, the higher the ionisation
energy.
• The Periodic Table can be subdivided into four blocks (s, p, d and
f) corresponding to the outer electronic configurations of the
elements within these blocks.
Trends in ionisation energy
The electrons in atoms and ions are attracted to the positive nucleus.
The process of removing electrons from atoms and ions is called
ionisation. The ionisation energy is the energy involved in
removing one mole of electrons from one mole of atoms in the
gaseous state. Looking a plot of the first ionisation energy against
atomic number (Figure 4.1), a regular repeating pattern is observed.
Figure 4.1 The first ionisation energy for the elements up to
barium
Figure 4.1 shows that ionisation energy is a periodic property. From
Higher we know that, in general, the ionisation energy increases
across a period. This is observed moving along period 2 from lithium
(Li) to neon (Ne) or along period 3 from sodium (Na) to argon (Ar);
there is an overall increase in energy. The explanation from Higher is
that the electrons are being removed from the same shell, and the
nuclear charge increases moving along the period (as more protons
are added) and so the outer electrons experience greater attraction
as the atom increases in atomic number.
However, on closer observation of the ionisation energies, they do
not show a perfect increase.
Small decreases are seen in the second period from beryllium (Be)
to boron (B) and then from nitrogen (N) to oxygen (O). Decreases
are also seen in the third period from magnesium (Mg) to aluminium
(Al) and then again from phosphorus (P) to sulfur (S).
The explanation for the decreases from group 2 to group 3 in in
period 2 (beryllium (Be) to boron (B)) and period 3 (magnesium (Mg)
to aluminium (Al)) is as follows: the added electron is in the new p
subshell which has a slightly higher energy level and is slightly
further from the nucleus. The s2 electrons also provide some
shielding for the group 3 elements. So despite the increased nuclear
charge, the ionisation energies are seen to decrease at this point.
Figure 4.2 Orbital box notation for boron and aluminium to help
explain the anomalies in ionisation energy
The explanation for the decreases from group 5 to group 6 in period
2 (nitrogen (N) to oxygen (O)) and period 3 (phosphorus (P) to sulfur
(S)) is due to the stability of the p orbitals. During ionisation, the
electrons for the group 6 elements are to be removed from the p4
configuration. There are four electrons in the outer p subshell, so two
must have paired opposite spins in the orbital.
Figure 4.3 Orbital box notation for oxygen and sulfur to help
explain the anomalies in ionisation energy
This fourth electron in the 3p orbital leads to electron–electron
repulsion. This lowers the attraction between that fourth electron and
the nucleus. So despite the increased nuclear charge, the electron is
easier to remove, therefore less energy is required.
These examples provide evidence that different electronic
configurations have relative stability associated with either a filled
subshell or a half-filled subshell. The p subshell is most stable when
it contains either three or six electrons. The d subshell is at its most
stable when it contains five or ten electrons. The more stable the
electronic configuration present, the harder it is to remove the outer
electron from its attraction to the nucleus and, subsequently, the
higher the ionisation energy.
Hints & tips
Half-filled and filled p or d subshells provide stability to an element,
and therefore such elements tend to have higher ionisation energy.
Four blocks of the Periodic
Table
The Periodic Table can be divided into four blocks, depending on
which subshell the outer electrons are located in.
• The electronic configuration for a group 1 sodium atom is 1s2 2s2
2p6 3s1 and for a group 2 calcium atom is 1s2 2s2 2p6 3s2 3p6 4s2.
The highest-energy-level electrons are always in an s subshell. So
these two groups form the s block.
• Groups 3 to 7 are referred to as the p block, as the highest-
energy-level electrons are found in the p subshell.
• The transition metals are described as the d block as the highest-
energy-level electrons are found in the d subshell. The transition
metals start at period 4. Consider manganese with an electronic
configuration of 1s2 2s2 2p6 3s2 3p6 3d5 4s2. It is known that for the
transition metals in the fourth period, the 4s subshell is filled before
the 3d subshell. When forming ions, the 4s subshell is emptied
before the 3d subshell.
• The lanthanide and actinide series have their highest-energy-level
electrons in the f subshell; they are called the f block. They fit
between the s-block and d-block elements in periods 6 and 7. The
lanthanides and actinides are all metal elements. The lanthanides
are found naturally on Earth, and promethium (Pm) is the only
radioactive member. The actinide series are all radioactive and
some are not found naturally. Plutonium was the first identified
actinide.
Figure 4.4 The Periodic Table can be divided into four different
blocks
Hints & tips
An element is classified as s, p, d or f block based on the location of
the highest-energy-level electrons – in the s, p, d or f subshells.
Study questions
1 Write the electronic configurations for atoms of oxygen, sodium
and manganese.
[3 marks]
2 State the block of the Periodic Table to which the following
elements belong:
a) tungsten
b) curium
c) strontium.
[1 mark]
3 The electronic configuration of an element is 1s2 2s2 2p6 3s2 3p6
3d10 4s2 4p6 4d10 5s2 5p5.
a) In which block of the Periodic Table is the element located?
[1 mark]
b) State the atomic number of the element and identify it by
name.
[1 mark]
4 Explain why the first ionisation energy of boron is lower than
expected.
[1 mark]
Chapter 5
Shapes of molecules and polyatomic
ions
Key points
• A dative covalent bond is one in which one atom of the bond
provides both electrons of the bonding pair.
• The shapes of molecules or polyatomic ions can be predicted
from the number of bonding electron pairs and the number of
non-bonding electron pairs (lone pairs).
• Pairs of electrons are negatively charged and so repel each
other. They are arranged in such a way that they are as far apart
as possible so repulsion is minimised.
• VSEPR theory is used to predict the shapes of molecules and
polyatomic ions.
• The arrangement of electron pairs is linear, trigonal, tetrahedral,
trigonal bipyramidal and octahedral when the total number of
electron pairs is 2, 3, 4, 5 and 6, respectively.
• Electron-pair repulsions decrease in strength in the order: lone
pair/lone pair > lone pair/bonding pair > bonding pair/bonding
pair.
• These different strengths of electron-pair repulsions account for
slight deviations from expected bond angles in molecules such as
NH3 and H2O.
Dative covalent bonds
A covalent bond is formed when two atomic orbitals combine
together to form a molecular orbital. This normally occurs when
both orbitals are half-filled before they overlap to form the molecular
orbital. A dative covalent bond is formed when one atom provides
both the electrons that form the bond. Once it has been formed, the
dative covalent bond is identical to other covalent bonds. An
example of this type of bond is found in the ammonium ion (Figure
5.1). A hydrogen ion in solution has no electrons and so cannot
share an electron with the bond. As a result, both the electrons in the
ammonium ion come from the lone pair on the nitrogen atom in the
ammonia molecule. Carbon monoxide and ozone, O3, are other
examples of molecules that contain a dative covalent bond.
Figure 5.1 Formation of an ammonium ion
The shapes of molecules
There is no direct relationship between the formula of a compound
and the shape of its molecules. The shape of a covalent ion or
molecule is dependent upon the repulsion between the electrons
around a central atom. This is called valence shell electron pair
repulsions (VSEPR). It is important not to confuse the number of
atoms with the number of electron pairs.
At first sight, CF4 and XeF4 may appear likely to be the same shape.
To determine whether they are, we need to work out the number of
electron pairs using the following formula:
number of electron pairs = (electrons of central atom + number
of atoms attached) ÷ 2
• For CF4, number of electron pairs = (4 electrons of central atom +
4 atoms attached) ÷ 2 = 4 electron pairs
• For XeF4, number of electron pairs = (8 electrons of central atom +
4 atoms attached) ÷ 2 = 6 electron pairs
The two molecules will, therefore, be different in shape as given by
Table 5.1.
Table 5.1 Working out the arrangement of electron pairs from the number of
electron pairs
Total number of electron pairs Arrangement of electron pairs
2 Linear
3 Trigonal
4 Tetrahedral
5 Trigonal bipyramidal
6 Octahedral
For an ion, we determine the number of electron pairs by the
following method: we subtract an electron from the central atom
value if the charge is +1 and add an electron to the central atom
value if the charge is −1.
• For BrF4−, number of electron pairs = ((7 electrons of central atom
+ 1 charge electron) + 4 atoms attached) ÷ 2 = 6 electron pairs
• For NH4+, number of electron pairs = ((5 electrons of central atom
− 1 charge electron) + 4 atoms attached) ÷ 2 = 4 electron pairs
Electron pairs are charge clouds around an atom and they need to
repel each other as far as is possible. Electrons can either be in a
bonding pair or a lone (non-bonding) pair of electrons. The shape of
an ion or molecule is determined from the number of electron pairs
around the central atom: the number of bonding pairs of electrons
and the number of lone pairs of electrons. Lone pairs are held closer
to the central atom, so they have a greater repulsion than the
bonding pairs of electrons.
Figure 5.2 Order of strength of repulsions
So lone pairs of electrons repel lone pairs of electrons more than
they repel bonding pairs of electrons. The lowest level of repulsion
occurs between bonding pairs of electrons. An ion or molecule tries
to minimise these repulsions in the shape it adopts.
When working out the shape of a molecule or ion, you must work out
the number of bonding pairs and lone pairs of electrons.
Hints & tips
Remember: repulsion between non-bonded electrons (lone pairs) is
greater than that between bonding pairs of electrons.
The arrangement of the electrons in covalently bonded molecules
can be shown in electron dot diagrams. For example, in hydrogen:
If we wish to show that each of the bonding electrons is from a
different hydrogen atom, the dot-and-cross variation can be used:
Further examples of dot-and-cross diagrams can be seen in the
following examples of the different shapes.
Two filled orbitals, both bonding pairs, are found in beryllium chloride
(BeCl2, Figure 5.3). The beryllium atom has two bonding pairs of
electrons. These bonding pairs repel each other equally, so the
molecule takes up a linear shape to minimise this effect. The bond
angle between the two bonds is 180°.
Figure 5.3 The linear shape of beryllium chloride
Three filled orbitals, all bonding pairs, are found in boron trifluoride
(BF3, Figure 5.4). There are three bonding pairs of electrons around
the central boron atom in BF3. These bonding pairs repel each other
equally and so take up a trigonal planar shape with a bond angle of
120°.
Figure 5.4 The trigonal planar shape of boron trifluoride
Four filled orbitals, all bonding pairs, are found in methane (CH4,
Figure 5.5). There are four bonding pairs of electrons around the
central carbon atom in CH4. These bonding pairs repel each other
equally and so take up a tetrahedral shape with a bond angle of
109.5°.
Figure 5.5 The tetrahedral shape of methane
Five filled orbitals, all bonding pairs, are found in phosphorus
pentafluoride (PF5, Figure 5.6).
Figure 5.6 The trigonal bipyramidal shape of phosphorus
pentafluoride
There are five bonding pairs of electrons around the central
phosphorus atom. These bonding pairs of electrons repel each other,
giving two bond angles of 90° and 120°; this results in a trigonal
bipyramidal shape.
Six filled orbitals, all bonding pairs, are found in sulfur hexafluoride
(SF6, Figure 5.7). There are six bonding pairs of electrons around
the central sulfur atom. These bonding pairs of electrons repel each
other equally giving an octahedral shape and a bond angle of 90°.
Figure 5.7 The octahedral shape of sulfur hexafluoride
Four filled orbitals – three bonding pairs and one lone pair – are
found in ammonia (NH3, Figure 5.8). There are three bonding pairs
of electrons and one lone pair of electrons around the central
nitrogen atom in ammonia. The arrangement of the four electron
pairs is tetrahedral but since there are only three bonds, the shape is
said to be pyramidal. There is greater repulsion between the lone
pair and the three bonding pairs than there is between the three
different bonding pairs, with the result that the bonds are pushed
closer together by the lone pair. So instead of a bond angle of 109.5°
found in a tetrahedral shape, the three bonds are angled closer
together at 107° to each other.
Figure 5.8 The pyramidal shape of ammonia
Four filled orbitals – two bonding pairs and two lone pairs – are
found in water (H2O, Figure 5.9). The basic arrangement of the
electron pairs is tetrahedral around the oxygen atom but as there is
no atom attached to the lone pairs, all you see is the two bonds.
Since there is greater repulsion between the two lone pairs than
between the lone pairs and the two bonding pairs, the outcome is
that the bonds are pushed even closer together in water than in
ammonia. In water the bond angle is 104.5° and the shape of the
water molecule is described as angular.
Figure 5.9 The angular shape of water
Hints & tips
You should know the bond angles for tetrahedral, pyramidal and
angular shapes.
Five filled orbitals – three bonding pairs and two lone pairs – are
found in bromine trifluoride (BrF3, Figure 5.10). In bromine trifluoride,
the central bromine atom has seven outer electrons and each
fluorine atom contributes one electron to make five electron pairs in
total. There will be three bonding pairs and two lone pairs. The five
electron pairs will be in a trigonal bipyramidal arrangement, but the
actual shape of the molecule depends on the arrangement of the
bonds. As there is greater repulsion between the two lone pairs than
between the lone pairs and the three bonding pairs, the outcome is
that the bonds are pushed even closer together. The lone pairs take
up positions of 120° from each other. The bonding fluorine atoms
take up the other positions in the trigonal bipyramidal shape but, as
the repulsion is greater with the lone pairs than the bonding pairs,
they have a bond angle of 86°. This molecule is described as T-
shaped.
Figure 5.10 The T-shape of bromine trifluoride
Hints & tips
Each extra lone pair causes the expected bond angle in a
tetrahedral molecule to decrease by 2.5°. So two lone pairs of
electrons cause a decrease of approximately 5° from 109.5°. As a
result, water and hydrogen sulfide have bond angles of 104.5°.
Example
1 Deduce the shape and bond angles of the PF6− ion.
Answer
Phosphorus is in group 5 so its atoms have five outer electrons.
There are six fluorine atoms attached. The negative charge on
the ion adds another electron.
So the number of electron pairs = ((5 electrons of central atom +
1 charge electron) + 6 atoms attached) ÷ 2 = 6 electron pairs
There are no lone pair of electrons so the shape will be
octahedral and the bond angles will be 90°.
2 Deduce the shape and bond angles of the hydrogen sulfide
molecule (H2S).
Answer
Sulfur is in group 6 so has six outer electrons on the central
atom. There are two hydrogen atoms attached to the sulfur.
There is no charge.
So the number of electron pairs = (6 electrons of central atom + 2
atoms attached) ÷ 2 = 4 electron pairs.
This means there are four pairs of electrons: two lone pairs and
two bonding pairs. Lone pairs repel the bonding pairs more
strongly than bonding pairs repel each other so the angle is less
than the tetrahedral shape. The bond angles will be 104.5° and
the shape is described as angular.
Study questions
1 Draw the following species showing their shapes:
a) PCl3
b) BrF4−
c) NH4+
[3 marks]
2 What is the shape and bond angle of the following molecules?
a) H2O
b) SiH4
c) BCl3
[3 marks]
Chapter 6
Transition metals, electronic
configuration and oxidation numbers
Key points
• The d-block transition metals are metals with an incomplete d
subshell in at least one of their ions. The filling of the d orbitals
follows the Aufbau principle, with the exception of chromium and
copper atoms. These exceptions are due to a special stability
associated with all the d orbitals being half-filled or completely
filled.
• When transition metals form ions it is the 4s-subshell electrons
which are lost first rather than the 3d-subshell electrons.
• An element is said to be in a particular oxidation state when it has
a specific oxidation number. The oxidation number is determined
by following certain rules. Transition metals exhibit variable
oxidation states of differing stability.
• Compounds of the same transition metal but in different oxidation
states may have different colours, for example vanadium.
• Oxidation can be considered as an increase in oxidation number
and reduction can be considered as a decrease in oxidation
number. Compounds containing metals in high oxidation states
tend to be oxidising agents whereas compounds with metals in
low oxidation states are often reducing agents.
Properties and electronic
configurations of the transition
metals
The general properties of transition metals are that they:
• have atoms or ions with an incomplete d subshell
• can form complexes
• have variable oxidation states
• show catalytic ability
• form coloured ions.
Hints & tips
A transition metal is defined as a metal that, in at least one of its
stable ions, has a partly filled d subshell.
The first period of transition metals runs from scandium to copper.
Details of the electronic configurations for these metals are shown in
Table 6.1.
The electronic configuration can be shortened by using [Ar] to
represent the configuration of argon [1s2 2s2 2p6 3s2 3p6]; using this
short form, scandium can be written as [Ar] 3d1 4s2 and manganese
can be written as [Ar] 3d5 4s2.
Chromium and copper atoms have only one electron in their 4s
subshell, 3d5 4s1 and 3d10 4s1, respectively, so do not follow the
Aufbau principle. They have a stability conferred by their half-filled
and filled d subshells because it is a lower-energy arrangement than
filling the 4s subshell with two electrons.
When a transition metal atom forms an ion, the electrons are lost
from the 4s subshell first. This can be seen in Table 6.1. So a
titanium ion would have the shortened configuration of [Ar] 3d2 while
a nickel ion could be written as [Ar] 3d8.
Hints & tips
Remember that chromium and copper atoms have only one
electron in their 4s subshell: 3d 5 4s1 or 3d 10 4s1, respectively. All
other transition metals have two electrons in their 4s subshell; Ti
has 3d2 4s2.
Example
1 Write the electronic configuration of a chromium(III) ion.
Answer
The name tells us that the valency of the ion is 3 and as it is a
metal ion, it has a positive charge, so forms a Cr3+ ion. A
chromium atom has the electronic arrangement 1s2 2s2 2p6 3s2
3p6 3d5 4s1 (remember, it does not have a full 4s subshell).
Remove the electrons required to form the ion charge, starting
with the 4s subshell.
Cr3+ has the configuration 1s2 2s2 2p6 3s2 3p6 3d3.
Oxidation numbers
Many transition metals have variable oxidation states in their
compounds. In ionic compounds, the oxidation number is equal to
the charge on the ion. Iron can form either an iron(II) ion, Fe2+, or an
iron(III) ion, Fe3+; copper can be copper(I), Cu+, or copper(II), Cu2+.
Manganese has four common oxidation states: manganese(II),
Mn2+, manganese(IV), Mn4+, manganese(VI), Mn6+, and
manganese(VII), Mn7+. It is important to be able to determine the
varying oxidation state of these and other ions. It gives a numerical
value for the degree of oxidation or reduction of an element.
The two terms, oxidation state and oxidation number, are
interchangeable so an element is said to be in a particular oxidation
state when it has a specific oxidation number.
There are certain rules for assigning and using oxidation numbers.
1 The oxidation number in a free or uncombined element is zero.
Thus, metallic magnesium has an oxidation number of zero as
does Cl in chlorine gas, Cl2.
2 For ions consisting of single atoms, the oxidation number is the
same as the charge on the ion. For example, the oxidation
number of chlorine in Cl− is −1, for oxygen in O2− it is −2 and for
aluminium in Al3+ it is +3.
3 In most compounds, the oxidation number for hydrogen is +1 and
for oxygen it is −2. Notable exceptions are metallic hydrides (−1
for hydrogen) and peroxides (−1 for oxygen).
4 In its compounds, fluorine always has the oxidation number −1.
5 The algebraic sum of all the oxidation numbers in a molecule must
be equal to zero.
6 The algebraic sum of all the oxidation numbers in a polyatomic ion
must be equal to the charge on the ion. For example, in SO42−,
the sum of the oxidation numbers of the one sulfur atom and four
oxygen atoms must equal −2.
Examples
2 Determine the oxidation state of the transition metal in the
compound sodium chromate (Na2CrO4).
Answer
As sodium chromate is a compound, the overall charge is zero.
There are two sodium ions with a +1 charge, so this is +2. Then
there are four oxygen ions with charges of −2, so this is −8. We
can do a simple algebraic sum to find the oxidation state of
chromium, the transition metal in the compound:
0 = (2 × +1) + (4 × −2) + Cr
(2 − 8) + Cr = 0, so therefore Cr = +6.
3 What is the oxidation state of vanadium in VO3−?
Answer
VO3− is a ion, so the overall charge is −1. There are three oxygen
ions with charges of −2, so this is −6. So if we do this simple
algebraic sum:
−1 = (3 × −2) + V
(−6) + V = −1
we can find the oxidation state of V, which is +5.
Oxidation is the loss of electrons and reduction is gain of electrons.
So oxidation can be described as an increase in oxidation number.
Reduction can be said to be a decrease in oxidation number.
Examples
4 Work out the oxidation number of manganese in MnO4− and in
Mn2+ and decide whether the conversion of MnO4− to Mn2+ is an
example of oxidation or reduction.
Answer
MnO4− is a ion, so the overall charge is −1. There are four
oxygen ions with a charge of −2, so this is −8. So if we do this
simple algebraic sum:
−1 = (4 × −2) + Mn
(−8) + Mn = −1
we can see that in the MnO4− ion, the oxidation number of Mn is
+7.
Mn2+ has a valency which provides its oxidation state, which is
+2. Looking at the change in oxidation state for manganese from
+7 to +2, we can see that it is a reduction as the oxidation
number has decreased.
5 Work out the oxidation number of chromium in Cr2O72− and in
Cr3+ and decide whether the conversion of Cr2O72− to Cr3+ is an
example of oxidation or reduction.
Answer
Cr2O72− is a ion, so the overall charge is −2. There are seven
oxygen ions with a charge of −2, so this is −14. So if we do this
simple algebraic sum:
−2 = (7 × −2) + 2Cr
(−14) + 2Cr = −2
2Cr = +12
we can see that Cr = +6 in the Cr2O72− ion.
Cr3+ has a valency which provides its oxidation state, which is +3.
Looking at the change in oxidation state for chromium from +6 to
+3, we can see that it is a reduction as the oxidation number has
decreased.
Colour changes in transition
metals
Compounds containing metals in a high oxidation state tend to be
oxidising agents; for example:
MnO4− + 8H+ + 5e− → Mn2+ + 4H2O
Compounds containing metals in a low oxidation state tend to be
reducing agents; for example:
Fe(OH)2 + OH− → Fe(OH)3 + e−
Acidified permanganate and acidified dichromate are two very good
oxidising agents which are often used in redox experiments.
Permanganate (MnO4−) has manganese with an oxidation state of
+7. When it is used as an oxidising agent, the manganese sees a
reduction in oxidation state to +2. In dichromate, the chromium ion
has an oxidation state of +6 which then reduces to +3. These
examples also show a change in colour as their oxidation state
changes. They are good examples of self-indicating reagents.
Another example of a transition metal which varies in colour
depending on its oxidation state is vanadium (Table 6.2).
Hints & tips
A useful mnemonic for remembering the colours in vanadium is:
You Better Get Vanadium. (Y for yellow, B for blue, G for green and
V for violet.)
Study questions
1 Write the electronic configuration of the following transition metal
ions:
a) an iron(II) ion
b) a manganese(II) ion
c) a cobalt(II) ion.
[3 marks]
2 Work out the oxidation number of chromium in Cr2O72− and in
CrO42− and decide whether the conversion of Cr2O72− to CrO42−
is an example of oxidation or reduction.
[2 marks]
3 Work out the oxidation state of the transition metals in
a) Co2O3
b) [CoCl4]2−
c) [Ag[NH3]2]+.
[3 marks]
Chapter 7
Ligands and transition metal
complexes
Key points
• A complex consists of a central metal ion or atom surrounded by
ligands.
• Ligands are electron donors and may be negative ions or
molecules with non-bonding pairs of electrons. Ligands can be
classified as monodentate, bidentate, etc. up to hexadentate.
• The number of bonds from the ligand to the central metal ion or
atom is known as the co-ordination number of the central ion.
• Complexes are written and named according to IUPAC rules.
Transition metal complexes
An important property of transition metals is their ability to form
complexes, sometimes referred to as co-ordination compounds.
A complex has a metal ion or metal atom at its centre with a number
of other molecules or ions surrounding it. The molecules or ions
surrounding the central metal ion or atom are called ligands.
Ligands have at least one lone pair of electrons which can form a
co-ordinate bond (dative covalent bond) to the metal. The lone pair
of electrons on the ligand is donated into an empty orbital in the
transition metal atom or ion.
Ligands that form one co-ordinate bond to a metal atom or ion are
called monodentate ligands. Examples of monodentate ligands are
NH3, H2O, Cl−, OH− and CN−. When ligands form two co-ordinate
bonds, they are said to be bidentate ligands. Examples of this type
of ligand are 1,2-diaminoethane (H2NCH2CH2NH2) and oxalate ions
(C2O42−). The lone pairs of electrons are on different atoms, allowing
the ligand to form more than one co-ordinate bond.
Figure 7.1 Bidentate ligands showing the lone pairs of
electrons
A quadridentate ligand has four lone pairs, so forms four co-
ordinate bonds with the central metal ion or atom. An example of this
occurs in haemoglobin.
A hexadentate ligand has six lone pairs of electrons, all of which
can form co-ordinate bonds with the same metal ion. The best
example is ethylenediaminetetraacetic acid, which is abbreviated to
EDTA (Figure 7.2). The EDTA ion wraps up a metal ion entirely using
all six of its lone pairs of electrons.
Figure 7.2 EDTA showing the six lone pairs of electrons that
form co-ordinate bonds with the central metal ion, making EDTA
multidentate and hexadentate
Hints & tips
The co-ordination number is the total number of dative covalent
bonds to the metal atom or ion. It is not the number of ligands
attached.
The total number of co-ordinate bonds from the ligand to the central
metal ion is known as the co-ordination number.
Small ligands like water and ammonia normally form octahedral
complexes in which the co-ordination number of the complex is 6. In
the complex shown in Figure 7.3, the six cyanide ions act as ligands
and form an octahedral complex with iron(III) ions.
Figure 7.3 In this octahedral complex, the co-ordination number
is 6
The co-ordination number is not the same as the number of ligands.
The co-ordination number for an EDTA complex is 6 because of the
six co-ordinate bonds being formed by the central metal ion. This
can be seen as it forms a complex with iron (Figure 7.4).
Figure 7.4 Co-ordination number 6
Complexes are written with the metal atom or ion and the ligands
inside square brackets. The charge on the complex ion (if there is
one) is placed as a superscript outside the square bracket, for
example [CoCl4(NH3)2]−.
Ligands are referred to as Lewis bases since they donate electrons
to the central metal atom. The metals, in turn, are Lewis acids since
they accept electrons.
Naming complexes
When naming co-ordination complexes, the rules set out by the
International Union of Pure and Applied Chemistry (IUPAC) are
used.
To name a complex:
1 The ligands are named before the metal ion.
2 The names of the ligands are written in the order: neutral,
negative, positive. If a complex contains multiple ligands of the
same charge, they are named in alphabetical order, ignoring any
numerical prefixes.
3 When monodentate ligands occur more than once, a prefix gives
the number of occurrences: di-, tri-, tetra-, penta- or hexa-.
Polydentate ligands (such as EDTA) have the prefix bis-, tris-,
tetrakis-, etc.
4 The endings of anion names change. For example, cyanide
becomes cyanido, sulfate become sulfato and nitrite becomes
nitrito.
5 Neutral ligands are given their usual name, but there are a few
exceptions: H2O becomes aqua or aquo, NH3 becomes ammine,
and CO becomes carbonyl.
6 Write the name of the central metal atom or ion. If the complex is
an anion, the central atom’s name will end in ‘-ate’. If available, its
Latin name will be used (with the exception of mercury).
7 If the oxidation state of the metal atom or ion needs to be
specified, it is written as a Roman numeral in brackets (II, IV, etc.).
Hints & tips
Remember, the spelling of ammonia as a ligand is ammine.
Examples
1 What is the name of the complex ion [NiCl4]2−?
Answer
The ligand is a chloride ion which is changed to chlorido in the
complex name. The complex has four of them, so that is
tetrachlorido. Add the suffix ‘-ate’ in the name of the metal to
make it nickelate. The oxidation state for nickel in this complex is
determined to be 2+, so this gives us the Roman numeral (II).
So the complex is a tetrachloridonickelate(II) ion.
2 What is the name of the complex [Co(NH3)5(OH2)]3+?
Answer
There are two ligands: one is ammonia and the second is water.
Alphabetically, the ligand name of ammonia, ammine, comes
before that of water, aqua. The complex has five ammonia
ligands and one water ligand so it is pentaammineaqua. The
ligands are both neutral so the name for the metal is just cobalt.
So the complex name is pentaammineaquacobalt(III).
As given in step 6 of the rules, the Latin name for the metal is used if
the complex is an anion. Table 7.1 gives the Latin names of some of
the common transition metals found in complexes.
Table 7.1
Transition metal Latin name
Iron Ferrate
Copper Cuprate
Tin Stannate
Silver Argentate
Lead Plumbate
Hints & tips
In negative ion (anion) complexes, the Latin name is used for some
transition metals.
Study questions
1 What is the name of the complex [Fe(CN)6]2−?
[1 mark]
2 What is the name of the complex [Ni(H2O)6]2+?
[1 mark]
3 What is the formula for potassium hexacyanidoferrate(III)?
[1 mark]
4 What is the formula for tetrachloridocuprate(II)?
[1 mark]
Chapter 8
Colours of transition metals and
catalysts
Key points
• In a transition metal complex, the d orbitals are no longer
degenerate. Different ligands produce different field splittings and
so complexes of the same metal ion with different ligands will
have different colours.
• Ligands that cause a large difference in energy between
subshells of d orbitals are strong field ligands. Weak field ligands
cause a small energy difference.
• The energy difference between subshells of d orbitals depends
on the position of the ligand in the spectrochemical series.
Colours of many transition metal complexes can be explained in
terms of d–d transitions. The effects of d–d transitions can be
studied using ultraviolet and visible absorption spectroscopy.
• Light is absorbed when electrons in a lower-energy d orbital are
promoted to a d orbital of higher energy.
• Ultraviolet and visible absorption spectroscopy involve transitions
between electronic energy levels in atoms and molecules where
the energy difference corresponds to the ultraviolet and visible
regions of the electromagnetic spectrum.
• An ultraviolet/visible spectrometer measures the intensity of
radiation transmitted through the sample and compares this with
the intensity of incident radiation.
• Transition metals or their compounds act as catalysts in many
chemical reactions. Catalysts can be homogeneous (in the same
state as the reactants) or heterogeneous (in a different state to
the reactants).
• It is believed that the presence of unpaired d electrons or unfilled
d orbitals allows intermediate complexes to form, providing
reaction pathways of lower energy compared to the uncatalysed
reaction.
• The variability of oxidation states of transition metals is another
important factor.
Transition metals and colour
Simple ions and complex ions of the transition metals are often
coloured. This is because they absorb light in certain parts of the
visible spectrum. The colour seen is the complementary colour to
that absorbed; in other words, it is a combination of the colours not
absorbed. To understand this outcome, it has to be appreciated that
white light is a combination of the three primary colours: red, blue
and green.
• If red light is absorbed, the colours transmitted are blue and green,
which is seen as green/blue or cyan.
• If blue light is absorbed, the colours transmitted are red and green,
which is seen as yellow.
• If all the colours are absorbed, the complex will appear black.
This information is summarised in Table 8.1.
There are five d orbitals and all are degenerate (have the same
energy) when the transition metal is not bonded to anything else.
However, when ligands approach along the x-, y- and z-axes, the
lone pair of the ligand repels the d orbitals, especially those that are
pointing directly along the axes. The d subshell is split into two
distinct sets of orbitals that have a difference in energy between
them (ΔE, Figure 8.1).
Figure 8.1 Splitting of d orbitals
The size of the energy gap (ΔE) between the two sets of d orbitals
varies with:
• the transition metal ion
• the oxidation state of the transition metal
• the type of ligand.
Energy in the visible region of the spectrum causes d-electron
transitions between lower energy (ground state) orbitals and d
orbitals of higher energy (excited state). So some visible light
frequencies are absorbed and the complementary colour is
observed. Different ligands bonded to the central transition metal ion
cause different splits in the d orbitals. It is the size of this split which
determines the colour produced by the compound or complex. The
ability of ligands to cause splitting is given by the spectrochemical
series. For the most common ligands this is:
• The greater the splitting, the more energy is needed to promote an
electron from the lower group of d orbitals to the higher ones.
• The greater the energy of the light absorbed, the shorter the
wavelengths. As the splitting increases, the light absorbed shifts
away from the red end of the spectrum towards orange, yellow and
so on.
• The concentration of transition metals in solution can be
determined by their absorbance of visible or ultraviolet light.
• The higher the concentration of the ions, the greater the
absorbance.
• The absorption of ultraviolet or visible radiation corresponds to
outer electrons becoming excited.
Complexes containing weak field ligands, such as water, are more
likely to absorb in the visible region (400–700 nm). Colorimetry can
be used to determine the concentration of the ions. A filter of the
complementary colour of the solution being tested at a defined
wavelength is used in the colorimeter. Further details regarding
colorimetry are found in Unit 4 Researching chemistry.
If the energy absorbed is in the ultraviolet part of the electromagnetic
spectrum, the compound will be colourless and so ultraviolet
spectroscopy is used. When the ligands surrounding the transition
metal ion are strong-field ligands, such as the cyanide ion, the d–d
transition has a tendency to occur in the ultraviolet region (200–400
nm). Different wavelengths from 200 to 400 nm are passed through
the sample and the ultraviolet light absorbed at the different
wavelengths is recorded.
This theory only works if the transition metal in the compound has
between one and nine d electrons. Ions that have no d electrons or
ions that have a complete d subshell cannot have d–d transitions.
The colour wheel on page 20 of the SQA data booklet may help you
to work out which colours are transmitted when a specific colour or
wavelength is absorbed. If wavelengths of one or more colours are
absorbed, the colours on the opposite side of the colour wheel are
observed.
Hints & tips
Changes in the metal, the metal oxidation state, the ligand type and
the co-ordination number all have an effect on energy difference in
d orbitals. This changes the light absorbed and so the visible colour
observed.
Example
1 Copper(II) ions form a light blue solution in water and form a
[Cu(OH2)6]2+ complex. When copper(II) ions react with excess
ammonia, the complex [Cu(NH3)4(OH2)]2+ is formed, which has a
very deep blue colour. Explain why both of the complexes are
coloured. Why does changing the ligand from water to ammonia
result in a change in colour?
Answer
An energy gap is created between the d orbitals so they are no
longer degenerate. This energy gap corresponds to a frequency
of light which is absorbed. The transmitted light is complementary
to that absorbed. Changing the ligands alters the energy gap
between the d orbitals. This leads to a different frequency of light
being absorbed.
Catalysts
Transition metals and their compounds are important catalysts in
many industrial chemical reactions, as detailed in Table 8.2.
Table 8.2 Transition metals and the reactions they catalyse
Process Catalyst used
Haber process Iron
Contact process Vanadium(V) oxide
Ostwald process Platinum gauze
Catalytic converter in cars Platinum, palladium, rhodium
Preparation of methanol Copper
Preparation of margarine Nickel
Polymerisation of alkenes Titanium compounds
Catalysts speed up chemical reactions by providing an alternative
reaction pathway of lower activation energy.
• A homogeneous catalyst is in the same state as the reactants.
• A heterogeneous catalyst is a catalyst in a different state to the
reactants. Most industrial processes are examples of
heterogenous catalysis.
Heterogeneous catalysts work through a process called
chemisorption. As transition metals can form a variable number of
bonds because of the availability of unoccupied or partially occupied
d orbitals, it is believed that the presence of unpaired electrons or
unfilled d orbitals allows intermediate complexes to form. Reactant
molecules are therefore adsorbed onto active sites on the surface of
the catalyst. The bonds in the reactant molecules are weakened.
The reactant molecules are held in a more favourable conformation
for the reaction and form an intermediate complex before forming the
product. The product molecules that have formed are then desorbed
from the surface of the catalyst.
The ability of the transition metals to have a variety of oxidation
states enables them to act as homogeneous catalysts. The transition
metal can provide an alternative reaction pathway with lower
activation energy and so speed up the reaction. The transition metal
returns to its original oxidation state once the reaction is complete.
A good example of homogeneous catalysis is the oxidation of
Rochelle salt (potassium sodium tartrate) by hydrogen peroxide. The
reaction progresses very slowly without a catalyst, with only a few
bubbles of gas being released. Almost immediately on the addition of
pink cobalt(II) chloride solution, vigorous effervescence is observed.
A colour change is also seen as the cobalt changes from pink to
green. When the reaction finishes, the effervescence stops and the
pink colour returns. In terms of oxidation states, the cobalt has
changed from +2 (pink) to +3 (green) and then back to +2 (pink) at
the end:
Hints & tips
Transition metals act as catalysts because due to them having
variable oxidation states, they have d orbitals available to form
intermediate complexes on the catalyst active sites.
A catalyst can become poisoned if compounds adsorb irreversibly
onto the active sites. This decreases or prevents the activity of the
catalyst.
Example
2 The transition metals vanadium and copper can have variable
oxidation states and have a wide range of uses. Vanadium(III)
ions can react with iron(III) ions in solution. The reaction is
catalysed by Cu2+ ions:
V3+(aq) + Fe3+(aq) → V4+(aq) + Fe2+(aq)
Explain why Cu2+(aq) can be classified as a homogeneous
catalyst in this reaction.
Answer
The two reactants are in solution and so are the Cu2+ ions. A
homogeneous catalyst is in the same state as the reactants.
Study questions
1 Which of the following correctly places the ligands in their order
in the spectrochemical series?
[1 mark]
A I− < Br− < H2O < F−
B Br− < Cl− < NH3 < H2O
C F− < Cl− < H2O < NH3
D I− < Cl− < H2O < CN−
2 Which one of the following hexa-aqua complexes is unlikely to be
coloured?
[1 mark]
A Cr3+
B Co2+
C Cu2+
D Al3+
3 When a transition metal complex is formed, the presence of
ligands ‘splits’ the d orbitals. The ability of a ligand to split the d
orbitals when forming a complex with a transition metal ion is
determined by the spectrochemical series. Three ligands from
the series, listed in order of their relative ability to split the d
orbitals, are NH3 > H2O > Cl−. A study of part of the absorption
spectrum for the complex ion formed between nickel(II) and six
chloride ligands shows a broad absorption band which peaks at
around 440 nm.
Explain what would happen to the peak absorption wavelength if
the chloride ligands were replaced by water ligands.
[2 marks]
Unit 1 Glossary
Absorption spectrum: A spectrum consisting of dark absorption
lines superimposed on a bright continuous spectrum; it shows the
absorption of radiation by a material over a range of wavelengths.
Aufbau principle: This states that orbitals are filled in order of
increasing energy.
Bidentate: A ligand that contains two atoms with lone pairs of
electrons capable of bonding to a metal atom or ion.
Complex: A complex consists of a central metal atom or ion
surrounded by ligands.
Co-ordinate bond: A covalent bond in which one of the atoms
supplies both of the electrons of the shared pair.
Co-ordination compounds: Compounds in which a central metal
atom or ion is attached to a group of surrounding molecules or ions
by dative covalent bonds (also known as co-ordinate bonds).
Co-ordination number: The co-ordination number is the number of
nearest neighbours by which an atom or ion is surrounded in a
structure.
Dative covalent bond: In this type of covalent bond, both the
shared electrons originally came from the same atom.
Degenerate: A set of atomic orbitals that are of equal energy to
each other are said to be degenerate.
Electromagnetic spectrum: This is the range of frequencies or
wavelengths of electromagnetic radiation.
Electronegativity: Electronegativity is a measure of the attraction
an atom involved in a bond has for the electrons of the bond.
Emission spectroscopy: The study of emission spectra produced
by excited substances (often gaseous atoms or molecules).
Excitation energy: The minimum energy required to change a
system from its ground state to a particular excited state.
Frequency: This is the number of wavelengths that pass a fixed
point in one unit of time.
Ground state: This is the lowest possible electronic configuration
the electrons in an atom can adopt.
Heisenberg’s uncertainty principle: This states that it is impossible
to state precisely the position and the momentum of an electron at
the same instant.
Hexadentate: A ligand that bonds to a metal ion using electron pairs
on six donor atoms.
Hund’s rule: When degenerate orbitals are available, electrons fill
each singly, keeping their spins parallel, before pairing starts.
Ionisation: The addition or removal of an electron to create an ion.
Ionisation energy: The energy involved in removing one mole of
electrons from one mole of atoms in the gaseous state.
Ligands: Molecules or ions that bond to the central metal atom or
ion in a complex.
Lone pair: A lone or non-bonding pair of electrons is a pair of outer
or valence shell electrons (that have opposing spins) which are not
used to form covalent bonds within the molecule.
Molecular orbital: A molecular orbital is a region in space between
the nuclei where there is a high probability of finding electrons. It is
formed by the overlap of atomic orbitals.
Monodentate: A ligand that bonds to a metal atom or ion using the
electron pair of a single donor atom.
Oxidation: This is the loss of electrons from a substance. It can also
be described as an increase in oxidation number.
Oxidation number: The formal charge assigned to each atom in a
compound according to certain rules.
Pauli exclusion principle: This states that an orbital holds a
maximum of two electrons.
Quanta: The smallest possible discrete unit of any physical property,
such as energy or matter.
Reduction: The gain of electrons by a substance. It can also be
described as a decrease in oxidation number.
Spectrochemical series: A list of ligands in order of the size of the
crystal field splitting caused in the d orbitals.
Wavelength: The distance between adjacent crests or troughs of a
wave.
Wavenumber: Wavenumber is the reciprocal of wavelength and has
the units of cm−1 (number of cycles per cm).
Unit 1 Exam-style questions
1 The ore pyrolusite, MnO2, was used 30 000 years ago as a black
pigment in the cave paintings of Lascaux, France.
The best known oxide of manganese is possibly potassium
permanganate, KMnO4, first made in 1740 for the glass industry. It
now has many uses, including disinfectants and the removal of
organic impurities from waste gases and effluent water.
a) (i) State the oxidation number of manganese in MnO2.
[1 mark]
(ii) Using orbital box notation, write the electronic
configuration for a manganese ion in MnO2.
[1 mark]
(iii) Explain how your answer is consistent with Hund’s rule.
[1 mark]
b) (i) The d orbitals in an isolated manganese atom are
degenerate. State the meaning of the term degenerate.
[1 mark]
(ii) The second quantum number, l, is related to the shape of
the orbitals. Draw the shape of an orbital when l = 1.
[1 mark]
2 Which one of the following metal salts will emit radiation of the
highest frequency when placed in a Bunsen flame?
[1 mark]
A Copper(II) chloride
B Potassium chloride
C Barium chloride
D Lithium chloride
3 An atom has electronic configuration 1s2 2s2 2p6 3s2 3p3. When
ionised, this is likely to form a particle with a charge of
[1 mark]
A 5+
B 5−
C 3+
D 3−
4 Atomic spectroscopy is a useful analytical tool for identifying and
quantifying the elements present in a sample. It also provides
information about atomic structure.
a) When a high voltage is applied to a lamp filled with helium gas,
a line of red light, wavelength 706 nm, is observed through a
spectroscope.
(i) Explain how the line of red light is produced.
[2 marks]
(ii) Calculate the energy, in kJ mol−1, associated with this
wavelength.
[2 marks]
b) A helium atom has two electrons in its ground state. One of the
electrons can be described by the four quantum numbers
What four quantum numbers describe the other electron?
[1 mark]
c) (i) Using orbital box notation, write the electronic configuration
for a phosphorus atom in its ground state.
[1 mark]
(ii) Explain how your answer is consistent with Hund’s rule.
[1 mark]
5 Which of the following statements for the first row transition metals
is always true?
[1 mark]
A 2+ ions are formed by removing electrons from 3d orbitals.
B The 3d orbitals remain degenerate when ligands bond to the
metal.
C All electronic configurations obey the Aufbau principle.
D They have at least one electron in the 4s orbital.
Unit 2 Physical chemistry
Chapter 9
Chemical equilibrium and factors
affecting equilibrium
Key points
• A reaction is in equilibrium when the concentrations of reactants
and products are constant.
• The equilibrium constant has the symbol K and a general
expression of reactants ⇋ products is given by the formula:
• The value of an equilibrium constant indicates the position of
equilibrium:
• K less than 1 indicates that equilibrium lies more to the left
(reactants) side.
• K approximately 1 indicates that neither reactants nor products
are favoured.
• K greater than 1 indicates that equilibrium lies more to the right
(products) side.
• The numerical value of the equilibrium constant depends on the
reaction temperature and is independent of concentration and/or
pressure.
• For endothermic reactions, a rise in temperature causes an
increase in K and the yield of the product is increased. For
exothermic reactions, a rise in temperature causes a decrease in
K and the yield of the product is decreased.
• The presence of a catalyst does not affect the value of the
equilibrium constant.
Chemical equilibrium
In Higher Chemistry we learnt that some reactions never go to
completion but are in a state of dynamic equilibrium. For an
equilibrium to establish, the reaction must take place in a closed
system. This is where the chemicals are trapped, and nothing is
added or removed, but heat is allowed to be transferred to or from
the surroundings.
A system is in equilibrium when the concentrations of reactants and
products are constant, so the forward reaction (reactants to
products) and the reverse reaction (products to reactants) do not
stop. At equilibrium, the rate of the forward reaction is equal to the
rates of the reverse reaction. In other words, reactants are still
converting to products and products are still converting to reactants,
but the rates of change are equal. Overall there is no change in the
composition of the equilibrium mixture.
According to Henri Le Chatelier,
‘If a system is at equilibrium and a change is made in any of the
conditions, then the system responds to counteract the change as
much as possible.’
In other words, if you make a change to a reaction at equilibrium, the
system will alter the concentrations to minimise the effect until
equilibrium is re-established. From Higher we know that the position
of equilibrium can be altered by changing the concentrations of
reacting substances, the pressure of reacting gases and the
temperature.
A catalyst will speed up the rate of both the forward and the reverse
reactions but has no effect on the position of equilibrium.
Table 9.1 summarises the effect on equilibrium position of changes
made to the system.
Table 9.1
Change applied Effect on equilibrium position
Increasing the concentration of Equilibrium shifts to the right
reactant or removing the product
Increasing the concentration of Equilibrium shifts to the left
product or removing the reactant
Increasing the temperature Equilibrium shifts in the direction
of the endothermic reaction
Decreasing the temperature Equilibrium shifts in the direction
of the exothermic reaction
Catalyst No effect on equilibrium position;
equilibrium more rapidly
obtained
The equilibrium constant
The equilibrium constant is given the symbol K and can be described
by the general expression
for the reversible reaction reactants ⇋ products.
If the reaction is then written as
aA + bB → cC + dD
the equilibrium expression can be written:
where the square brackets [ ] indicate the concentration of the
species present at equilibrium; this is then raised to the power of the
number of moles of that species from the balanced equation. The
concentrations are usually measured in mol l−1, but for gaseous
reactions, partial pressures can be used. Remember, though, that
the equilibrium expression represents a ratio and so K has no units.
Hints & tips
Remember that equilibrium concentrations of reactants and
products are hardly ever the same.
Example
In a reaction, 1 mole of compound A and 1 mole of compound B
were reacted in a closed 1 litre container. At equilibrium, the mixture
was found to contain 0.3 moles of product C along with unreacted A
and B.
A + 2B ⇋ 2C
Calculate the equilibrium constant for this reaction.
Answer
Step 1: Work out the moles/concentrations of each reactant and
product once equilibrium has been established.
The equilibrium mixture is made up of 0.85 mol A + 0.7 mol B
(reactants) + 0.3 mol C (product).
Step 2: Write an expression for the equilibrium constant, K.
Step 3: Using the equilibrium concentrations, calculate the
equilibrium constant, K.
0.85 mol A and 0.7 mol B remain along with 0.3 mol of C produced,
all in 1 litre. So
In a homogeneous equilibrium, all the species are in the same
gaseous phase. For example:
N2(g) + 3H2(g) ⇋ 2NH3(g)
The equilibrium expression for this reaction is:
In a heterogeneous equilibrium, the species are in more than one
phase. For example:
Br2(l) ⇋ Br2(g)
The concentrations of pure solids and pure liquids acting as solvents
for the reaction are constant and as such are given the value of 1 in
the equilibrium equation. An expression for the equilibrium constant
for the above reaction would be
If we look at the equilibrium produced when carbon is heated with
steam:
H2O(g) + C(s) ⇋ H2(g) + CO(g)
everything is exactly the same as before in the equilibrium constant
expression, except that you leave out the solid carbon. So the
equilibrium expression for this reaction is
Equilibrium constants are independent of changes of concentrations
or partial pressures of species in any given reaction. When the
concentration or partial pressure is changed in a reaction, the
position of equilibrium also changes until the ratio of the products to
reactants is re-established to its original value. The position of
equilibrium is used to describe how an equilibrium solution behaves
when reaction concentrations change. So while the numerical value
of the equilibrium constant K remains the same, the position of
equilibrium is all that is adjusted.
Equilibrium constants state where the thermodynamically most
favourable ratio of reactants to products is. So a high value of K
means that the fraction used to calculate it has more product, i.e.
equilibrium lies to the right. A low value of K means the equilibrium
lies more to the left. Table 9.2 summarises this for different reactions.
Knowing the value of the equilibrium constant, K, allows us to
determine:
• the direction in which a reaction will proceed to achieve equilibrium
• the ratio of the concentrations of reactants and products when
equilibrium is reached.
Exothermic and endothermic
reactions
The value of the equilibrium constant, K, is dependent upon
temperature. Consider the following equilibrium:
ΔH for the forward reaction is positive so the reaction is endothermic.
Increasing the temperature favours the endothermic reaction and
decreasing the temperature will favour the exothermic direction:
• At higher temperatures, products are favoured so K will increase.
• At lower temperatures, reactants are favoured so K will decrease.
For exothermic reactions, a rise in temperature causes a decrease in
K and the yield of the product is decreased.
Catalysts
Catalysts speed up the rates of the forward and reverse reactions by
an alternative reaction pathway with lower activation energy. The
position of equilibrium is not changed, and so there is no effect on
the equilibrium constant. In summary, catalysts have no effect on
equilibrium constant or equilibrium position and only allow
equilibrium position to be achieved at a faster rate.
Hints & tips
Catalysts have no effect on the equilibrium constant, K.
Study questions
1 2 moles of A and 1 mole of B were placed into a flask and
equilibrium was established. At equilibrium, 0.5 moles of D were
formed.
2A + B ⇋ 2C + D
Determine the composition of the equilibrium mixture.
[2 marks]
2 When 2 moles of phosphorus pentachloride were heated to 523
K in a closed vessel, 25% of it dissociated.
PCl5(g) ⇋ PCl3(g) + Cl2(g)
How many moles of gas would be present in the equilibrium
mixture?
A 1.0 moles
B 3.0 moles
C 2.5 moles
D 3.5 moles
[1 mark]
3 If the equilibrium concentrations found in the reaction
A(g) + 2B(g) ⇋ 2C(g)
are [A] = 0.025 M, [B] = 0.15 M and [C] = 0.25 M, calculate the
value of K.
[1 mark]
4 Nitrogen monoxide decomposes into nitrogen and oxygen
according to the equilibrium
2NO(g) ⇋ N2(g) + O2(g)
Write an expression for the equilibrium constant K for this
reaction.
[1 mark]
5 If 2.5 mol of nitrogen monoxide is left in a sealed container at
room temperature, 0.75 mol of oxygen is found to be present at
equilibrium.
a) Calculate the amounts, in moles, of nitrogen monoxide and
nitrogen present in the equilibrium mixture.
[2 marks]
b) Calculate a value for K for this equilibrium at this temperature.
[2 marks]
Chapter 10
Chemical equilibrium and the ionic
product of water
Key points
• The Brønsted–Lowry definitions of acids and bases state that an
acid is a proton donor and a base is a proton acceptor.
• For every acid there is a conjugate base formed by the loss of a
proton, and for every base there is a conjugate acid formed by
the gain of a proton.
• In water and aqueous solutions there is an equilibrium between
the water molecules and hydronium (hydrogen) and hydroxide
ions. This can be represented by:
H2O(l) + H2O(l) ⇋ H3O+(aq) + OH−(aq)
• Water is amphoteric and can act as an acid or a base.
• The dissociation constant for the ionisation of water is known as
the ionic product and is represented by Kw.
• The value of Kw varies with temperature and at 25 °C it is
approximately 1 × 10−14.
• The relationship between pH and the hydrogen ion concentration
is given by:
pH = −log10[H3O+] and [H3O+] = 10−pH
• In water and aqueous solutions with a pH value of 7, the
concentrations of H3O+ and OH− are both 10−7 mol l−1 at 25 °C.
• If the concentration of H3O+(aq) or OH−(aq) is known, the
concentration of the other ion can be calculated using Kw or by
using pH + pOH = 14.
The Brønsted–Lowry
definitions of acids and bases
From National 5 Chemistry you understand that acids (pH < 7) have
a higher concentration of hydrogen ions (H+) than pure water and
that alkalis (pH > 7) have a higher concentration of hydroxide ions
(OH−) than pure water.
For Advanced Higher, consider the hydrogen ion as simply a proton
(a hydrogen atom that has lost an electron). In an aqueous solution,
a hydrogen ion only exists when surrounded by water molecules.
These are known as hydronium ions and are written as H3O+(aq).
When the hydrogen ion (H+) concentration is discussed, what is
really being referred to is the hydronium ion, but for simplification, it
is generally shortened to H+(aq).
So, in a theory put forward by Brønsted and Lowry in 1932, an acid
can be considered as a proton donor and a base as a proton
acceptor. When an acid donates a proton, the species left is called
the conjugate base. When a base accepts a proton, the species
formed is a conjugate acid.
Hints & tips
A shorthand representation of H3O+ is H+. Stoichiometric equations
and equilibrium expressions can be written using H+ instead of
H3O+ where the meaning is clear.
In the reaction between hydroiodic acid (HI) and ethanoic acid
(CH3COOH):
it can be seen that HI has donated a proton to CH3COOH, so the HI
is an acid and the CH3COOH is acting as a base. The products are a
conjugate acid (CH3COOH2+) and a conjugate base (I−).
Water as an acid and a base
Water dissociates very slightly according to this equilibrium equation:
In pure water, the concentrations of hydroxide and hydronium ions
are equal and the pH is neutral or 7. So, water can be called
amphoteric, as it can behave as either a base or an acid.
In water and aqueous solutions, there is an equilibrium between the
water molecules and hydronium (hydrogen) and hydroxide ions.
Water acts both as a proton donor (acid) forming a conjugate base
and as a proton acceptor (base) forming a conjugate acid.
For simplicity we write the above equation as
H2O(l) ⇋ H+(aq) + OH−(aq) ∆H +ve
The dissociation constant for the ionisation of water is known as the
ionic product. It is represented by Kw:
Kw = [H3O+(aq)][OH−(aq)]
or simply
Kw = [H+][OH−]
Hints & tips
Remember, Kw has no units as it is an equilibrium constant.
The reactants are not included in the equilibrium expression due to
water being a pure liquid and having a value of 1.
The equilibrium constant Kw is temperature dependent and the value
is approximately 1 × 10−14 at 25 °C.
As the reaction is endothermic, an increase in temperature moves
the equilibrium to the right (products side) and a decrease moves it
to the left (reactants side).
The pH scale
The pH scale gives a measure of the concentration of hydrogen ions
in an aqueous solution. It is a continuous scale, with values ranging
from less than zero to more than 14.
• If [H+] > [OH−], the solution will be acidic; pH < 7.
• If [OH−] > [H+], the solution will be alkaline or basic; pH > 7.
• If [H+] = [OH−], the solution is neutral; pH = 7.
The greater the [H3O+] or [H+] ion concentration, the lower the pH
value.
The pH of any aqueous solution can be calculated using the
expression
pH = −log10[H3O+] (or pH = −log10[H+]) and [H3O+] = 10−pH
The pH scale is a logarithmic scale; this means that any change in
pH unit requires a factor of 10 change in the hydrogen ion
concentration.
Using the ionic product of water, we can calculate the [H+], pH and
[OH−] for acidic and alkaline solutions.
Examples
1 Calculate the pH of a solution in which the hydrogen ion
concentration is 0.45 mol l−1.
Answer
2 Calculate the hydrogen ion concentration in a solution with a pH
of 9.1.
Answer
3 Calculate the pH in a solution where the hydroxide concentration
is 1.3 × 10−4 mol l−1.
Answer
Kw = [H+][OH−]
1.00 × 10−14 = [H+][1.3 × 10−4]
[H+] = 1.00 ×10−14/1.3 × 10−4
[H+] = 7.68 × 10−11 mol l−1
pH = −log10[H+] = −log10(7.68 × 10−11)
pH = 10.1
Study questions
1 Calculate the pH of a solution in which the hydrogen ion
concentration is 1.5 mol l−1.
[1 mark]
2 Calculate the hydrogen ion concentration of a solution with pH
9.3.
[1 mark]
3 Calculate the pH of a solution in which the hydroxide ion
concentration is 1.8 × 10−4 mol l−1.
[1 mark]
Chapter 11
Strong and weak acids and bases
Key points
• Strong acids and bases fully dissociate in water to ions. Weak
acids and bases only dissociate partially.
• Nitric, sulfuric and hydrochloric acids are examples of strong
acids; sodium hydroxide and potassium hydroxide are examples
of strong bases.
• Ethanoic acid, carbonic acid and sulfurous acid are examples of
weak acids; ammonia and amines are examples of weak bases.
• For acids and bases that fully ionise in solution, the [H+] or [OH–]
can be determined from the initial concentration of the acid or
base.
• Equations of the equilibria of reactions can be used to show the
weakly acidic/alkaline nature of solutions.
• The definition of pH for strong acids and bases is
pH = –log10[H+].
Acids
Strong acids
Acids are proton donors. Some acids are classified as strong and
some are classified as weak. Any acid, weak or strong, which is
capable of donating only one hydrogen ion per molecule is termed a
monoprotic or monobasic acid. An acid like sulfuric acid which can
donate two hydrogen ions is called diprotic or dibasic.
The degree of ionisation is the ability of the non-ionised acid
molecule to produce hydrogen ions. This is what determines whether
an acid is classified as strong or weak. The more an acid molecule
tends to produce hydrogen ions, H+(aq), the stronger it will be.
Hints & tips
Learn which acids and bases are strong. Nitric, sulfuric and
hydrochloric acids are strong acids, and sodium hydroxide and
potassium hydroxide are strong bases.
Strong acids tend to be lab acids, like hydrochloric acid, nitric acid
and sulfuric acid. They will dissociate (or ionise) fully in water. This
means that all the acid molecules become ions. This is shown by a
one-way arrow in the equation for the ionisation of the acid.
For nitric acid and sulfuric acid, the equations are given as
HNO3 → H+ + NO3–
H2SO4 → 2H+ + SO42 –
Nitric acid is a strong monoprotic acid as 1 mol of the acid will
produce 1 mol of hydrogen ions (protons). Sulfuric acid is described
as a strong diprotic acid as 1 mol of the acid produces 2 mol of
hydrogen ions (protons).
The strength of an acid is not related to the concentration of the
acid. The concentration of an acid is determined by the volume of
water that has been added to the undiluted acid. More water will
produce a more dilute acid, but it does not alter the strength of the
acid.
Weak acids
In a weak acid, the molecules only partially dissociate in the aqueous
solution. Weak acids are an equilibrium mixture of non-ionised acid
molecules, hydrogen ions and the conjugate base of the acid. Most
carboxylic acids are weak acids. The most common example is
ethanoic acid (CH3COOH). In ethanoic acid, the acidic hydrogen
atom is the one bonded to the oxygen atom. Only some of the
ethanoic acid molecules dissociate into ions. Ethanoic acid can be
described as a weak monoprotic acid as 1 mol of ethanoic acid could
produce 1 mol of hydrogen ions if it dissociated completely.
The equilibrium reaction for the ionisation of ethanoic acid is given
below. Note the reversible arrow in the equation.
CH3COOH ⇋ H+ + CH3COO–
Bases and alkalis
Strong bases
Any substance that reacts with an acid and accepts the proton from
the acid is classified as a base. A soluble base is called an alkali as
it will produce hydroxide ions when in an aqueous solution. Some
substances can act as a base, but are not soluble in water so cannot
be termed an alkali.
Hints & tips
All alkalis are bases but not all bases are alkalis.
A strong base dissociates completely in aqueous solution. Strong
bases are typically group 1 metal hydroxides, such as sodium
hydroxide or potassium hydroxide. Sodium hydroxide produces
hydroxide ions when dissolved in water as it dissociates completely,
as shown in the following equation:
NaOH → Na+ + OH–
Weak bases
The most common weak bases are ammonia (NH3) and the amino
alkanes (organic amines), i.e. compounds which contain the amine
functional group (–NH2). Weak bases are only partially dissociated in
aqueous solution. An equilibrium mixture of non-ionised base
molecules, hydroxide ions and the conjugate acid of the base is set
up, as shown in the example below for ammonia:
NH3(g) + H2O(l) → NH4OH(aq) ⇋ NH4+(aq) + OH– (aq)
The pH of strong acids and
bases
For acids and bases that ionise fully in solution, the [H+] or [OH–] can
be determined from the initial concentration of the acid or base.
The definition of pH for strong acids and bases is given by
pH = – log10[H+]
where the square brackets indicate the concentration of hydrogen
ions in solution in mol l–1.
To calculate the hydrogen ion concentration from the pH we reverse
the calculation:
[H+] = 10(–pH)
To calculate the pH of strong bases, we use Kw and the [OH–] to
calculate the [H+], which is then used to determine the pH.
Examples
1 A solution of sulfuric acid (H2SO4) has a pH of 2.7. Calculate
a) the concentration of hydrogen ions
b) the concentration of the acid.
Answer
a) [H+] = 10(–pH) = 10(–2.7) = 2.0 × 10–3 mol l–1
b) The acid concentration is not the same as the concentration
of hydrogen ions in a diprotic acid like H2SO4, as 2 moles of
H+ ions are dissociated in aqueous solution for every 1 mole
of H2SO4. The concentration of H2SO4 is half the
concentration of H+ ions in solution. So, the concentration of
acid is
0.5 × 2.0 × 10–3 mol l–1 = 1.0 × 10–3 mol l–1
2 A solution of sodium hydroxide has a concentration of 0.635 mol
l–1. Calculate
a) the concentration of hydrogen ions
b) the pH of the solution.
Answer
a) Kw = [H+][OH–]
1.00 × 10–14 = [H+](0.635)
[H+] = 1.00 × 10–14/0.635
[H+] = 1.575 × 10–14 mol l–1
b) pH = –log10[H+]
= –log10(1.575 × 10–14)
pH = 13.8
3 A solution of sodium hydroxide has a pH of 12.4. Calculate
a) the concentration of hydrogen ions
b) the concentration of hydroxide ions.
Answer
a) [H+] = 10(– pH) = 10(– 12.4) = 3.98 × 10–13 mol l–1
b) Kw = [H+][OH–]
1.00 × 10–14 = (3.98 × 10–13)[OH–]
[OH–] = 1.00 × 10–14/3.98 × 10–13
[OH–] = 0.0251 mol l–1
Hints & tips
Remember that in aqueous solutions, strong acids and strong
bases completely dissociate and weak acids and weak bases only
partially dissociate.
Study questions
1 At 50 °C, Kw = 5.48 × 10–14. Calculate the pH of 0.165 mol l–1
potassium hydroxide solution at 50 °C.
[1 mark]
2 Determine the concentration of nitric acid which has pH of 0.73.
Give your answer to three significant figures.
[2 marks]
3 Calculate the pH of sulfuric acid which has an acid concentration
of 0.250 mol l–1.
[2 marks]
Chapter 12
Acid dissociation constant and salts
Key points
• The acid dissociation constant, Ka, is given by
or by pKa where
pKa = – log10Ka
• The approximate pH of a weak acid can be found by using
• Equimolar solutions of weak and strong acids (or bases) have
different pH values, conductivity and reaction rates, but the
stoichiometry of reactions are the same.
• A soluble salt of a strong parent acid and a strong parent base
dissolves in water to produce a neutral solution. A soluble salt of
a weak parent acid and a strong parent base dissolves in water to
produce an alkaline solution. A soluble salt of a strong parent acid
and a weak parent base dissolves in water to produce an acidic
solution.
Weak acids and bases
As we learnt in the previous chapter, weak acids and bases do not
fully ionise in solution, so the [H+] or [OH–] cannot be determined
from the initial concentration of the acid or base. Examples of weak
acids were encountered in N5 and Higher and included carbon
dioxide, which is a soluble gas that will dissolve in/react with water to
produce an acidic solution of carbonic acid. Similarly, sulfur dioxide
from the air dissolves in solution to form the weak acid sulfurous
acid.
As these are reversible reactions, they set up equilibria and so an
equilibrium constant can be derived; this is called the acid
dissociation constant and is given the symbol Ka.
As only some of the hydrogen ions are found in solution, the
hydrogen ion concentration, [H+], is much lower than the
concentration of the acid.
This can be simplified to
For carbonic acid:
For sulfurous acid:
In a solution of a weak acid, only a very small proportion of the
original acid molecules dissociate into ions, and so the acid
concentration [HA] is the same as the original acid concentration and
is give the symbol c.
Every weak acid which dissociates into ions forms one hydrogen
conjugate acid ion and one conjugate base ion. So, ignoring any
hydrogen ions from water molecules, it can be said that
[H+] = [A–]
and that
Rearranging this equation for a weak acid gives
This enables us to calculate the hydrogen ion concentration [H+] if
the Ka is known.
The acid dissociation constant can also be represented by
pKa = – log10 Ka
Within the weak acids, the value of Ka gives a measure of how weak
(or how dissociated) the acid is: the smaller the value of Ka, the
weaker the acid. pKa helps these become whole numbers so the
larger the value of pKa, the weaker the acid or the less it dissociates
into ions; the equilibrium constant lies towards the left (the
reactants).
The pKa is also used for calculating the approximate pH of a weak
acid using
Examples
1 Calculate the pH of 0.20 mol l−1 butanoic acid.
Answer
Step 1: Look up the pKa of butanoic acid on page 13 of the data
booklet; pKa = 4.83
Step 2: Input numbers into
2 A solution of propanoic acid has a pH of 2.61. Calculate the
concentration of the acid.
Answer
Step 1: Look up the pKa of propanoic acid on page 13 of the data
booklet; pKa = 4.87
Step 2: Input numbers into
Divide each side by –
– 0.35 = log10c
c = 10(– 0.35) = 0.447 mol l–1
The most important weak base to know is ammonia, NH3. Ammonia
is a polar covalent gas molecule that dissolves readily in water. Most
of the ammonia (99.6%) remains as molecules, but a small amount
reacts with water to produce hydroxide ions. Other weak acids
include the organic ‘ammonia derivatives’, the amines.
The weaker the base, the stronger the conjugate acid that will be
formed with a higher Ka value. In other words, the conjugate acid of
a weak base is a strong acid.
Comparing the pH and conductivity results, it is seen that strong
acids with greater dissociation have a greater concentration of
hydrogen ions and so have a lower pH and greater conductivity.
Weak bases produce fewer OH– ions than strong bases and are,
therefore, less alkaline than strong bases of equal concentration.
Weak bases have a lower pH and therefore conductivity. Reaction
rates are faster with higher conductivity as there are more ions
present at the start of the reaction.
It is important to note, however, that the stoichiometry of acid–alkali
reactions is not changed by changing from a strong acid to a weak
acid. It takes the same number of moles of alkali to neutralise one
mole of a weak acid as it takes to neutralise one mole of a strong
acid. This is due to the equilibrium equation of weak acids shifting to
form more hydrogen ions as they form water molecules in the
neutralisation reaction.
1 mole NaOH neutralises 1 mole HCl (strong acid):
HCl + NaOH → NaCl + H2O
1 mole NaOH neutralises 1 mole CH3COOH (weak acid):
CH3COOH + NaOH → CH3COONa + H2O
pH of salt solutions
The pH of a salt solution depends on the strength of the parent acid
and parent base from which it was formed, as summarised in Table
12.1. All salts are strong electrolytes and dissociate completely in
solution.
Hints & tips
The pH of a salt solution depends upon the strength of the parent
acid and the parent base.
In soluble salts formed from strong acids and strong bases, both the
parent acid and parent base dissociate fully in water and the water
equilibrium is unaffected.
However, when a weak acid or weak base is dissolved, the ions will
set up an equilibrium with the ions in water. This disrupts the water
equilibrium.
Consider ammonium nitrate, which is acidic in solution and so has
more H+ ions than OH– ions. This means that the water equilibrium
has been disrupted and some of the OH– ions present have been
removed. The positive ammonium ions present in ammonium nitrate
have reacted with the negative hydroxide ions to form ammonia
molecules.
Water equilibrium:
H2O(l) ⇋ H+(aq) + OH– (aq)
Ammonium equilibrium:
NH4+(aq) + OH– (aq) ⇋ NH3(aq) + H2O(l)
This equilibrium removes some of the negative hydroxide ions from
the water equilibrium. Now there are more hydrogen ions than
hydroxide ions, so the solution is acidic.
A solution of a salt of a weak acid and a strong base has a pH
greater than 7. When sodium ethanoate dissolves in water, it forms
an alkaline solution. This is because the ethanoate ions react with
hydrogen ions in the water equilibrium to form ethanoic acid
molecules. This means there are now fewer H+ ions than OH– ions
and so the salt solution will have a pH greater than 7.
Study questions
1 Two salts can be formed when sulfurous acid reacts with sodium
hydroxide. Sodium sulfite is one of the salts and it dissolves in
water to give an alkaline pH. Explain clearly why sodium sulfite
forms an alkaline solution. (Include reference to both of the
equilibrium reactions which occur.)
[2 marks]
2 Calculate the pH of 0.15 mol l–1 methanoic acid.
[1 mark]
3 Which of the following will dissolve in water to give a solution of
pH < 7?
A Potassium ethanoate
B Sodium sulfate
C Sodium hydroxide
D Ammonium chloride
[1 mark]
Chapter 13
Indicators and buffers
Key points
• A buffer solution is one in which the pH remains approximately
constant when small amounts of acid or base are added.
• An acid buffer consists of a solution of a weak acid and one of its
salts, and a basic buffer consists of a solution of a weak base and
one of its salts. The required composition of an acid buffer
solution can be calculated from the desired pH and the acid
dissociation constant.
• An approximate pH of an acid buffer solution can be calculated
from the acid dissociation constant:
• Indicators are weak acids for which the dissociation can be
represented as:
HIn(aq) + H2O(l) ⇋ H3O+(aq) + In– (aq)
• In aqueous solution the colour of the acid is distinctly different
from that of its conjugate base.
• The colour of the indicator is determined by the ratio of [HIn] to
[In–].
• The theoretical point at which colour change occurs is when
[H3O+] = KIn.
• The colour change is assumed to be distinguishable when [HIn]
and [In–] differ by a factor of 10.
• The pH range over which a colour change occurs can be
estimated by the expression:
Buffer solutions
Buffer solutions are defined as solutions that resist the change in
pH, in other words remain at a nearly constant pH, on adding small
amounts of either acid or base or by diluting the solution with water.
An interesting application of buffers is their use in skin creams. Anti-
wrinkle creams are buffered at pH values lower than 5.5. At this pH,
the skin is irritated so swells and thus hides the wrinkles.
Buffer solutions exist as one of two main types according to their
components: acid buffers and basic buffers:
• In an acid buffer, the solution contains a mixture of a weak acid
and its conjugate base. The pH of an acid buffer solution is lower
than 7.
• In a basic buffer, the solution contains a mixture of a weak base
and its conjugate acid. The pH of a basic buffer solution is higher
than 7.
A buffer works by setting up an equilibrium in each buffer solution.
The buffering action of an acidic buffer can be represented by this
general equilibrium equation:
For a buffer to stabilise the pH of a solution, it must be able to react
with added OH– ions and added H+ ions.
A typical acid buffer is a mixture of a weak acid like ethanoic acid
and its salt sodium ethanoate. This uses the equilibrium of the weak
acid in solution. There are lots of ethanoic acid molecules present
since it only partially dissociates into ethanoate ions (conjugate
base) and hydronium ions (conjugate acid):
CH3COOH(aq) + H2O (l) ⇋ CH3COO– (aq) + H3O+(aq)
Addition of a small volume of dilute acid to this buffer solution
increases the concentration of the hydrogen ions in the equilibrium
shown above. This disrupts the equilibrium position, shifting it to the
left-hand side so the ethanoate ions remove the excess hydrogen
ions:
CH3COO– (aq) + H3O+(aq) ⇋ CH3COOH(aq) + H2O(l)
Hints & tips
Remember that a carboxylic acid and one of its corresponding salts
can be used (when combined) to make a buffer.
This then keeps the pH almost constant as the hydrogen ion
concentration in the new equilibrium mixture is similar to that at the
start.
Addition of a small volume of dilute base to this buffer solution
increases the concentration of hydroxide ions in the equilibrium. The
hydroxide ions react with the hydrogen ions in the buffer, so
removing some of the hydrogen ions. This causes the position of
equilibrium to move to the right to replace the hydrogen ions which
have reacted with the hydroxide ions, so keeping the pH constant.
An example of a basic buffer is formed from a weak base such as
ammonia and its salt ammonium chloride. The buffering action of this
basic buffer is shown as:
NH4+(aq) + H2O(l) ⇋ NH3(aq) + H3O+(aq)
Hints & tips
Hydrogen ions are removed when a small amount of acid is added
to a buffer solution due to the salt of the weak acid providing the
conjugate base to remove them.
When small volumes of acid are added, the additional hydrogen ions
are removed as they react with the ammonia in the solution to form
ammonium ions. The position of the equilibrium moves from right to
left and so the pH remains almost constant.
When small volumes of base are added, the system tries to minimise
the effect by removing the hydroxide ions, which react with the
hydrogen ions in the buffer to form water. The equilibrium moves
from left to right to replace the removed hydrogen ions.
The pH of an acid buffer can be calculated using the following
equation:
where the square brackets mean ‘concentration of’. Given the
concentrations of acid and salt, the pH of the buffer can be obtained
using the pKa of the weak acid as given in the data booklet.
Examples
1 A buffer solution was prepared by dissolving 0.013 mol of sodium
ethanoate in 100 cm3 of 0.058 mol l–1 ethanoic acid. Using the
pKa of ethanoic acid from the data booklet, calculate the pH of
this buffer solution.
Answer
Step 1: Find the pKa of ethanoic acid from page 13 of the data
booklet; pKa = 4.76
Step 2: Work out the concentration of ethanoate ions:
Step 3: Input data into the formula:
2 What concentration of methanoic acid and sodium methanoate
would be required to make a buffer solution with a pH of 4.5?
Answer
Step 1: Find the pKa of methanoic acid on page 13 of the data
booklet; pKa = 3.75
Step 2: Using this equation, input the numerical data:
A ratio of 5.62 mol of sodium methanoate to 1 mol of methanoic
acid will give the required pH for this buffer. Any mixture in this
ratio in any volume of water will give the correct pH.
Hints & tips
Practise being able to calculate the pH of an acid buffer solution.
Indicators
Indicators are weak acids or bases that are able to give a measure
of the pH of a solution by their colour. There are many different
indicators which change from one colour to another at different pH
values. When deciding which indicator to use in a titration, the
equivalence point in the titration must be considered. The
equivalence point is the point in a titration where the amount of titrant
added is enough to neutralise the analyte solution completely. The
shape of the titration curve indicates the type of titration (Figure
13.1). The vertical region in a titration curve shows where there is a
rapid change in pH. This area also shows the equivalence point in
the titration.
Figure 13.1 Titration curves
• For a strong acid–strong base titration, the vertical region occurs
between pH3 and pH10. So a suitable indicator would change
colour within this pH range.
• For a weak acid–strong base titration, the vertical region occurs
between pH7 and pH10.
• For a strong acid–weak base titration, the vertical region occurs
between pH3 and pH8.
• For a weak acid–weak base titration, there is no sharp vertical
region. This means that the pH does not change rapidly at any
point and so no suitable indicator is available. To determine the
equivalence point for a weak acid–weak base titration, a pH meter
is used to produce a titration curve. The curve is plotted in order to
identify the equivalence point.
The equivalence point depends on the strengths of the acid and
base used and the nature of the salt formed at the end point.
When choosing an indicator, a colour change at a pH equal to or
very close to the pH of the equivalence point of that acid–base
reaction is desired. This means that the colour change occurs in the
vertical region of the titration curve. The equivalence point should be
so sharp and distinct that a colour change is seen with the addition
of one drop of base. A list of indicators and their pH ranges is given
on page 20 of the data booklet. Table 13.1 shows some common
laboratory indicators along with their colour change and pH ranges.
The dissociation of litmus can be represented by the following
equilibrium equation:
where litmus is represented by HIn (‘In’ from ‘indicator’).
Adding acid to this equilibrium will increase the concentration of
hydronium ions (H3O+) and will shift the position of equilibrium to the
left-hand side; the solution will become red in colour.
Adding a base will increase the concentration of hydroxide ions
(OH–), which remove the H3O+ ions from the equilibrium. This
causes the equilibrium position to shift to the right-hand side, so a
blue colour is now observed.
The equilibrium constant can be determined by the equation
This can be rearranged to
The indicator colour during the titration depends on the concentration
of the non-ionised indicator molecules [HIn] and the conjugate base
molecules [In–]. The theoretical point at which the pH changes is
when [HIn] = [In–]. Therefore KIn = [H3O+]. In other words, the colour
change takes place when pKIn = pH. As pH and pKIn are based on a
logarithmic scale, in practice the colour change can only be seen
when [HIn] differs from [In–] by a factor of 10.
The pH range over which a colour change can be seen is therefore
established using the expression
pH = pKa ± 1
The appropriate indicator for any titration can then be chosen so the
colour change occurs over the pH range when the pH is changing
very rapidly.
Hints & tips
There is no indicator which is suitable for a weak acid–weak base
titration.
Study questions
1 A buffer solution was prepared by dissolving 0.02 mol of sodium
benzoate in 200 cm3 of 0.40 mol l–1 benzoic acid. Using the pKa
of benzoic acid from the data booklet, calculate the pH of this
buffer solution.
[2 marks]
2 A buffer solution with a pH of 4.76 contains a mixture of butanoic
acid and potassium butanoate. Calculate the molar proportion of
acid to salt that would be required to make this buffer solution.
[2 marks]
3 Explain which of these indicators – methyl red, phenolphthalein,
bromothymol blue and p-nitrophenol – could be used to detect
the end point of a titration involving these solutions:
a) nitric acid and ammonia
b) methanoic acid and potassium hydroxide
c) sulfuric acid and sodium hydroxide.
[3 marks]
Chapter 14
Reaction feasibility, standard enthalpy
of formation and entropy
Key points
• Thermochemistry concerns the study of changes in energy which
occur during chemical reactions.
• The standard molar enthalpy of formation, ΔHf, refers to the
enthalpy change which occurs when one mole of a substance is
prepared from its elements in their standard states.
• The standard state of a substance is its most stable state at a
pressure of 1 atmosphere and a specific temperature, usually 298
K.
• The standard enthalpy of formation of a substance can be
calculated from standard enthalpy changes which are determined
experimentally:
ΔH° = ΣΔH°f(products) – ΣΔH°f(reactants)
• The entropy (S) of a system is the degree of disorder of the
system. The greater the disorder, the greater the entropy.
• Solids have low disorder and gases have high disorder.
• Entropy increases as temperature increases.
• Changes of state involve changes in entropy. Melting and
evaporation are accompanied by increases in entropy.
• The second law of thermodynamics states that the total entropy
of a reaction system and its surroundings always increases for a
spontaneous process.
• Heat energy released by the reaction system into the
surroundings increases the entropy of the surroundings. Heat
energy absorbed by the reaction system from the surroundings
decreases the entropy of the surroundings.
• The third law of thermodynamics states that the entropy of a
perfect crystal at 0 K is zero.
• The change in standard entropy for a reaction system can be
calculated from the standard entropy values for the reactants and
products:
ΔS° = ΣS°(products) – ΣS°(reactants)
Standard enthalpy of formation
When a chemical reaction takes place, there is often an exchange of
heat energy between the reaction and its surroundings. The study of
changes in energy which occur during such reactions is called
thermochemistry.
The change in energy is also called the enthalpy change and is
given the symbol ∆H°. An enthalpy change can be either
exothermic (heat lost to the surroundings so ∆H is negative) or
endothermic (heat taken in from the surroundings so ∆H is
positive).
Enthalpy changes can be measured directly or indirectly:
• Direct methods include calorimetry, in which temperature changes
are measured as a reaction happens. You will have met this
method at N5 and Higher.
• Indirect methods include Hess’s law calculations and bond
enthalpy calculations, covered in the Higher Chemistry course.
The standard enthalpy of formation of a compound (ΔH°f) is the
energy given out or taken in when one mole of a compound is
formed from its elements in their standard states. The standard
state of an element is the most stable state of the substance under
standard conditions. Standard conditions are a pressure of one
atmosphere and a specific temperature, usually room temperature
298 K (25 °C). Enthalpies of formation are used in conjunction with
Hess’s law to provide standard enthalpy changes (measured in kJ
mol–1) of a reaction. The enthalpy change for a reaction is given by
this equation:
ΔH° = ΣΔH°f(products) – ΣΔH°f(reactants)
Table 14.1 summarises the energy changes for exothermic and
endothermic reactions.
Enthalpies of formation can be found in the data booklet or will be
stated in the question. The enthalpy of formation of an element in its
most stable form is zero.
Examples
1 Calculate the standard enthalpy of reaction for the decomposition
of copper carbonate from the given standard enthalpies of
formation.
CaCO3(s) → CaO(s) + CO2(g)
Substance ∆H°f/kJ mol–1
CaCO3(s) –1207
CaO(s) –635
CO2(g) –394
Answer
2 Calculate the standard enthalpy of reaction for the reduction of
aluminium oxide from the given standard enthalpies of formation.
Al2O3(s) + 3H2(g) → 2Al(s) + 3H2O(l)
Substance ∆H°f/kJ mol–1
Al2O3(s) –1676
H2(g) 0
Al(s) 0
H2O(l) –286
Answer
Entropy
A reaction that occurs at a particular temperature is called a feasible
reaction. However, the enthalpy change by itself is not enough to
determine whether a reaction will be feasible. The entropy of a
system also has to be taken into account.
Entropy is the degree of disorder or randomness in a system.
Entropy is given the symbol S and is measured in J K–1 mol–1.
Standard entropy values (S°) or absolute entropy values are
calculated for one mole of the substance based on a scale where the
substance has an entropy value of zero at 0 K. This assumes that
the crystalline form of any substance at 0 K is a perfect crystal. This
is given by the third law of thermodynamics.
The higher the absolute entropy value, the greater the degree of
disorder in a substance. It is seen that solids are more ordered than
liquids, which are more ordered than gases. As a general rule there
is an increase in entropy from solid to liquid; the increase in entropy
from liquid to gas is much larger as gases are very disordered. Table
14.2 gives the entropy values for water in its various states. Entropy
can be seen to increase from ice to water to water vapour. Figure
14.1 illustrates this idea with a particle diagram.
Table 14.2 Entropy values for water in its three states
Water state S°/J K–1 mol–1
Solid, H2O(s) 48
Liquid, H2O(l) 70
Gas, H2O(g) 189
Figure 14.1 The changes in the arrangement of water
molecules when ice melts to form a liquid, and then evaporates,
show an increase in disorder
Figure 14.2 is a graph of entropy values plotted against temperature.
It can be seen that at 0 K, the entropy of a perfect crystal is zero,
and there is a general increase in entropy as temperature increases.
As the entropy increases in a solid, the particles gain energy and
vibrate more, so increasing their disorder.
Figure 14.2 A graph of the entropy values of water against
temperature
As the solid melts and forms a liquid, there is a large increase in
disorder at that particular temperature and this is observed by a
rapid vertical line. This temperature is the melting point. As the
liquid continues to increase in temperature, so does the disorder until
another rapid rise is observed. This temperature, where the liquid
turns to gas, is the boiling point.
Entropy is a state function (solid, liquid or gas) and values can vary
with temperature and pressure. Solids have less entropy than
liquids, which have less entropy than gases. In general, the higher
the molecular weight of substances in the same state, the higher the
entropy value. The entropy change in a reaction can be calculated
from absolute entropy values:
ΔS° = ΣS°(products) – ΣS°(reactants)
Examples
3 Calculate the change in standard entropy for the decomposition
of copper carbonate by heating from the given standard entropy
values.
CaCO3(s) → CaO(s) + CO2(g)
Substance S°/J K–1 mol–1
CaCO3(s) 93
CaO(s) 40
CO2(g) 214
Answer
This reaction shows an increase in entropy as the solid calcium
carbonate decomposes to produce calcium oxide solid and
carbon dioxide gas. The products (1 mol of gas and 1 mol of
solid) have more disorder than the reactant (1 mol of solid).
4 Calculate the change in standard entropy for the Haber process
from the given standard entropy values.
N2(g) + 3H2(g) → 2NH3(g)
Substance S°/J K–1 mol–1
N2(g) 192
H2(g) 131
NH3(g) 193
Answer
This reaction shows a decrease in entropy as the system
becomes more ordered: 4 mol of reactant gas become 2 mol of
product gas. Fewer moles of gas have less disorder, so giving a
decrease in entropy.
From a reaction equation, it is often possible to predict whether the
entropy change is positive or negative. A reaction is sometimes
called a system. So it is often said that the system has decreased or
increased in entropy, meaning that ΔS° is either negative or positive,
respectively.
The fact that both entropy and enthalpy are involved in reaction
feasibility is summarised in the second law of thermodynamics.
This means for a reaction to be spontaneous there must be an
increase in the overall entropy of the reaction system and its
surroundings. This relationship is examined in more detail in the next
chapter, on Gibbs free energy.
Hints & tips
Remember, it is important to multiply the standard enthalpy of
formation or entropy by its corresponding stoichiometric coefficient.
For example, if two moles of a reactant appear in the balanced
equation for the reaction, the enthalpy or entropy value must be
multiplied by 2 before being used in the calculations.
Study questions
1 Calculate the standard enthalpy of reaction for the combustion of
ethene from the given standard enthalpies of formation.
[3 marks]
C2H4(g) + 3O2(g) → 2CO2(g) + 2H2O(l)
Substance ∆H°f/kJ mol–1
C2H4(g) +52
CO2(g) –394
H2O(l) –286
2 Ammonia reacts with oxygen to form nitrogen monoxide and
steam according to this equation:
4NH3(g) + 5O2(g) → 4NO(g) + 6H2O(g)
Substance ∆H°f/kJ mol–1
NH3(g) –46
NO(g) +90
H2O(g) –242
Using the standard enthalpies of formation, calculate the
enthalpy change for this reaction.
[3 marks]
3 Calculate the change in standard entropy for the oxidation of iron
from the given standard entropy values.
[3 marks]
3Fe(s) + 2O2(g) → Fe3O4(s)
Substance S°/J K 1 mol 1
Fe(s) 27
O2(g) 205
Fe3O4(s) 146
4 2 Predict whether the entropy change of the following reactions is
positive or negative and explain your answer:
a) NH4NO3(s) → N2O(g) + 2H2O(g)
b) CaO(s) + H2O(l) → Ca(OH)2(s)
c) 2NaHCO3(s) → Na2CO3(s) + CO2(g) + H2O(l)
[3 marks]
Chapter 15
Gibbs free energy
Key points
• The change in entropy of the surroundings that occurs as a result
of a chemical reaction can be calculated from the temperature
and the enthalpy change for the reaction system.
• The total entropy change is proportional to the change in free
energy (ΔG°) of the reaction system.
• The direction of spontaneous change is in the direction of
decreasing free energy.
• The change in standard free energy for a reaction is related to
the standard enthalpy and entropy changes by
ΔG° = ΔH° – TΔS°
• The feasibility of a chemical reaction under standard conditions
can be predicted from the calculated value of the change in
standard free energy.
• The standard free energy change for a reaction can be calculated
from the standard free energies of formation of the reactants and
the products:
ΔG° = ΣΔG°(products) – ΣΔG°(reactants)
• The temperature at which the reaction becomes feasible for a
reaction for which both ΔH° and ΔS° have positive values can be
calculated using
• A reaction is feasible under standard conditions if the change in
standard free energy between reactants and products is negative.
• This means that the equilibrium composition favours the products
over the reactants.
• Under non-standard conditions, any reaction is feasible if ΔG° is
negative.
• An equilibrium reaction will proceed spontaneously in the forward
direction until the composition is reached where ΔG° = 0.
• Industrial systems can employ the removal of a product to ‘force’
the equilibrium in the desired direction.
Gibbs standard free energy
In order for a chemical reaction to be feasible, both enthalpy and
entropy have to be taken into consideration. The feasibility of a
reaction depends on the enthalpy change of the reaction, the entropy
change and the temperature at which the reaction is occurring.
Gibbs standard free energy is a measure of the potential work that
may be done by a system at constant temperature and pressure. It is
a thermodynamic property that was defined in 1876 by Josiah
Willard Gibbs to predict whether a process will occur spontaneously
at constant temperature and pressure.
Standard free energy (ΔG) is defined by the following relationship:
ΔG° = ΔH° – TΔS°
The equation can be used to calculate ΔG° at a certain temperature
and so determine whether a reaction is feasible. For a reaction to be
feasible, ΔG° must be equal to or less than zero; this shows that the
reaction is likely to happen spontaneously. A positive value for ΔG°
shows that the reaction is unlikely to happen unless external energy
is provided to the system (in other words, unless the reaction is
heated).
There is a difference between a reaction being feasible and a
reaction occurring spontaneously. A reaction may be feasible at a
particular temperature but may not proceed due to the activation
energy being too high. For a reaction to be spontaneous, ΔG° must
be less than zero and the activation energy for the reaction must be
low enough for the reaction to proceed under standard conditions
(298 K and 1 atm pressure).
Hints & tips
When ΔG° = 0, a system is at equilibrium so both the forward and
reverse reactions are feasible.
To deduce the temperature at which a reaction occurs
spontaneously, assuming that ΔG° = 0, the equation can be rewritten
as
ΔH° = TΔS°
which gives us
This is the temperature at which the standard free energy, ΔG°, will
start to be negative and the reaction will be spontaneous.
Table 15.1 summarises the factors that affect ΔG° and the feasibility
of a reaction.
Hints & tips
Both exothermic and endothermic reactions can occur
spontaneously.
Examples
1 The thermal decomposition of calcium carbonate is given by the
equation below. Use the data given to:
a) calculate ΔG° at 298 K
b) calculate the temperature range over which the reaction is
spontaneous.
CaCO3(s) → CaO(s) + CO2(g)
ΔH° = +178 kJ mol–1;
ΔS° = +160 J K–1 mol–1
Answer
a) ΔG° = ΔH°–TΔS°
= +178 – (298 × (160/1000)) = 178 – 47.7 = +130 kJ mol–1
b) ΔAs both ΔH° and ΔS° are positive, when ΔG° = 0
ΔS° = +160/1000 = 0.160 kJ K–1 mol–1
Therefore T = ΔH°/ΔS° = 178/0.160 = 1113 K so the reaction is
spontaneous thermodynamically above 1113 K.
2 Given the data provided, calculate ΔG° at 298 K for the reaction
ZnO(s) + CO(g) → Zn(s) + CO2(g)
ΔH° = +65 kJ mol–1;
ΔS° = +14 J K–1 mol–1
Answer
This reaction is not spontaneous at 298 K as the sign for ΔG° is
positive. So, it is not possible for carbon monoxide to reduce zinc
oxide at 298 K.
3 Given the data provided, calculate the temperature at which
carbon monoxide can reduce zinc oxide to zinc.
ZnO(s) + CO(g) → Zn(s) + CO2(g)
ΔH° = +65 kJ mol–1;
ΔS° = +14 J K–1 mol–1
Answer
ΔG° = 0 AND ΔS° = +14/1000 = 0.014 kJ K–1 mol–1
Therefore T = ΔH°/ΔS° = 65/0.014 = 4643 K
4 Barium carbonate decomposes when heated according to the
following equation:
BaCO3 → BaO(s) + CO2(g).
Calculate the temperature at which the decomposition of barium
carbonate becomes just feasible when ΔH° = +266 kJ mol–1 and
ΔS° = +174 J K–1 mol–1.
Answer
ΔG° = 0 AND ΔS° = +174/1000 = 0.174 kJ K–1 mol–1 Therefore T
= ΔH°/ΔS° = 266/0.174 = 1529 K
Hints & tips
The step which is most often forgotten in such calculations is
converting ΔS° to kJ K–1 mol¬[minus]1 by dividing the given value by
1000 before using it in the ΔG° expression.
The overall standard free energy change can also be determined
from the standard free energies of formation of reactants and
products:
ΔG° = ΣΔG°(products) – ΣΔG°(reactants)
ΔG° is the Gibbs free energy change under standard conditions,
comparing the free energies of pure products and reactants.
The standard free energy of formation of an element is the most
stable form, so is always zero.
Free energy and equilibrium
Most reactions can be considered as equilibrium reactions to some
extent.
We know from previous chapters that at equilibrium both the forward
and the reverse reactions are possible, the concentrations of
reactants and products are constant (not the same!) and the system
has maximum disorder (highest entropy).
Equilibrium is reached when the free energy of the reactants has
fallen (by the reactants being used up) to the same value as the
products (i.e. ΔG = 0).
At equilibrium, the free energy of the system has reached its
minimum value under the given conditions:
ΔG = ΣΔG(products) – ΣΔG(reactants) = 0
In any chemical system, once the reaction has started ΔG will be
lower than ΔG°. This is as a result of the change from standard
conditions and the increased entropy due to the mixing of reactant
and product.
This change can be observed by looking at the variation of free
energy against equilibrium composition (Figure 15.1).
Figure 15.1 Variation of free energy with composition
Hints & tips
In data tables, the values for free energy, enthalpy and entropy are
always quoted at what is known as standard conditions and are
given the symbols ΔH°, ΔS° and ΔG°. This is because the values
can vary with temperature, pressure and concentration. Standard
conditions are a temperature of 298 K (25 °C), a pressure of 1 atm
and a concentration of 1 mol l−1.
At equilibrium, standard conditions do not apply and so ΔG° is no
longer used; ΔG is used instead.
From Figure 15.1 it can be seen that as the reaction proceeds to
equilibrium, the free energy of the system drops to the lowest free
energy point in the diagram (indicated by *), and this is where
equilibrium is established (where ΔG = 0 not ΔG° = 0).
• In a system where ΔG° is negative, the reaction is
thermodynamically feasible and the equilibrium mixture favours the
formation of products. The equilibrium position will lie to the side of
the products (the right-hand side) and the equilibrium constant will
be larger than 1.
• If ΔG° is positive, then the forward reaction will not be feasible and
so the equilibrium mixture will favour the reactants. The equilibrium
position will lie to the side of the reactants (left-hand side) and the
equilibrium constant will be very small, at less than 1. The fact that
standard free energy change is positive does not mean that no
reaction will take place. It only means that the reaction favours
reactants over products.
• When ΔG° is zero, the reaction mixture is 50% reactants and 50%
products and so neither side is favoured. The equilibrium position
exists at the middle of the reaction and is called the balance
point, i.e. products and reactants are equally favoured. The
equilibrium constant will be equal to 1.
These thermodynamic calculations do not provide any information
about either the reaction pathway or the reaction rate.
Thermodynamics tells us only how far a reaction will proceed (given
enough infinite time); it never tells us how fast a reaction will take
place or by which route. The fact that the standard free energy
change is greater than zero does not mean that no reaction takes
place. It means that the products are less favoured than the
reactants.
Hints & tips
If a reaction is feasible, but does not occur spontaneously, the
activation energy is high.
In industry, equilibrium is rarely allowed to occur and removal of a
product in an industrial process can ‘force’ the equilibrium in the
desired direction by using Le Chatelier’s principle. An example would
be the Haber process, where removal of the product ammonia forces
the equilibrium position to the right-hand side to minimise the
change.
Study questions
1 Magnesium carbonate decomposes on heating according the
equation:
MgCO3(s) → MgO(s) + CO2(g)
ΔH° = +100 kJ mol–1;
ΔS° = +175 J K–1mol–1
a) Calculate the standard free energy, ΔG°, at 900 K.
b) Calculate the minimum temperature needed for MgCO3(s) to
decompose.
[3 marks]
2 When heated strongly, copper(II) sulfate decomposes according
to the equation:
CuSO4(s) → CuO(s) + SO3(g)
ΔH° = +210 kJ mol–1;
ΔS° = +185 J K–1mol–1
Calculate the minimum temperature needed for CuSO4(s) to
decompose.
[2 marks]
3 Nuclear fuel is made from uranium ore via this equation:
UO2(s) + 4HF(g) → UF4(s) + 2H2O(g)
ΔH° = – 244 kJ mol–1;
ΔS° = –243 J K–1mol–1
Determine, by calculation, whether this reaction is feasible at 298
K.
[2 marks]
Chapter 16
Kinetics
Key points
• The rate of a chemical reaction normally depends on the
concentrations of the reactants.
• The rate can be expressed as
rate = k[A]
where k is the rate constant and [A] is the concentration of
reactant A.
• Orders of reaction are used to relate the rate of a reaction to the
reacting species.
• If changing the concentration of a reactant, A, has no effect on
the rate of reaction, the reaction is said to be zero order with
respect to A. This is written as [A]0.
• If doubling the concentration of A doubles the rate of reaction,
the reaction is said to be first order with respect to A. This is
written as [A]1.
• If doubling the concentration of A increases the rate of reaction
fourfold, the reaction is said to be second order with respect to
A. This is written as [A]2.
• The overall order of a reaction is the sum of the powers to which
the concentrations of the reactants are raised in the rate
equation:
rate = k[A]m[B]n
where m and n are rate coefficients, derived from practical work
or the slow rate-determining step from the mechanism of the
reaction.
• The order of a reaction can only be determined from
experimental data.
• Reactions usually occur by a series of steps called a reaction
mechanism. The rate of a reaction is dependent on the slowest
step, called the rate-determining step.
Determining rate equation,
order of reaction and rate
constant
Kinetics is the study of factors which alter the rates of chemical
reactions. The rate of a chemical reaction is monitored by the rate of
change of the concentration.
Consider a reaction where
A+B→C
The production of C is not directly proportional to the rate of
decrease of the concentrations of A and B. However, the formation
of C has a relationship with the concentrations of the reactants. So
the rate of formation of C can be derived from the order at which the
concentrations of the reactants decrease. So the relationship is
rate of formation of C ∝ [A]m + [B]n
where m and n are the orders of reaction with respect to the
reactants.
So an overall rate equation can be determined as
rate = k[A]m[B]n
where k is the rate constant, which is dependent upon temperature,
and m and n are the orders of the reaction with respect to the
reactants.
Orders of reaction are used to relate the rate of a reaction to the
reacting species:
• If changing the concentration of a reactant, A, has no effect on the
rate of reaction, the reaction is said to be zero order with respect to
A. This is written as [A]0.
• If doubling the concentration of A doubles the rate of reaction, the
reaction is said to be first order with respect to A. This is written as
[A]1.
• If doubling the concentration of A increases the rate of reaction
fourfold, the reaction is said to be second order with respect to A.
This is written as [A]2.
Hints & tips
The units of the rate constant, k, vary according to the total order of
reaction.
Table 16.1 Units of the rate constant, k
Overall reaction order k units
1st order s–1 (or min–1)
2nd order mol–1 l s–1 (or mol–1 l min–1)
3rd order mol–2 l2 s–1 (or mol–2 l2 min–1)
The overall order of a reaction is the sum of the powers to which
the concentrations of the reactants are raised in the rate equation:
rate = k[A]m[B]n
In other words, the overall order is given by m + n. For example, if m
= 2 and n = 1, the rate would be 2nd order with respect to A and 1st
order with respect to B and the overall rate order would be 3 (2 + 1).
To obtain the order of the reaction, find any two experiments to
compare. Then look at the effect that the concentration change has
had on the rate of the reaction (Table 16.2).
Let’s consider the data given in Table 16.3 to obtain the rate
equation for the reaction.
Comparing experiments 1 and 2, it can be seen that doubling [A]
doubles the rate. This gives the information that the order with
respect to A is 1.
Comparing experiments 2 and 3, it can be seen that doubling [B] has
no effect. The order of reaction with respect to B is, therefore, 0.
From these results we can see that the overall rate equation is
rate = k[A]1[B]0
This can be simplified to
rate = k[A]1
[B] disappears from the equation as any number to the power of 0
has a value of 1.
The units for the rate constant can be worked out using the
rearranged equation:
The results from each experiment should give the same rate
constant, as each experiment was carried out at the same
temperature.
• Using experiment 1 data:
• Using experiment 4 data:
To determine the units, insert the units of rate and concentration into
the equation and solve for k:
Reaction mechanisms
A reaction mechanism is a set of theoretical steps proposed to
account for the reactants being converted into products. The slowest
step is the one which controls the overall rate of reaction and is
called the rate-determining step.
Consider the reaction of nitrogen dioxide and hydrogen:
NO2(g) + H2(g) → NO(g) + H2O(l)
When examining the kinetics of the experiment, it can be seen that
the reaction must occur using a two-step process.
Step 1: NO2(g) + NO2(g) → NO3(g) + NO(g)
This step is slow and is the rate-determining step.
Step 2: NO3(g) + 2H2(g) → NO(g) + 2H2O(l)
This step is fast.
Hints & tips
No relationship exists between the rate equation and the
stoichiometric coefficients in a balanced equation.
The first step involves two molecules of NO2, and so doubling the
concentration of NO2 will make collisions in this step four times more
likely. So the reaction is second order with respect to NO2. H2 is not
involved in step 1 so the reaction is zero order with respect to H2.
The rate equation is, therefore, determined to be
rate = k[NO2]2
A rate equation does not provide definitive proof that this is the
reaction mechanism. Alternative mechanisms may give the same
rate equation.
Study questions
1 The following rate equation was determined from a series of
experiments:
rate = k[X][Y]
The units of the overall rate constant for the reaction above are
A s–1
B mol–1 l s–1
C mol–1 l–1 s–1
D mol–2 l2 s–1
[1 mark]
2 Consider the following reaction between phosphorus and
hydrogen:
P4 + 6H2 → 4PH3
When the reaction was carried out experimentally, the data in the
table were obtained.
Write the rate equation and calculate a value for the rate
constant, with units.
[3 marks]
Unit 2 Glossary
Buffer solution: A solution in which the pH remains approximately
constant when small amounts of acid or base are added.
Closed system: A closed system has no exchange of matter or
energy with its surroundings.
Conjugate acid: For every base, there is a conjugate acid formed
by the gain of a proton (H+).
Conjugate base: For every acid, there is a conjugate base formed
by the loss of a proton (H+ ion).
Diprotic: An acid which is able to produce two H+ ions from each
molecule.
Dynamic equilibrium: A dynamic equilibrium is achieved when the
rates of two opposing processes become equal, so that no net
change results.
End point: The end point of a titration is the point at which the
reaction is shown to be complete. This is frequently signalled by the
change in colour of an indicator.
Entropy: The entropy of a system is the degree of disorder of the
system. The greater the disorder, the greater the entropy. Low
entropy is associated with strongly ordered substances.
Equivalence point: The equivalence point in a titration experiment
is reached when the reaction between the titrant (added from the
burette) and the titrate (in the flask) is just complete.
First law of thermodynamics: This states that the total energy of
the Universe is constant; energy can neither be created nor
destroyed.
Gibbs standard free energy: The standard Gibbs free energy
change for a reaction is related to the standard enthalpy and entropy
changes.
Hess’s law: Hess’s law states that the overall reaction enthalpy is
the sum of the reaction enthalpies of each step of a reaction.
Mean bond enthalpy: An average value that is quoted for a bond
that can occur in different molecular environments.
Molar bond enthalpy: The bond enthalpy is the energy required to
break one mole of diatomic molecule bonds.
Monoprotic: An acid which is able to produce one H+ ion from each
molecule.
Order of a reaction: The power to which the concentration of a
particular reactant is raised in the rate equation.
Overall order of reaction: The sum of the powers to which the
concentrations of all reactants are raised in the rate equation.
Quantitative reaction: A quantitative reaction is one where the
reactants react completely according to ratios in the balanced
stoichiometric equation.
Rate constant: In a rate equation, k is the rate constant and has a
constant value for a given reaction at a particular temperature.
Rate-determining step: The slowest step in a reaction mechanism,
which governs the overall rate.
Rate equation: An equation that displays how the reaction rate
depends on the concentration of each reactant.
Reaction mechanism: The series of simple steps by which a
chemical reaction occurs.
Second law of thermodynamics: The total entropy of a reaction
system and its surroundings always increases for a spontaneous
change.
Standard conditions: The conditions of a reaction process at a
pressure of 1 atmosphere and a specific temperature (298 K; 25 °C).
Standard enthalpy change: The enthalpy change for a reaction in
which reactants and products are considered to be in their standard
states at a specified temperature.
Standard enthalpy of combustion: The enthalpy change when one
mole of a substance is completely burned in oxygen under standard
conditions.
Standard enthalpy of formation: The enthalpy change that occurs
when one mole of a substance is produced from its elements in their
standard states.
Standard solution: A solution with an accurately known
concentration. It can be prepared by weighing a primary standard
and dissolving it in a known volume of solution, or by titrating against
another standard solution.
Standard state: The most stable state of a substance or element
under standard conditions.
Third law of thermodynamics: The entropy of a perfect crystal at 0
K is zero.
Unit 2 Exam-style questions
1 The pH of 0.025 mol l–1 H2SO4(aq) is
A 1.3
B –1.6
C 1.6
D 1.9
[1 mark]
2 Which of the following would not act as a buffer solution?
A Ethanoic acid and potassium ethanoate
B Nitric acid and potassium nitrate
C Ammonia and ammonium nitrate
D Hexanoic acid and magnesium hexanoate
[1 mark]
3 a) A salt of methanoic acid, sodium methanoate is commonly
used as a food preservative
Explain why sodium methanoate solution has a pH greater than
7.
In your answer you should mention the two equilibria involved.
[2 marks]
b) When sodium methanoate is added to methanoic acid, a buffer
solution is formed.
0.50 moles of sodium methanoate was added to 200 cm3 of a
0.25 mol l–1 solution of methanoic acid at 298 K, pKa = 3.75.
Calculate the pH of the buffer solution formed.
[2 marks]
4 Barium carbonate decomposes on heating.
BaCO3(s) → BaO(s) + CO2(g) ΔH° = +266 kJ mol–1.
a) Using the data from the table below, calculate the standard
entropy change, ΔS°, in J K–1 mol–1, for the reaction.
[2 marks]
Substance Standard entropy, S°/J K–1
mol 1
BaCO3(s) 112.0
BaO(s) 72.1
CO2(g) 213.8
b) Calculate the temperature at which the decomposition of
barium carbonate just becomes feasible.
[3 marks]
Unit 3 Organic chemistry and
instrumental analysis
Chapter 17
Molecular orbitals and hybridisation
Key points
• VSEPR cannot explain the bonding in all compounds. Molecular
orbital theory can provide an explanation for more complex
molecules.
• Molecular orbitals form when atomic orbitals combine. The
combination of two atomic orbitals results in the formation of a
bonding molecular orbital and an antibonding orbital.
• Each molecular orbital can hold a maximum of two electrons.
• Molecular orbitals that form by end-on overlap of atomic orbitals
along the axis of the covalent bond are called sigma (σ)
molecular orbitals or sigma bonds.
• Molecular orbitals that form by side-on overlap of parallel atomic
orbitals that lie perpendicular to the axis of the covalent bond are
called pi (π) molecular orbitals or pi bonds.
• The electronic configuration of an isolated carbon atom cannot
explain the number of bonds formed by carbon atoms in
molecules. The bonding and shape of molecules of carbon can
be explained by hybridisation.
• Hybridisation is the process of mixing atomic orbitals within an
atom to generate a set of new atomic orbitals called hybrid
orbitals. These hybrid orbitals are degenerate.
• The bonding in alkanes can be described in terms of sp3
hybridisation.
• The bonding in alkenes can be described in terms of sp2
hybridisation.
• The bonding in alkynes can be described in terms of sp
hybridisation.
Formation of molecular orbitals
When atoms join together to form molecules, the electrons then go
into molecular orbitals. These molecular orbitals have lots of
similarities with atomic orbitals. Molecular orbital theory describes
how atomic orbitals merge and overlap when atoms bond to form
molecules. It shows how the valence electrons are distributed within
molecular orbitals spread across the molecule.
The bonding orbital is where the electrons of different spin spend
most of their time between the two nuclei. The bonding orbitals are
where it is most likely that the electrons will be found. The shape is
determined by quantum mechanics. The bonding molecular orbital
has lower energy than the separate atomic orbitals. This means that
a hydrogen molecule is more stable than hydrogen atoms (Figure
17.1). The number of molecular orbitals is the same as the number
of atomic orbitals that combine. If we have two atomic orbitals, we
must therefore have two molecular orbitals. So bonding orbitals and
antibonding orbitals are formed. Antibonding orbitals have higher
energy than the bonding orbitals and atomic orbitals. They are also a
different shape and do not contain electrons.
Figure 17.1 Bonding in a hydrogen molecule
Sigma and pi bonds
There are three different types of molecular bonding orbitals that
come about by end-on overlapping. These are called sigma (σ)
bonds.
Figure 17.2 shows the three different types of sigma bond formation.
In part (a), the s orbitals are overlapping as seen in a hydrogen
molecule. In part (b), the overlap is between an s orbital and a p
orbital as seen in a hydrogen chloride molecule. In part (c), the
overlap is between two p orbitals as seen in a chlorine molecule.
Figure 17.2 Sigma bond formation
It is also possible for p orbitals to overlap side-on or laterally O. This
type of bond is called a pi (π) bond. Pi bonds occur when atoms
form multiple bonds.
Figure 17.3 Pi bond formation
Pi bonds are weaker than sigma bonds. This is shown by the fact
that the carbon–carbon double bond (612 kJ mol−1) is not twice as
strong as the carbon–carbon single bond (348 kJ mol−1).
Hybridisation
Carbon has an electronic configuration of 1s2 2s2 2p2 in its ground
state. You might expect carbon to form two bonds as it has only two
unpaired electrons. However, there is only a small energy difference
between the 2s and 2p subshells. This makes it easy for an electron
to be promoted from the 2s to the 2p. This then provides four atomic
orbitals with unpaired electrons. Carbon now has a valency of four
and can form four bonds.
Spectroscopy of methane has shown that the molecular orbitals are
all the same shape. This must be due to the mixing of the two types
of orbitals, s and p, to give hybrid orbitals. This is called
hybridisation.
2s 2px 2py 2pz orbitals → 4 sp3 orbitals
The four sp3 hybrid orbitals are degenerate and are identical to each
other (Figure 17.4). In methane, molecular orbitals form between
carbon and hydrogen. This type of hybridisation, sp3 hybridisation,
is found in saturated hydrocarbons. In alkanes all the bonds are
sigma bonds.
Figure 17.4 sp3 hybridisation in alkanes
Figure 17.5 sp2 hybridisation in ethene
Carbon can also form other types of hybridisation. In alkenes, sp2
hybridisation is found. This is where the s orbital is hybridised with
two of the p orbitals, and one p orbital is not involved in the
hybridisation.
2s 2px 2py orbitals → 3 sp2 orbitals
In this form of carbon, an electron is promoted from the 2s orbital
into a 2p orbital. This leads to 2s and two 2p orbitals mixing to form
three sp2 orbitals which are all degenerate and singly occupied
(Figure 17.5). The sp2 orbitals form sigma bonds end-on-end. Each
carbon atom has an electron in one p orbital that is not hybridised.
The unhybridised p orbitals overlap side-on-side to form a pi bond.
Sigma bonds are very efficient at overlapping so have a much
stronger bond between the carbons. Pi bonds are formed from the
unhybridised p orbitals overlapping side-on-side, which is less
efficient so they form a weaker bond. This is why a carbon–carbon
double bond is not double the value of a carbon–carbon single bond.
In ethene all the carbon atoms have sp2 hybridisation. However, this
is not the case in the other members of the alkene homologous
series. In propene, for example, only two of the carbons have sp2
hybridisation as they form one carbon–carbon double bond (sigma
and pi bonds) whereas the other carbon has sp3 hybridisation (only
sigma bonds).
In alkynes a triple bond exists between two carbon atoms. This
indicates that another form of hybridisation is present. This is called
sp hybridisation. It is formed by the mixing of one 2s orbital and
one 2p orbital to form two sp orbitals, leaving two of the 2p orbitals
unhybridised.
2s 2px orbitals → 2 sp orbitals
There are two p orbitals on each carbon that are not hybridised.
These overlap sideways and form two pi bonds. The hybrid sp
orbitals overlap end-on to form sigma bonds.
Table 17.1 summarises the types of hybridisation covered in this
chapter.
Study questions
1 How many sigma and pi bonds are found in pent-2-ene?
[2 marks]
2 Describe the bonding found in butane.
[2 marks]
Chapter 18
Molecular orbitals and colour
Key points
• Molecular orbital theory can be used to explain why organic
molecules are colourless or coloured. Electrons fill bonding
molecular orbitals, leaving higher energy antibonding orbitals
unfilled.
• The highest bonding molecular orbital containing electrons is
called the highest occupied molecular orbital (HOMO). The lowest
antibonding molecular orbital is called the lowest unoccupied
molecular orbital (LUMO).
• Absorption of electromagnetic energy can cause electrons to be
promoted from the HOMO to the LUMO.
• Most organic molecules appear colourless because the energy
difference between the HOMO and the LUMO is relatively large.
This results in absorption of light from the ultraviolet region of the
spectrum.
• Some organic molecules contain chromophores. A chromophore
is a group of atoms within a molecule that is responsible for
absorption of light in the visible region of the spectrum.
• Light can be absorbed when electrons in a chromophore are
promoted from the HOMO to the LUMO.
• Chromophores exist in molecules containing a conjugated
system – a system of adjacent unhybridised p orbitals that
overlap side-on to form a molecular orbital across a number of
carbon atoms. Electrons within this conjugated system are
delocalised. Aromatic molecules and molecules with alternating
single and double bonds have conjugated systems.
• The more atoms in the conjugated system, the smaller the
energy gap between the HOMO and the LUMO. A lower
frequency of light (longer wavelength, lower energy) is absorbed
by the compound. When the wavelength of light absorbed is in
the visible region, the compound will exhibit the complementary
colour.
Colour in organic molecules
Many chemical compounds are coloured because they absorb
visible light. However, most organic molecules appear colourless.
By considering molecular orbital theory, we can explain why some
organic compounds are coloured. Electrons can be found in the
bonding molecular orbitals. The sigma and pi bonding orbitals
contain normal bonding pairs of electrons. The non-bonding orbital
contains lone pairs of electrons. The sigma and pi antibonding
orbitals (σ* and π*, respectively, indicated with an asterisk) usually
have no electrons and so are empty.
Several transitions are possible, each with an electron being excited
from a full orbital into an empty one. The energy from photons is
used to promote electrons from bonding or non-bonding orbitals into
the higher energy antibonding orbitals. When absorptions occur,
electrons are excited and then promoted from the filled orbital (so an
electron in a sigma or pi bonding orbital or from a lone pair in a non-
bonding orbital) into the higher energy antibonding orbital.
Figure 18.1 Electron transitions
Saturated compounds that contain only sigma bonds are colourless.
In such compounds, the sigma bonding orbital (σ) is the highest
occupied molecular orbital (HOMO), and the lowest unoccupied
molecular orbital (LUMO) is the sigma antibonding orbital (σ*). The
transition between these orbitals (as shown in Figure 18.1, labelled
A) is quite large (high energy) and is found to correspond to the
ultraviolet part of the spectrum. No visible light is absorbed and so
the compound is colourless.
With compounds containing simple pi bonds, like ethene and butene,
there is still a large transition to promote an electron from the HOMO
(pi bonding orbital, π) to the LUMO (sigma antibonding orbital, σ*),
as shown in Figure 18,=.a, labelled B, and so these compounds also
absorb in the ultraviolet part of the spectrum and are colourless.
Conjugation and chromophores
Organic molecules which are coloured have alternating double and
single bonds in long chains. This is called conjugation. The
alternating sigma and pi bonds of benzene rings allow the electrons
to be delocalised among a number of carbon atoms. The molecular
orbitals containing the delocalised electrons stretch along the length
of the conjugated system. The greater the number of atoms with
delocalised electrons (the larger the conjugated system), the smaller
the energy difference between the HOMO which contains the
delocalised electrons and the LUMO. So the excited delocalised
electrons will therefore require less energy. When this falls within the
visible part of the electromagnetic spectrum, the compounds appear
coloured.
A chromophore is a group of atoms within a molecule that is
responsible for the absorption of light in the visible range and,
therefore, its colour (so alternating single and double bonds in the
conjugated system). By comparing chromophores, it is possible to
find out about the energy of light that is being absorbed. Vitamin A
has a conjugated system that spans over five carbon–carbon double
bonds (Figure 18.2). So it appears yellow.
Figure 18.2 A vitamin A molecule with chromophore shown in
red
β-carotene (Figure 18.3) is found in carrots, pumpkins, squashes
and sweet potatoes. It has 11 carbon–carbon double bonds as the
conjugated system. It has a greater number of carbon–carbon
double bonds than vitamin A, so it appears orange.
Figure 18.3 A β-carotene molecule with chromophore shown in
red
Lycopene (Figure 18.4) is found in watermelon, pink grapefruit, red
cabbage and tomatoes. It has a conjugated system which spreads
over 22 carbon atoms, with 11 carbon–carbon double bonds. It
appears red.
Figure 18.4 A lycopene molecule with chromophore shown in
red
The colours observed are the complementary colours to those
absorbed by the molecule. If the chromophore absorbs light of one
colour, then the complementary colour is observed. Since violet light
has higher energy than blue or green, when it is absorbed, as in
vitamin A, the colour observed is yellow. The complementary colours
can be determined using the colour wheel on page 20 of the SQA
data booklet. As molecules with more conjugation absorb lower-
energy light (longer wavelength), the greater the degree of
conjugation, the more likely the compound is to have absorbed
green light so a red colour is observed. Similarly, less conjugation
results in compounds appearing yellow. Because organic
compounds absorb photons of energy from the electromagnetic
spectrum, these compounds can be analysed using a UV-VIS
spectrophotometer. Many pH indicators are coloured due to
conjugated electron systems.
Hints & tips
The molecules with more conjugation absorb lower energy light
(longer wavelength) as they have a smaller energy gap between
the HOMO and the LUMO.
Study questions
1 Coloured organic compounds have a system of alternating
double and single bonds. What word is used to describe this type
of system?
[1 mark]
2 Anthracene is colourless, while naphthacene is yellow-orange.
Which of the two compounds is likely to have the greatest degree
of conjugation?
[1 mark]
3 Lycopene and β-carotene are both coloured organic compounds
found in ripened tomatoes. Which compound has the largest
energy gap between the HOMO and the LUMO?
[1 mark]
4 Lycopene is present in tomatoes. What colour of light does
lycopene absorb?
[1 mark]
Chapter 19
Homolytic and heterolytic fission
Key points
• When an organic reaction takes place, bonds in the reactant
molecules are broken and bonds in the product molecules are
made. The process of bond breaking is known as bond fission.
• There are two types of bond fission: homolytic and heterolytic.
• Homolytic fission results in the formation of two neutral radicals.
This occurs when each atom retains one electron from the
covalent bond and the bond breaks evenly. It normally occurs
when non-polar covalent bonds are broken.
• Reactions involving homolytic fission tend to result in the
formation of very complex mixtures of products, making them
unsuitable for organic synthesis.
• Heterolytic fission results in the formation of two oppositely
charged ions. This occurs when one atom retains both electrons
from the σ covalent bond and the bond breaks unevenly. It
normally occurs when polar covalent bonds are broken.
• Reactions involving heterolytic fission tend to result in far fewer
products than reactions involving homolytic fission, and so are
better suited for organic synthesis.
• The movement of electrons during bond fission and bond making
can be represented using curly arrow notation.
During bond breaking or bond fission, the electrons are redistributed
between the two atoms involved. There are two possible ways in
which the electrons can be redistributed: homolytic fission and
heterolytic fission.
Homolytic fission
Homolytic fission is the breaking of a covalent bond to produce
two free radicals. Each atom in the bond leaves with one electron
from the shared pair. A dot is used to represent the unpaired electron
in a radical. The single-headed curly arrow is used to indicate the
movement of a single electron (Figure 19.1). The tail of the arrow
shows the source of the electron, while the head indicates its
destination.
Figure 19.1 Homolytic fission
The free radicals produced are highly reactive as the unpaired
electrons tend to attack other species. This was seen in Higher
Chemistry. When alkanes react with halogens such as chlorine, the
halogen atoms replace the hydrogen atoms. The process involves a
number of steps. In the initiation step (Figure 19.2), the chlorine
absorbs energy supplied by UV light. The chlorine–chlorine bond
breaks to form the free radicals.
Figure 19.2 Initiation step
Then, during the propagation step (Figure 19.3), the highly reactive
chlorine free radicals react with the methane molecules, forming
hydrogen chloride gas and leaving a methyl free radical. This methyl
free radical can then react with a second chlorine molecule, forming
chloromethane and another chlorine free radical.
Figure 19.3 Propagation step
For the reaction to stop, two free radicals must collide and react to
form a molecule. This is called the termination step. There are
several possible combinations of free radicals that can react, as
shown in Figure 19.4.
Figure 19.4 Termination step
Heterolytic fission
Heterolytic fission occurs when a covalent bond breaks and both
electrons in the bond move to one of the atoms. This results in two
oppositely charged ions being formed, as shown in Figure 19.5.
Figure 19.5 Heterolytic fission
The full, double-headed arrow shows the movement of a pair of
electrons. Heterolytic fission is most likely to occur in polar
molecules, with the electrons going to the more electronegative
atom. When the polar bond between a carbon and a halide atom
breaks, heterolytic fission results in the formation of a positively
charged carbon atom. This is due to the halide atom having greater
electronegativity than the carbon atom. The positively charged
carbon atom is known as a carbocation.
If the carbon atom is more electronegative than the other bonded
atoms then it will have a greater attraction for the bonding electrons.
This will result in the formation of a negatively charged carbon atom.
This is called a carbanion.
Hints & tips
A double-headed curly arrow shows that two electrons are being
redistributed whereas a single-headed arrow shows the movement
of one electron. In both instances, the tail of the arrow shows the
source of the electron(s) while the head indicates the destination.
Study questions
1 Write the equation for the homolytic fission of bromine by UV
light using curly arrow notation.
[2 marks]
2 What is the name given to the positive carbon atom formed
during heterolytic fission?
[1 mark]
Chapter 20
Electrophiles, nucleophiles and
reaction types
Key points
• In reactions involving heterolytic bond fission, attacking groups
are classified as nucleophiles or electrophiles.
• Nucleophiles are negatively charged ions or neutral molecules
that are electron rich. Examples are
Cl−, Br−, OH−, CN−, NH3 and H2O.
• Nucleophiles are attracted towards atoms bearing a partial (δ+)
or full positive charge. They are capable of donating an electron
pair to form a new covalent bond.
• Electrophiles are positively charged ions or neutral molecules
that are electron deficient. Examples of electrophiles are H+,
NO2+ and SO3.
• Electrophiles are attracted towards atoms bearing a partial (δ−)
or full negative charge. They are capable of accepting an electron
pair to form a new covalent bond.
• A substitution reaction is one in which an atom or group of atoms
is replaced by another atom or group of atoms.
• An addition reaction is a reaction in which two molecules join
together to make one larger molecule.
• An elimination reaction is one in which a small molecule is
removed from an organic compound.
Nucleophiles and electrophiles
A nucleophile is a lone-pair donor. Nucleophiles are molecules or
negatively charged ions that have at least one lone pair of electrons.
The negatively charged ion or group is then attracted to an electron-
deficient centre, where it donates its electron pair to form a new
covalent bond. Nucleophiles will seek out electron-deficient sites in
organic molecules. Examples of common nucleophiles are H2O,
NH3, OH− and the halide ions.
An electrophile is a substance which can accept a pair of electrons.
Electrophiles are positively charged ions or neutral molecules that
are electron deficient. The electrophile molecule is attracted towards
an area of high electron density with a negative or partial negative
charge (δ−). Examples of electrophiles are H2O+, NO2+ and SO3.
Polar molecules such as ammonia have both nucleophilic and
electrophilic centres as they contain areas which are both positively
and negatively charged. In ammonia, the nitrogen is nucleophilic as
it has partial negative charge and so is electron rich while the
hydrogen atoms are electrophilic as they have a partial positive
charge and so are electron deficient.
Figure 20.1 An ammonia molecule
Types of reactions
In a substitution reaction, an atom or group of atoms is replaced
by another atom or group of atoms. For example:
In an addition reaction, two molecules join together to make one
larger molecule. For example:
In an elimination reaction, a small molecule is removed from an
organic compound. For example:
Study questions
1 Identify the type of reaction in which 1,2-dibromobutane is
formed from but-1-ene and bromine.
[1 mark]
2 Identify the nucleophile centre on a carbonyl group in propanone.
[1 mark]
Chapter 21
Alkenes
Key points
• Alkenes are a homologous series of unsaturated hydrocarbons.
• Alkenes can be prepared by dehydration of alcohols using
aluminium oxide, concentrated sulfuric acid or concentrated
phosphoric acid. They can also be prepared by base-induced
elimination of hydrogen halides from monohaloalkanes.
• Markovnikov’s rule is that when a hydrogen halide or water is
added to an asymmetric alkene, the hydrogen atom becomes
attached to the carbon with the most hydrogen atoms attached to
it already. During the reaction of a hydrogen halide or water with
alkenes, Markovnikov’s rule is used to predict the major and
minor products formed.
• Alkenes take part in electrophilic addition reactions with:
• hydrogen to form alkanes in the presence of a catalyst
• halogens to form dihaloalkanes.
• They can also react with hydrogen halides via an addition
reaction to form monohaloalkanes.
• They undergo a hydration reaction with water using an acid
catalyst to form alcohols. This is also an addition reaction.
• The reaction mechanisms for the addition of a hydrogen halide
and the acid-catalysed addition of water can be represented
using curly arrows. These mechanisms also show the
intermediate carbocation. The inductive stabilisation of
intermediate carbocations formed during these reactions is used
to explain the products formed.
• The reaction mechanism for the addition of a halogen can be
represented using curly arrows.
Alkenes are unsaturated hydrocarbons containing at least one
carbon–carbon double bond, C=C. The double bond consists of both
a pi bond and a sigma bond. The chemical reactivity of alkenes is
due to the pi electrons being open to attack from electrophiles.
Alkenes are obtained industrially either by cracking the naphtha
fraction from crude oil or cracking the corresponding alkane. Alkenes
are used as feedstocks for other chemicals, such as
pharmaceuticals and polymers.
Preparation of alkenes
Alkenes can be prepared by several methods.
Alcohols can be dehydrated using concentrated sulfuric acid as a
catalyst to form an alkene. For example, propan-1-ol can be
dehydrated to form propene:
Figure 21.1 Apparatus used to form ethene from ethanol
In the lab, ethanol can be dehydrated using porcelain chips to form
ethene (Figure 21.1).
Monohaloalkanes can undergo an elimination reaction to form an
alkene. The solvent used must be hot ethanol so that elimination is
promoted over nucleophilic substitution. Potassium hydroxide
solution is dissolved in ethanol; the hydroxide ion acts as a base and
accepts a hydrogen ion to form water. This is where the haloalkane
loses a hydrogen atom and a halogen atom and forms an alkene:
Markovnikov’s rule
Addition reactions can produce isomers when the alkene is
asymmetric (in other words, when the double bond is not
symmetrically placed in the molecule). Markovnikov’s rule can be
used to predict the major product of this type of reaction. It states
that ‘the H atom of H–X adds to the carbon atom which already has
more hydrogen atoms’.
If we take but-1-ene and react it with hydrogen bromide, two
products are possible: 1-bromobutane and 2-bromobutane. This is
because either carbon atom in the double bond can be protonated in
the first step. This then leaves the bromide ion free to bond to either
of the resulting carbocations. However, Markovnikov’s rule states
that the major product will be the one where the hydrogen is added
to the carbon with most hydrogens attached, so 2-bromopropane will
be the major product (approximately 90%) and 1-bromopropane will
be the minor product, as shown in Figure 21.2.
Figure 21.2 Two products are formed when hydrogen bromide
reacts with but-1-ene
Hints & tips
When you write mechanisms with nucleophiles, the lone pair must
be shown.
Reactions of alkenes
Hydrogenation
Alkenes can undergo addition reactions with hydrogen to form
alkanes. So ethene will react with hydrogen to form ethane. This
reaction, also known as a hydrogenation reaction, is catalysed by
nickel or platinum.
Halogenation
Halogenation is when a halogen is reacted with an alkene in an
addition reaction. So reacting bromine with ethene will form 1,2-
dibromoethane.
Hydration
Hydration is an addition reaction where water is added across the
carbon–carbon double bond of an alkene to form an alcohol. This
reaction is catalysed by acids, such as phosphoric acid.
Ethanol can be formed by the reaction of steam with ethene in the
presence of concentrated phosphoric acid (Figure 21.3). Acids form
hydronium ions in water:
Figure 21.3 Acid catalysed hydration of an alkene (ethene)
Hints & tips
Addition of water to an alkene using an acid catalyst follows
Markovnikov’s rule.
The hydronium ion is an electrophile which attacks the pi electrons in
the carbon–carbon double bond of any alkene (Step 1 of Figure
21.3). The concentration of water molecules in an aqueous solution
is many times greater than that of the negative ions from the acid
and so the carbocation formed in the first step is much more likely to
react with a water molecule (Step 2). This produces a protonated
alcohol molecule, which readily loses a hydrogen ion to form the
alcohol (Step 3).
Hydrohalogenation
Hydrohalogenation is where a hydrogen halide is added to an
alkene to form a monohaloalkane. So if hydrogen bromide is reacted
with but-2-ene it produces 2-bromobutane. Hydrogen bromide is a
polar electrophile. The reaction takes place at a suitable rate at room
temperature. It proceeds via the method shown in Figure 21.4.
Figure 21.4 Hydrogen bromide reacts with but-2-ene
Carbocations can be classified as primary, secondary and tertiary
depending on the number of alkyl groups, –R, attached (Figure
21.5).
The alkyl groups, such as methyl or ethyl, have an electron-donating
effect. So they push away electrons away from the group towards
the opposite end of the bond. Electron-donating groups reduce the
size of the positive charge on the carbon and so the more alkyl
groups, the more stable the carbocation: a tertiary carbocation is
more stable than a secondary carbocation, which is more stable than
a primary carbocation.
Hints & tips
Alkyl groups such as methyl have an electron-donating effect,
which is the opposite effect to that of an electronegative atom such
as chlorine.
Figure 21.5 Stability of carbocations
The relative stability of the carbocation intermediate affects the
product formed. The more stable carbocation remains for a longer
time in the reaction mixture and so has a greater chance of reacting
with an anion to form a product.
For example, when propene reacts with hydrogen bromide, it forms
two possible carbocation intermediates: in one intermediate, the
carbocation is attached to only one alkyl group; in the second
intermediate, the carbocation is attached to two alkyl groups (Figure
21.6). The more stable carbocation is the secondary carbocation.
The bromide ion preferentially attaches to the secondary carbocation
so 2-bromopropane is the major product (similar to when using
Markovnikov’s rule).
Figure 21.6 Electrophilic addition of hydrogen bromide to
propene
Key links
Unit 3 Chapter 23: Alcohols
Unit 3 Chapter 22: Haloalkanes
Study questions
1 a State the name of the type of mechanism used for the reaction
of ethene with hydrogen bromide.
[1 mark]
b What is the name given to the organic intermediate produced
in this reaction?
[1 mark]
2 Draw and label the apparatus required for the formation of
ethene from ethanol in the lab.
[2 marks]
Chapter 22
Haloalkanes including SN1 and SN2
mechanisms
Key points
• Haloalkanes (alkyl halides) are substituted alkanes where one or
more of the hydrogen atoms is substituted with a halogen atom.
• Monohaloalkanes contain only one halogen atom. They can be
classified as primary, secondary or tertiary according to the
number of alkyl groups attached to the carbon atom bonded to
the halogen atom.
• They take part in elimination reactions to form alkenes using a
strong base, such as sodium or potassium hydroxide, in an
ethanol solvent.
• Monohaloalkanes take part in nucleophilic substitution reactions
with:
• aqueous alkalis to form alcohols
• ethanolic cyanide to form nitriles (chain length increased by
one carbon atom) that can be hydrolysed to carboxylic acids
• alcoholic alkoxides to form ethers.
• Monohaloalkanes undergo nucleophilic substitution reactions by
one of two different mechanisms.
• SN1 is a nucleophilic substitution reaction with one species in
the rate-determining step. This occurs in a minimum of two
steps via a trigonal planar carbocation intermediate.
• SN2 is a nucleophilic substitution reaction with two species in
the rate-determining step. This occurs in a single step via a
single five-centred, trigonal bipyramidal transition state.
• The reaction mechanisms for SN1 and SN2 reactions are drawn
using curly arrows to represent the movement of electrons. Steric
hindrance and the inductive stabilisation of the carbocation
intermediate explains which mechanism will be preferred for
primary, secondary or tertiary haloalkanes.
Haloalkanes
Haloalkanes are a homologous series of saturated carbon
compounds containing one or more halogen atoms. They are used
as anaesthetics, pharmaceuticals, flame retardants and solvents.
Haloalkanes are named in a similar fashion to branched alkanes by
the use of numbers to indicate the position of the halogen atom. This
is illustrated in Figure 22.1.
Figure 22.1 Examples of haloalkanes
Figure 22.2 Examples of primary, secondary and tertiary
haloalkanes
Haloalkanes can be classed as primary, secondary or tertiary. This is
dependent upon how many alkyl groups are attached to the carbon
atom that is bonded to the halogen atom. The haloalkanes shown in
Figure 22.2 are classed as primary, secondary and tertiary,
respectively.
Haloalkanes are polar covalent molecules. They contain a halogen
atom which has greater electronegativity than the carbon atom it is
attached to. So this bond is polarised, with the halogen atom having
a slightly negative charge and the carbon atom having a slightly
positive charge. Haloalkanes can react by having this carbon–
halogen bond broken. Iodoalkanes have the weakest bond, so react
the fastest. Fluoroalkanes have the strongest bond, so react the
slowest. The order of reactivity of the haloalkanes is summarised a
iodoalkanes > bromoalkanes > chloroalkanes > fluoroalkanes
Reactions of haloalkanes
Haloalkanes undergoes two main types of reactions: elimination
and substitution.
Elimination reactions
In elimination reactions, haloalkanes form alkenes using a strong
base, such as potassium or sodium hydroxide, dissolved in hot
ethanol solvent. For example, when ethanolic potassium hydroxide is
added to 2-bromopropane and refluxed, it will form an alkene
(propene), water and bromide ions (Figure 22.3).
When dissolved in ethanol, the hydroxide ions act as a base and
accept a proton to form water. The solvent, ethanol, promotes the
elimination process over the nucleophilic substitution which would
occur if aqueous solvent was used.
Figure 22.3 Mechanism to show an elimination process
Hints & tips
A hot ethanol solvent promotes elimination whereas warm aqueous
conditions promote nucleophilic substitution.
Substitution reactions
There are two different types of substitution reactions: SN1
reactions and SN2 reactions (Figure 22.4).
Figure 22.4 Nucleophilic substitution
In the case of nucleophilic substitution reactions, a nucleophile
attacks the slightly charged positive carbon atom. The nucleophiles
used tend to be hydroxide ions (OH−), cyanide ions (CN−) and
ammonia (NH3). The first two nucleophiles are negatively charged
while the ammonia is neutral. The negatively charged nucleophiles
react via the mechanism shown in Figure 22.5, where Nu represents
the nucleophile.
Figure 22.5 General mechanism for tertiary haloalkane
nucleophilic substitution, SN1
In the SN1 reaction mechanism, there are two steps: a slow step and
a faster step. In the first step, the carbon–halogen bond breaks
heterolytically to form a negative halogen atom and a carbocation
intermediate. A nucleophile can attack the carbocation intermediate
from either side to form the product. This slow step is also called the
rate-determining step. The transition state (carbocation
intermediate) is trigonal planar with three groups attached to the
carbon atom.
The second step is fast so it is not included in the rate order. The
reaction is, therefore, first order overall. This mechanism is least
likely for primary haloalkanes and most likely for tertiary haloalkanes.
In the SN2 reaction mechanism (Figure 22.6), the nucleophile forms
a bond with the carbon atom attached to the halogen atom. At the
same time the carbon–halogen bond breaks, resulting in a transition
state. In this instance, the transition state will be trigonal bipyramidal,
with five groups attached to the carbon atom. As two molecules are
involved in the rate-determining step, the reaction is second order
overall. The nucleophile will have to attach to the opposite side to the
departing hydrogen. SN1 mechanisms are most likely for primary
haloalkanes and least likely for tertiary haloalkanes.
Figure 22.6 General mechanism for primary haloalkane
nucleophilic substitution, SN2
Hints & tips
When you draw the transition state, be very careful to make a clear
difference between the dotted lines showing the half-made and half-
broken bonds.
The nature of the haloalkane and the stability of the corresponding
carbocation will influence the mechanism chosen.
When a nucleophile attacks a primary haloalkane, it approaches the
positive carbon atom from the side away from the halogen atom.
This is impossible with a tertiary haloalkane, as the back of the
molecule is completely cluttered with alkyl groups. This means that
this type of attack is unlikely to occur, which explains why tertiary
haloalkanes undergo a two-step process using the SN1 mechanism.
This effect is known as steric hindrance.
Key link
Unit 2 Chapter 16: Kinetics
If a primary haloalkane was to proceed via the SN1 reaction
mechanism, it would form a carbocation. This carbocation is much
more energetically unstable than the carbocation produced by
tertiary haloalkanes. This is due to the fact that the alkyl groups are
said to have a positive inductive effect. This means the alkyl groups
can push electrons towards the positively charged carbon atom, so
stabilising the carbocation in a process known as inductive
stabilisation. The order of stability of carbocations is:
primary < secondary < tertiary
Hints & tips
Alkyl groups such as methyl have an electron-donating effect which
is the opposite effect to that of an electronegative atom such as
chlorine.
Formation of alcohols
When refluxed with an alkali, a monohaloalkane undergoes
nucleophilic substitution to form the corresponding alcohol. So when
chloromethane is reacted with hot aqueous sodium hydroxide
solution, the hydroxide ion from the alkali acts as a nucleophile and
attacks the carbon attached to the halogen. The two bonding
electrons in the carbon–halogen bond then move on to the halogen
atom and a halide ion forms. This method shows how alcohols can
be formed from haloalkanes. In this example, methanol would be
formed from chloromethane (Figure 22.7).
Figure 22.7 The nucleophilic substitution of chloromethane by
hydroxide ions
Formation of nitriles
Monohaloalkanes can also undergo nucleophilic substitution
reactions with ethanolic cyanide to form nitriles. As a result of such a
reaction, the chain length is increased by one carbon atom. These
nitriles can then be hydrolysed to form carboxylic acids. When
iodomethane is gently refluxed with sodium cyanide dissolved in a
small volume of ethanol, it produces the nitrile, ethanenitrile, and
sodium iodide (Figure 22.8).
Figure 22.8 The nucleophilic substitution of iodomethane by
cyanide ions
Formation of ethers
When a monohaloalkane is reacted with alcoholic alkoxides, it forms
ethers.
Alkoxides are formed by adding an alkali metal, such as sodium or
potassium, to an alcohol such as methanol or ethanol. For example
when sodium is added to methanol, sodium methoxide is formed in
the following reaction:
The methoxide ion can then act as a nucleophile. So if iodopropane
was refluxed gently with sodium methoxide in methanol, the ether
methyloxypropane would be formed (Figure 22.9).
Figure 22.9 The nucleophilic substitution of iodopropane by
methoxide ions
Ethers are discussed later on in this unit (Chapter 25).
Study questions
1 Which of the following is most likely to undergo hydrolysis via an
SN1 reaction mechanism?
A CH3CH2I
B CH3CH2Cl
C (CH3)3Cl
D CH3CHClCH2CH3
[1 mark]
2 When 1-bromobutane is reacted with potassium methoxide, what
is the name of the ether formed?
[1 mark]
3 What is the type of reaction that occurs when a bromoalkane
reacts with potassium hydroxide dissolved in ethanol?
[1 mark]
Chapter 23
Alcohols
Key points
• Alcohols are substituted alkanes in which one or more of the
hydrogen atoms is replaced with a hydroxyl functional group.
• Hydroxyl groups make alcohols polar, which gives rise to
hydrogen bonding. Hydrogen bonding explains the boiling points,
melting points, viscosity and solubility, or miscibility in water of
alcohols.
• Alcohols can be prepared from:
• haloalkanes by substitution
• alkenes by acid-catalysed hydration (addition)
• aldehydes and ketones undergoing reduction using a reducing
agent. A suitable reducing agent is lithium aluminium hydride.
• Alcohols can be dehydrated to form alkenes using aluminium
oxide, concentrated sulfuric acid or concentrated phosphoric acid.
• Oxidation of primary alcohols forms aldehydes and then
carboxylic acids. Secondary alcohols undergo oxidation to form
ketones. Oxidising agents used could be acidified permanganate,
acidified dichromate or hot copper(II) oxide.
• Alcoholic alkoxides can be formed by reaction with alkali metals
such as potassium or sodium. These can then be reacted with
monohaloalkanes to form ethers.
• Alcohols form esters by reaction with carboxylic acids using
concentrated sulfuric acid or concentrated phosphoric acid as a
catalyst.
• Alcohols can also form esters by reaction with acid chlorides.
This method gives a faster reaction than reaction with carboxylic
acids, and no catalyst is needed.
The effect of hydrogen bonding
on the properties of alcohols
Alcohols have higher boiling points than alkanes of a similar relative
molecular mass due to the presence of hydrogen bonding between
the alcohol molecules (Figure 23.1). Alkane molecules have only
London dispersion forces between their molecules. Hydrogen
bonding exists between alcohol molecules due to the polar hydroxyl
group. As hydrogen bonding is stronger than London dispersion
forces, extra energy is needed to overcome the intermolecular forces
of attraction between the alcohol molecules, leading to higher boiling
points.
Figure 23.1 Hydrogen bonding between ethanol molecules
In short-chain alcohols, such as methanol and ethanol, the hydroxyl
group has a considerable effect on the solubility of the molecule.
Both are soluble in all proportions in water while alkanes are
insoluble. This is because the hydroxyl group can hydrogen-bond
with the water molecules. Starting with heptan-1-ol, longer-chain
alcohols are insoluble in water. As the chain length increases, the
solubility decreases as the long non-polar hydrocarbon part of the
molecule becomes dominant over the polar hydroxyl group.
The viscosity of alcohols also increases as the size of the molecules
increases. This is because the strength of the intermolecular forces
increases as the size increases, so more intermolecular forces are
holding the molecules more firmly in place.
Alcohols are substances which are isomeric with ethers. (Isomerism
is covered in greater depth in Chapter 28.)
Preparation of alcohols
Alcohols can be prepared in several different ways:
• Alcohols can be made from monohaloalkanes via nucleophilic
substitution by refluxing with aqueous sodium hydroxide solutions.
Here the hydroxide ion from the alkali acts as a nucleophile.
R–CH2X + NaOH → R–CH2OH + NaX
• Alcohols can also be prepared by hydration reactions of alkenes
(Figure 23.2). Alkenes are refluxed with steam in the present of
acid catalysts such as sulfuric acid. This can also be considered
an addition reaction. Industrial ethanol is produced via this
process.
Figure 23.2 Hydration of ethene
• Aldehydes and ketones can be reduced to alcohols using a
reducing agent. A suitable reducing agent is lithium aluminium
hydride (LiAlH4) dissolved in ethoxyethane. In equations, the
reducing agent is represented by [H]. An aldehyde is reduced to
form a primary alcohol. A ketone is reduced to form a secondary
alcohol. These processes are shown in Figures 23.3 and 23.4.
Figure 23.3 Reduction of aldehyde to primary alcohol
Figure 23.4 Reduction of ketone to secondary alcohol
Reactions of alcohols
Alcohols can undergo a number of different reactions:
• Reactive metals such as sodium and potassium (group 1 alkali
metals) can be reacted with alcohols to form alcoholic alkoxides.
This is classified as a displacement or redox reaction.
• Alcohols can undergo an elimination reaction to form an alkene
(Figure 23.5). This is also classed as dehydration, as a molecule
of water is eliminated from the alcohol to produce the alkene.
Figure 23.5 Dehydration of an alcohol to form an alkene
• Alcohols are dehydrated by using concentrated sulfuric acid as a
catalyst at a temperature of 170 °C or by passing the alcohol
vapour over a heated aluminium oxide catalyst (Al2O3) at 600 °C
(Figure 23.6)
Figure 23.6 Apparatus used for the dehydration of alcohol to
form alkene using an aluminium oxide catalyst
• When alcohols react with carboxylic acids or acid chlorides they
form esters (Figure 23.7). These are classified as esterification or
condensation reactions. When reacting carboxylic acid with an
alcohol, a concentrated sulfuric acid or concentrated phosphoric
acid catalyst is required. This method is much slower than the
reaction of the acid chloride.
Figure 23.7 Ester formation using carboxylic acid
• The reaction of alcohol with an acid chloride does not require a
catalyst and is much faster than the carboxylic acid method (Figure
23.8).
Figure 23.8 Ester formation using acid chloride
• During condensation reactions, a small molecule is lost. This small
molecule does not always have to be a water molecule; for
example, hydrogen chloride is lost in the condensation reaction of
alcohol and acid chloride.
Key links
Unit 3 Chapter 25: Ethers
Unit 3 Chapter 21: Alkenes
Study questions
1 Which reagent can be used to convert an aldehyde into an
alcohol?
A Aluminium oxide
B Tollens’ reagent
C Chromium(III) ions.
D Lithium aluminium hydride
[1 mark]
2 A molecule of propanone was reacted with lithium aluminium
hydride. What would be the name of the final product?
[1 mark]
3 State the type of reaction taking place when 2-bromopropane
reacts with sodium hydroxide.
[1 mark]
Chapter 24
Carboxylic acids
Key points
• Carboxylic acids contain the carboxyl functional group.
• They are weak acids which dissociate partially in solution.
• Carboxylic acids can be prepared by:
• oxidising primary alcohols using acidified permanganate,
acidified dichromate or hot copper(II) oxide
• oxidising aldehydes using acidified permanganate, acidified
dichromate, Fehling’s solution or Tollens’ reagent
• hydrolysing nitriles, esters or amides.
• The reactions of carboxylic acids include:
• the formation of salts by reactions with metals or bases
• condensation reactions with alcohols to form esters in the
presence of concentrated sulfuric or concentrated phosphoric
acid
• reactions with amines to form alkylammonium salts that form
amides when heated
• reduction with lithium aluminium hydride to form primary
alcohols.
Carboxylic acids contain the carboxyl functional group (Figure
24.1).
Figure 24.1 The carboxyl functional group
The carboxyl group is a combination of a carbonyl group (C=O) and
a hydroxyl group (O–H).
Carboxylic acids are weak acids which dissociate partially in
solution. Many carboxylic acids are found in nature, such as citric
acid in oranges and lemons, methanoic acid in ant stings and
ethanedioic acid (oxalic acid) in rhubarb.
Short-chain carboxylic acids are very soluble in water. This is due to
them having highly polar carbonyl and hydroxyl groups which can
form hydrogen bonds with water. As the non-polar hydrocarbon
chain length increases, the solubility in water decreases. Carboxylic
acids are isomers of esters.
Preparation of carboxylic acids
Aldehydes can be oxidised into carboxylic acids. When warmed
with acidified potassium dichromate(VI), ethanal will be oxidised to
ethanoic acid. Other suitable oxidising agents are acidified
permanganate, Fehling’s solution or Tollens’ reagent.
Carboxylic acids can also be prepared by the oxidisation of primary
alcohols. Suitable oxidising agents are acidified permanganate,
acidified dichromate or hot copper(II) oxide. The oxidising reaction is
hard to stop at the aldehyde state so a mixture of carboxylic acid and
aldehyde can sometimes be obtained. The mixture can usually be
distilled to separate the carboxylic acid from the aldehyde. The
carboxylic acid will be the main product after prolonged heating
under reflux.
Figure 24.2 Oxidation of a primary alcohol
A third way of preparing carboxylic acids is by hydrolysing nitriles.
The nitrile is reacted with an aqueous acid and heated under reflux
until it undergoes hydrolysis. The hydrogen ions from the acid
catalyse this reaction.
Hydrolysing esters can also form carboxylic acids. An aqueous acid
or alkali is used as a catalyst while heating under reflux.
When esters are hydrolysed in acidic or basic solution, an alcohol is
always formed. The carboxylic acid is only formed in acidic
conditions (Figure 24.3) whereas the carboxylate ion is formed under
basic conditions (Figure 24.4). The carboxylic acid is easily formed
from the carboxylate ion by reacting it with a strong acid such as
hydrochloric acid.
Figure 24.3 Acid hydrolysis of an ester
Figure 24.4 Base hydrolysis of an ester
In acidic conditions, esters are not completely hydrolysed as they
form an equilibrium mixture in which some ester is still present. The
ester is not soluble in water, however, and will form the top layer on
the water-soluble carboxylic acid and alcohol mixture.
Amides can be hydrolysed to form carboxylic acids by heating under
reflux with either an aqueous acid or aqueous alkali as catalyst
(Figure 24.5). The use of an alkali catalyst produces more carboxylic
acid as it shifts the equilibrium position to the right.
Figure 24.5 Hydrolysis of an amide
Reactions of carboxylic acids
Carboxylic acids show all the normal reactions of acids. So with
reactive metals they form a carboxylate salt and hydrogen:
Metal oxides react with carboxylic acids to form a carboxylate salt
and water:
Metal carbonates react with carboxylic acids to form a carboxylate
salt, water and carbon dioxide:
They also react with alkalis to form a salt and water:
These are all neutralisation reactions.
Carboxylic acids react with alcohols to form esters, with either a
concentrated sulfuric acid or concentrated phosphoric acid catalyst.
An equilibrium mixture is formed in which the products are an ester
and water (Figure 24.6).
Figure 24.6 Esterification process
Carboxylic acids react with amines to form alkylammonium salts,
which on heating will form amides:
Carboxylic acids can be reduced to primary alcohols using lithium
aluminium hydride (LiAlH4). This is a strong reduction reaction which
proceeds straight from the carboxylic acid to the primary alcohol.
Figure 24.7 Reduction of carboxylic acid to primary alcohol
Key link
Unit 3 Chapter 23: Alcohols
Study questions
1 The type of reaction occurring when butanoic acid reacts with
potassium hydroxide is
A neutralisation
B substitution
C elimination
D hydrolysis.
[1 mark]
2 Which reagent can be used to convert alkanoic acids into
alkanals?
A Aluminium oxide
B Hot copper(II) oxide
C Chromium(III) ions
D Lithium aluminium hydride
[1 mark]
3 Which of the following will not result in the formation of a
carboxylic acid?
A Oxidation of a primary alcohol with acidified dichromate
solution
B Hydrolysis of a nitrile with sulfuric acid
C Reduction of an aldehyde with lithium aluminium hydride
D Hydrolysis of an ester with concentrated phosphoric acid
[1 mark]
Chapter 25
Ethers
Key points
• Ethers can be regarded as substituted alkanes. This is where a
hydrogen atom is replaced with an alkoxy functional group, –OR.
Ethers have the general structure R’–O–R’’, where R’ and R’’ are
alkyl groups.
• Ethers are named as substituted alkanes. The alkoxy group is
named by adding the ending ‘oxy’ to the alkyl substituent, and this
prefixes the name of the longer carbon chain.
• Ethers are isomers of alcohols.
• Due to the lack of hydrogen bonding between ether molecules,
they have lower boiling points than the corresponding isomeric
alcohols.
• Methoxymethane and methoxyethane are soluble in water.
However, larger ethers are insoluble in water due to their
increased molecular size.
• Ethers are commonly used as solvents since they are relatively
inert chemically and will dissolve many organic compounds.
• Ethers can be prepared in a nucleophilic substitution reaction by
reacting a monohaloalkane with an alkoxide.
Properties of ethers
Ethers can be regarded as substituted alkanes in which a hydrogen
atom is replaced with an alkoxy functional group, –OR. Ethers have
the general structure R’–O–R’’, where R’ and R’’ are alkyl groups. As
a result of this arrangement of two alkyl groups around the oxygen
atom, they are relatively non-polar molecules.
When naming an ether, the smaller alkyl chain, which would normally
end with the -yl ending, is changed to -oxy. The longer alkyl group is
used to name the parent alkane (the second part of the name). This
is shown in Table 25.1.
Table 25.1 Naming ethers
Systematic name Formula
Methoxymethane CH3–O–CH3
Methoxyethane CH3–O–C2H5
Ethoxypropane C2H5–O–C3H7
Ethoxyhexane C2H5–O–C6H13
Propoxybutane C3H7–O–C4H9
Ethers are isomers of alcohols. They have much lower boiling points
than their corresponding isomeric alcohol because they cannot form
hydrogen bonds between their molecules. However, smaller ethers
such as methoxymethane and methoxyethane can form hydrogen
bonds with water molecules and so are soluble in water.
Hints & tips
Remember that alcohols and ethers are isomers of each other.
The shape of ether molecules is very similar to that of water
molecules. This allows for the polarisation of the C-O bond. Larger
ethers are insoluble in water as the non-polar alkyl chains increase
the masking of the C-O bond.
Larger ethers are useful solvents for extracting organic compounds
from aqueous mixtures. Ethers tend to be highly flammable and
volatile. Due to their volatile nature, they are easily removed from
organic compounds by distillation.
Preparation of ethers
When a monohaloalkane is reacted with alcoholic alkoxides it forms
ethers.
Alkoxides are formed by adding an alkali metal, such as sodium or
potassium, to an alcohol such as methanol or ethanol. For example,
when potassium is added to methanol, potassium methoxide is
formed:
The methoxide ion can then act as a nucleophile. So if
chloropropane was refluxed gently with potassium methoxide in
methanol, the ether methyloxypropane would form:
This is classified as a nucleophilic substitution reaction.
Key links
Unit 3 Chapter 23: Alcohols
Unit 3 Chapter 22: Haloalkanes
Study questions
1 When 1-chloropentane is reacted with potassium ethoxide, what
is the name of the ether formed?
[1 mark]
2 Suggest a reagent suitable for reacting with ethanol to form the
ethoxide ion.
[1 mark]
3 Name the ether with the shortened structural formula CH3–O–
C6H13.
[1 mark]
Chapter 26
Amines
Key points
• Amines are organic derivatives of ammonia in which one or more
hydrogen atoms of ammonia have been replaced by an alkyl
group.
• They can be classified as primary, secondary or tertiary based
upon the number of alkyl groups attached to the nitrogen atom.
• Amines react with acids to form salts.
• Primary and secondary amines, but not tertiary amines, have
hydrogen bonding. As a result, primary and secondary amines
have higher boiling points than isomeric tertiary amines.
• Primary, secondary and tertiary amine molecules can hydrogen
bond with water molecules. This explains the solubility of the
shorter-chain-length amines in water.
• Amines like ammonia are weak bases. They dissociate slightly in
aqueous solution. The nitrogen atom has a lone pair of electrons
which can accept a proton from water, producing hydroxide ions.
Primary, secondary and tertiary
amines
In amines, the hydrogen atoms in ammonia are replaced one at a
time by hydrocarbon groups. Amines play an important role in the
creation of amino acids, which are the building blocks of proteins,
and they are found in many chemicals in the body. Many analgesics
are amines and many synthetic dyes, such as methyl orange, are
made from amines.
Amines are classified as primary, secondary or tertiary according to
the number of alkyl groups bonding to the central nitrogen atom
(Figure 26.1).
A primary amine contains one alkyl group attached to the nitrogen
atom as only one hydrogen atom in ammonia is replaced. In a
secondary amine, two alkyl groups are attached to the nitrogen atom
so two hydrogen atoms in the ammonia have been replaced. In a
tertiary amine, three alkyl groups are attached to the nitrogen atom
as all the hydrogen atoms in ammonia have been replaced.
Figure 26.1 The structure of primary, secondary and tertiary
amines
Amines are usually named according to the alkyl group or groups
(arranged in alphabetical order) followed by ‘-amine’. Amines can
also be named using a number to give the position of the carbon to
which the –NH2 group is attached, using the ‘amino’ form of naming.
For example:
• CH3CH2NH2 is called ethylamine or aminoethane
• H3C–NH–CH3 is called dimethylamine or N-methylaminomethane
• CH3CH2NHCH2CH3 is called diethylamine or N-ethylaminoethane
• CH3N(CH3)CH3 is trimethylamine.
Table 26.1 shows that primary and secondary amines have higher
boiling points than tertiary amines of similar molecular mass. This
can be explained by the hydrogen bonding between the polar –NH
groups in primary and secondary amines. Tertiary amines do not
have a hydrogen atom attached to the N atom so are unable to form
hydrogen bonds between their molecules (Figure 26.2).
Figure 26.2 Hydrogen bonding between amine molecules
Short-chain primary and secondary amines are water soluble. They
make hydrogen bonds between their polar –NH group and the
hydroxyl groups in water molecules. Their solubility decreases as the
non-polar hydrocarbon chain length increases.
Tertiary amines with low molecular mass are soluble in water.
Hints & tips
Only primary and secondary amines form hydrogen bonds between
their molecules.
The basic nature of amines
Amines, like ammonia, are weak bases. The lone pair of electrons
on the nitrogen atom of ammonia and amines can accept a proton. A
dative covalent bond forms.
When amines react with water, they dissociate slightly in aqueous
solution. They accept a hydrogen ion from water to produce an
alkylammonium ion and a hydroxide ion. An equilibrium is set up and
the equilibrium position lies on the left-hand side.
This forms an alkaline solution with a low concentration of hydroxide
ions.
Basicity of aliphatic amines depends upon the availability of the lone
pair on the nitrogen atom which is used to bond with the proton. The
relative strengths of amines can be compared by looking at the pKa
values on page 13 of the data booklet. The stronger the base, the
higher the pKa value. So a methylammonium ion is a stronger base
than an ammonium ion. This is due to the alkyl group being electron
donating so there is slightly more electron density on the nitrogen
atom. As a result, the lone pair is more available to accept a proton.
So aliphatic amines increase in base strength as the number of alkyl
groups attached to the nitrogen increases. Aminobenzene is a
weaker base than ammonia as the non-bonding electrons become
delocalised around the ring.
Reactions of amines
Amines will react (just like ammonia) with strong acids to form salts.
These are neutralisation reactions. For example, hydrochloric acid
will react with methylamine to form methylammonium chloride:
Amines react with carboxylic acids to form salts which on heating
lose water and form amides. Ethylamine and methanoic acid react to
form the salt ethylammonium methanoate which, upon heating,
forms ethylamide:
Key links
Unit 2 Chapter 9: Chemical equilibrium and factors affecting
equilibrium
Unit 2 Chapter 11: Strong and weak acids and bases
Study questions
1 Which of the following amines has the lowest boiling point?
A C4H9NH2
B C3H7NHCH3
C N(CH3)3
D CH3NH2
[1 mark]
2 Which of the following bases is the strongest?
A C4H9NH2
B NH3
C C2H5NH2
D CH3NH2
[1 mark]
Chapter 27
Benzene
Key points
• Benzene (C6H6) is the simplest member of the class of aromatic
hydrocarbons.
• The benzene ring is stable due to the delocalisation of electrons
in the conjugated system. The delocalised electrons being
present explains why the benzene ring does not take part in
addition reactions.
• Bonding in benzene can be described in terms of sp2
hybridisation, with sigma and pi bonds, and electron
delocalisation.
• A benzene ring in which one hydrogen atom has been substituted
by another group is known as the phenyl group. The phenyl group
has the formula –C6H5.
• Benzene rings can take part in electrophilic substitution reactions:
• Benzene rings undergo halogenation by reaction with a
halogen using aluminium chloride or iron(III) chloride for
chlorination.
• They also undergo halogenation with aluminium bromide or
iron(III) bromide for bromination.
• Benzene rings take part in alkylation by reaction of a haloalkane
using aluminium chloride.
• Nitration of benzene uses concentrated sulfuric acid and
concentrated nitric acid.
• The sulfonation of benzene is with concentrated sulfuric acid.
The structure of benzene
Benzene was discovered by Michael Faraday in 1825. Scientists
speculated about its structure for many years. Experimental
evidence showed that benzene was rather unreactive. In 1865
Friedrich Kekulé proposed a structure of six carbon atoms joined by
alternating single and double bonds. However, since benzene does
not rapidly decolourise bromine water, it suggests that no carbon–
carbon double bonds are present.
Figure 27.1 Kekulé’s proposed structure of benzene
In 1922 Kathleen Lonsdale used X-ray diffraction to measure the
bond lengths in benzene. She found that all the carbon–carbon
bonds were the same length. This proved that the Kekulé structure
was incorrect.
Benzene has a molecular formula of six carbons and six hydrogen
atoms (C6H6). It is a colourless liquid with a sweet smell. The
standard way of drawing benzene is now a hexagon with a circle
within the middle, as shown in Figure 27.2.
Figure 27.2 Structure of benzene
Each carbon atom in benzene is sp2 hybridised and the three half-
filled sp2 hybrid orbitals form sigma bonds with a hydrogen atom and
their two neighbouring carbon atoms. This leaves one electron
occupying the unhybridised p orbital on each carbon atom. These p
orbitals overlap sideways with an adjacent carbon atom, forming a pi
cloud above and below the benzene ring. The six p electrons spread
over the whole of the ring. They are delocalised and give regions of
electron density above and below the ring. They are shared by all six
carbon atoms.
Figure 27.3 The formation of the delocalised electron structure
of benzene
Hints & tips
The sigma bonds in benzene are formed from the end-on overlap of
the sp2 orbitals and the pi bonds are formed by the p orbitals
overlapping sideways, forming a delocalised cloud above and
below the sigma bonds.
When the benzene replaces a hydrogen atom with another atom or
group, then a benzene derivative is formed (Figure 27.4). A benzene
ring in which one hydrogen atom has been substituted by another
group is known as the phenyl group and has the formula –C6H5.
When the hydrogen atom is replaced by a hydroxyl group, it forms
phenol (C6H5OH). Phenol is a weak acid and can lose a hydrogen
ion as the phenoxide ion formed is stabilised because the negative
charge on the oxygen atom is delocalised around the ring.
Figure 27.4 Compounds based upon benzene
Reactions of benzene
Most functional groups attach to the benzene ring by substitution
reactions. During substitution reactions the pi cloud of electrons is
disturbed but re-forms in the final product. These reactions are
examples of electrophilic substitution reactions.
In halogenation reactions benzene is less reactive than alkenes. A
catalyst is required to polarise the halogen bond.
• Chlorobenzene can be produced by reacting chlorine with
benzene in the presence of an iron(III) chloride catalyst or an
aluminium chloride catalyst.
Formation of the electrophile:
• Bromobenzene can be produced by reacting bromine with
benzene in the presence of an iron(III) bromide catalyst or an
aluminium bromide catalyst.
Formation of the electrophile:
As it approaches the benzene ring, the electrophile (E+) becomes
attracted by the delocalised electron cloud. A pair of electrons from
the ring of delocalised pi electrons forms a bond with the
electrophile, breaking the electron ring. This produces a highly
unstable intermediate which has only a partially delocalised electron
system containing four delocalised electrons. A carbon–hydrogen
bond breaks in the unstable intermediate and the two electrons in
the bond move back into the pi electron system, re-forming the
stable delocalised electron ring. The hydrogen is lost as H+. This
then bonds with the halide ion, regenerating the catalyst.
Figure 27.5 The mechanism of electrophilic substitution
Benzene rings also take part in alkylation by reaction of a
haloalkane, using aluminium chloride:
This is another example of an electrophilic substitution reaction.
The AlCl3 catalyst polarises the carbon–halide bond and forms the
electrophile (Figure 27.6).
The hydrogen is removed by the [AlCl4]− ion which was formed at
the same time as the electrophile. The aluminium chloride catalyst is
re-generated during the second stage.
Figure 27.6 Electrophilic substitution mechanism for the
alkylation of benzene
Nitration of benzene is where a hydrogen is replaced with a nitro
group (–NO2). To nitrate benzene, a mixture of concentrated sulfuric
acid and concentrated nitric acid is heated under reflux at 50 °C. The
electrophile is the nitronium ion, NO2+. It is produced by reacting the
concentrated sulfuric acid with nitric acid in the following reactions:
Figure 27.7 Nitration of benzene
Figure 27.8 Electrophilic substitution mechanism for the
nitration of benzene
A further reaction of benzene is sulfonation with concentrated
sulfuric acid. When benzene is heated under reflux for several hours
with concentrated H2SO4, benzenesulfonic acid is formed. The
electrophile is formed from concentrated sulfuric acid:
Sulfonation of benzene can also be done by reaction with the
reagent oleum. Oleum is concentrated sulfuric acid with sulfur
trioxide added to form the electrophile SO3+.
Figure 27.9 Electrophilic substitution mechanism for the
sulfonation of benzene
Hints & tips
The reaction mechanisms in this chapter are virtually the same as
each other; the only change is the nature of the electrophile.
Figure 27.10 Summary of aromatic reactions
Key links
Unit 3 Chapter 17: Molecular orbitals and hybridisation
Unit 3 Chapter 22: Haloalkanes
Study questions
1 Chlorobenzene, nitrobenzene and ethylbenzene can all be
formed from benzene by
A nucleophilic substitution
B nucleophilic addition
C electrophilic substitution
D electrophilic addition.
[1 mark]
2 Which of the following is an electrophile?
A NH3
B CH3CH2+
C CH3NH2
D HS−
[1 mark]
3 State the conditions and reagents required for the nitration of
benzene.
[1 mark]
Chapter 28
Stereochemistry (structural,
geometric and optical isomers)
Key points
• Molecules that have the same molecular formula but different
structural formulae are called isomers.
• Structural isomers arise when the atoms are bonded together in a
different order in each isomer.
• Stereoisomers are when the order of the bonding in the atoms is
the same but the spatial arrangement of the atoms is different in
each isomer. There are two types of stereoisomers: geometric
and optical.
• Geometric isomers:
• occur when there is restricted rotation around a carbon–carbon
double bond or a carbon–carbon single bond in a cyclic
compound
• have two different groups attached to each of the carbon atoms
that make up the bond with restricted rotation
• can be labelled cis or trans according to whether the
substituent groups are on the same side (cis) or on different
sides (trans) of the bond with restricted rotation
• have differences in physical properties, such as melting point
and boiling point
• have different chemical properties.
• Optical isomers:
• occur in compounds in which four different groups are arranged
tetrahedrally around a central carbon atom (chiral carbon or
chiral centre)
• are asymmetric and are non-superimposable mirror images of
each other
• can be described as enantiomers
• have identical physical properties, except for their effect on
plane-polarised light
• have identical chemical properties, except when in a chiral
environment such as that found in biological systems; in
biological systems only one optical isomer is usually present
• can rotate plane-polarised light; enantiomers rotate the plane-
polarised light by the same degree but in opposite directions
and so are optically active.
• form racemic mixtures when enantiomers are mixed in equal
volumes. A racemic mixture is optically inactive because the
rotational effect on the plane-polarised light cancels out.
Isomerism
Isomerism occurs when there are several ways to organise a given
number of atoms:
• Structural isomers differ by the order in which the atoms in the
molecule are joined together.
• In stereoisomers, the atoms making up the isomers are joined up
in the same order, but still manage to have a different three-
dimensional arrangement in space. There are two forms of
stereoisomerism: geometric isomerism and optical isomerism.
Structural isomers
In structural isomers, the atoms are arranged in a completely
different order. The shape of the carbon backbone may be altered by
the addition of branches. This is seen in the example of butane and
2-methylpropane. These two substances are isomers of each other;
they both have the molecular formula C4H10 but have different
structural formulae.
By moving a functional group, we can also form isomers. This is
shown by the double-bond position in but-1-ene and but-2-ene
(Figure 28.1).
Figure 28.1 But-1-ene and but-2-ene are structural isomers
Isomers can also be formed by alteration of the functional group in a
compound. Carboxylic acids and esters are isomers of each other,
as are ethers and alcohols.
Geometric isomerism
Geometric isomerism, sometimes called cis/trans isomerism, is a
form of stereoisomerism. Such isomers occur when there is
restricted rotation somewhere in a molecule. This often happens if
there is a carbon–carbon double bond. In unsaturated hydrocarbons,
a pi bond between neighbouring carbon atoms prevents rotation
about the bond.
The methyl chains of but-2-ene can be arranged in two different
ways, as shown in Figure 28.2.
Figure 28.2 But-2-ene
These two molecules are not the same. The carbon–carbon double
bond will not rotate. In structure A, the methyl groups are locked on
the same side of the double bond. This is known as the cis isomer. In
structure B, the two methyl groups are locked on opposite sides of
the double bond. This is known as the trans isomer.
Figure 28.3 Cis and trans isomers in rings
Geometric isomerism can also occur when carbon atoms are joined
in rings (Figure 28.3).
In the cis isomer, the two groups point in the same direction relative
to the plane of the ring. In the trans isomer, the two groups point in
different directions relative to the plane of the ring. As mentioned
above, to get geometric isomers you must have restricted rotation,
often resulting from a carbon–carbon double bond or a ring. There
must also be two different groups on the left-hand end of the bond
and two different groups on the right-hand end. It doesn’t matter
whether the left-hand groups are the same as the right-hand ones or
not.
Table 28.1 shows the melting points and boiling points of the cis and
trans isomers of 1,2-dichloroethene and but-2-ene.
It can be seen from the table that the trans isomer has the higher
melting point and the cis isomer has the higher boiling point. The
straighter shape of the trans isomer means that it packs better than
the cis isomer. As a result of the poorer packing in the cis isomer, the
intermolecular forces are not as effective and so less energy is
needed to melt the solid, giving a lower melting point.
The cis isomer has a higher boiling point so there must be stronger
intermolecular forces between the molecules than between trans
isomers. In 1,2-dichloromethane, the cis isomer is a polar molecule
whereas the trans isomer is non-polar. Both molecules contain polar
chlorine–carbon bonds; however, in the cis isomer, they are on the
same side of the molecule. One side of the molecule will therefore
have a slight negative charge while the other side is slightly positive,
making the cis isomer polar. There will be dipole–dipole interactions
as well as London dispersion forces, so the cis isomer requires more
energy to overcome these attractive forces. This will lead to the
boiling point being raised relative to that of the trans isomer.
In but-2-ene, the inductive effect of the methyl groups pushes the
electrons towards the double bond, causing the cis isomer to
polarise. The resulting dipole–dipole interactions, in addition to the
London dispersion forces, mean that the cis isomer requires more
energy to overcome these attractive forces and so the boiling point is
raised.
Optical isomerism
Optical isomers are so named because of their effect on plane-
polarised light. An optically active substance can rotate the plane of
polarisation of plane-polarised light.
Figure 28.4 Light can be plane-polarised by passing it through
a polariser
Optical isomerism occurs in molecules that have a central carbon
atom with four different atoms or groups attached tetrahedrally. A
molecule that exhibits optical isomerism is asymmetrical. It has no
central plane or axis of symmetry. So, two possible tetrahedral
arrangements occur in space. One is the mirror image of the other
and the mirror images cannot be superimposed on each other. The
two mirror-image molecules which are optical isomers are called
enantiomers (Figure 28.5). The carbon atom which has four
different atoms or groups attached is called the chiral centre (or
chiral carbon).
Figure 28.5 Enantiomers
• A solution of one enantiomer rotates the plane of polarisation in a
clockwise direction (to the right). This enantiomer is known as the
(+) form.
• A solution of the other enantiomer rotates the plane of polarisation
in an anticlockwise direction (to the left). This enantiomer is known
as the (−) form.
Mixing equal quantities of the same concentration of two
enantiomers makes an optically inactive mixture. This mixture has no
effect on the plane of polarised light as the two opposite rotations
cancel each other out. This is known as a racemic mixture or
racemate.
When a chiral compound is synthesised in the lab, it forms a mixture
of optical isomers. Many pharmaceutical compounds have optical
isomers. This can lead to difficulties in separating mixtures as optical
isomers have similar physical properties.
Key links
Unit 3 Chapter 21: Alkenes
Unit 3 Chapter 31: Pharmaceutical chemistry
Study questions
1 Lactic acid (CH3CHOHCOOH) is a weak acid which builds up in
the muscles after exercise. It has a chiral centre. Draw the two
optical isomers and mark on the chiral centre.
[1 mark]
2 How could you determine whether a solution of a drug was a
racemate and not a single enantiomer?
[1 mark]
Chapter 29
Experimental determination of
structure using empirical
microanalysis and mass spectrometry
Key points
• In organic chemistry, a number of experimental techniques are
carried out to verify the chemical structure of a substance.
• Elemental microanalysis is used to determine the masses of C,
H, O, S and N in a sample of an organic compound in order to
determine its empirical formula.
• An empirical formula is the simplest ratio of the elements present
in a molecule.
• Elemental microanalysis can be determined from combustion
product masses and percentage product by mass.
• Mass spectrometry can be used to determine the accurate gram
formula mass (GFM) and structural features of an organic
compound.
• In mass spectrometry, a small sample of an organic compound is
bombarded with high-energy electrons. This removes electrons
from the organic molecule, generating positively charged
molecular ions, sometimes known as parent ions. These
molecular ions then break into smaller positively charged ion
fragments and a mass spectrum is obtained showing a plot of the
relative abundance of the ions detected against the mass-to-
charge (m/z) ratio.
• The mass-to-charge ratio of the parent ion can be used to
determine the GFM of the molecular ion and so a molecular
formula can be determined using the empirical formula.
• The fragmentation data from mass spectrometry can be
interpreted to gain structural information.
Empirical formula
The empirical formula of a compound is the simplest whole number
ratio of the different atoms that are present in the compound.
Octene, for example, has the molecular formula C8H16 and hence
its C : H ratio is 1 : 2. The empirical formula of octene would be CH2.
Elemental microanalysis can be used to determine the masses of C,
H, O, S and N in a sample of a compound. From the data obtained, it
is possible to then determine the empirical formula of the compound.
A small sample of a compound which may contain C, H, N and S
(approximately 2 mg) is weighed accurately before being oxidised at
a high temperature in an atmosphere of pure oxygen. Depending on
the compound under analysis, this will produce a mixture of the
gases N2, SO2, CO2 and H2O. These gases are then separated by
gas chromatography. The mass of each component of combustion is
then measured using a thermal conductivity detector. By converting
the mass of each product gas into the mass of the original element,
the masses of C, H, N and S in the compound can be determined by
a simple calculation. If there is oxygen present, then its mass is
given by subtracting each of these masses from the original mass of
the sample. The empirical formula is then determined from the
calculated element masses using the method shown in the following
example.
Hints & tips
Remember that the structural formula of a compound can be the
empirical formula or it may be a multiple of the empirical formula.
For example, if the empirical formula of a compound is HO, which
would have a mass of 17, but its relative molecular mass is 34, we
can work out that the sample has the structural formula H2O2
(hydrogen peroxide).
Example
A sample of the solvent used in a perfume contained 0.60 g of
carbon, 0.15 g of hydrogen and 0.40 g of oxygen. Calculate the
empirical (simplest) formula of the solvent.
Answer
The empirical formula of the solvent is C2H6O.
Mass spectrometry
Mass spectrometry is a technique used for determining the
accurate gram formula mass of a compound. It may also give some
information about the structural arrangement of the compound. In
mass spectrometry, the injected sample is vaporised and then
bombarded with high-energy electrons. This removes electrons from
the molecule, resulting in positively charged molecular ions
(sometimes known as parent ions):
M → M+ + e−
These molecular ions break into smaller positive ion fragments. The
fragments are then accelerated and focused into a beam which is
deflected by a magnetic field. The degree of deflection is dependent
upon the mass/charge (m/z) ratio of the ions. This results in a
spectrum being produced which consists of several lines from both
the original molecule and the molecular ion fragments. The line with
the largest mass/charge ratio is that of the molecular ion, a molecule
which has lost one electron.
Figure 29.1 The mass spectrum of pentan-2-one
Hints & tips
A molecular ion is a whole (unfragmented) molecule which has lost
one electron so gives the molecular mass of the compound. It is
always the furthest along on the right-hand side of the mass/charge
ratio axis.
Figure 29.2 The structural formula of pentan-2one
Figures 29.1 and 29.2 show the mass spectrum and structural
formula of pentan-2-one (CH3COCH2CH2CH3). The peak due to the
molecular ion [CH3COCH2CH2CH3]+, in which one electron has been
removed, can be seen at 86 on the mass/charge ratio axis. This
peak gives the molecular mass of pentan-2-one and therefore the
molecular formula of the compound.
Hints & tips
A peak at m/z = 15 is common in the mass spectra of organic
compounds and is given by the methyl fragment [CH3]+.
Other peaks arise from the smaller ion fragments that are produced
in the process. Table 29.1 suggests some of the groups which could
account for these other peaks.
Table 29.1
Mass difference Suggested group
15 [CH3]+
28 [C=O]+ or [C2H4]+
43 [CH3CO]+
71 [COCH2CH2CH3]+
86 [CH3COCH2CH2CH3]+
Mass spectrometry can be used to identify the chemical composition
of banned drugs in sport, and detect toxins in food and water
samples.
Hints & tips
Remember: always include the brackets and a positive sign when
writing the ion fragment or molecular ion.
Study questions
1 Analysis of a compound shows that it contains 76.0% lead (Pb),
13.0% chlorine (Cl), 2.2% carbon (C) and 8.8% oxygen (O).
Calculate the empirical formula of this compound. To gain full
marks you must show all your working.
[2 marks]
2 A liquid has a gram formula mass of 44 g mol−1 and contains
54.5% carbon, 36.4% oxygen and 9.1% hydrogen. Calculate the
empirical formula and then the molecular formula and, hence,
name the liquid.
[2 marks]
3 A sample of gas was analysed and the molecular ion was
detected at a mass/charge ratio of 28. What gas is the sample
made up of?
A methane
B ethane
C carbon dioxide
D carbon monoxide
[1 mark]
Chapter 30
Experimental determination of
structure using NMR and infrared
Key points
• Infrared spectroscopy can identify certain functional groups in an
organic compound.
• When infrared radiation is absorbed by organic compounds,
bonds within the molecule vibrate (stretch and bend). The
wavelengths of infrared radiation that are absorbed depend on
the type of atoms that make up the bond and the strength of the
bond.
• In infrared spectroscopy, infrared radiation is passed through a
sample of the organic compound and then into a detector that
measures the intensity of the transmitted radiation at different
wavelengths. The absorbance of infrared radiation is measured in
wavenumbers, the reciprocal of wavelength, in units of cm−1.
• Proton nuclear magnetic resonance spectroscopy (proton NMR
or 1H NMR) can give information about the different chemical
environments of hydrogen atoms (protons or 1H) in an organic
molecule. It can also identify how many hydrogen atoms there are
in each of these environments.
• Proton nuclei behave like tiny magnets. So when placed in a
strong magnetic field some align with the field (lower energy) and
the rest align against it (higher energy).
• Absorption of radiation in the radio frequency region of the
electromagnetic spectrum causes the 1H nuclei to ‘flip’ from the
lower to the higher energy alignment. As they fall back from the
higher to the lower energy alignment, the emitted radiation is
detected and this is then plotted on a spectrum.
• In a 1H NMR spectrum the chemical shift, δ, (peak position) is
related to the environment of the 1H atom. This is measured in
parts per million (ppm).
• The area under the peak is related to the number of 1H atoms in
that environment and is often given by an integration curve on a
spectrum. The height of an integration curve is proportional to the
number of 1H atoms in that environment, and so a ratio of 1H
atoms in each environment can be determined.
• The standard reference substance used in proton NMR
spectroscopy is tetramethylsilane (TMS). This is then assigned a
chemical shift value equal to zero.
• NMR spectra can be obtained using low-resolution or high-
resolution NMR.
• High-resolution NMR uses higher radio frequencies than those
used in low-resolution NMR and provides more detailed spectra.
• In high-resolution proton NMR an interaction with 1H atoms on
neighbouring carbon atoms can result in the splitting of peaks into
multiplets. The number of 1H atoms on neighbouring carbon
atoms will determine the number of peaks within a multiplet and
can be determined using the n+1 rule, where n is the number of
1H atoms on the neighbouring carbon atoms.
• Low- and high-resolution NMR spectra can be analysed. It is
possible to sketch a low-resolution NMR spectrum for any given
compound.
Infrared spectroscopy
Pairs of atoms in a molecule are always vibrating about a fixed
position. The frequency of this vibration depends on the atoms that
are bonded together. The greater the mass of the atoms, the lower
the frequency of the vibrations. The stronger the bond, the higher the
frequency of the vibrations. The vibrational frequency of atoms lies in
the infrared (IR) region of the electromagnetic spectrum.
A simple IR spectrometer consists of two beams of light which pass
through a reference and a sample cell. The infrared beam is split in
two. The monochromator selects the wavelength that is passed
through the sample and reference cells. These beams of light go
through the monochromator through the sample and reference cells
to the detector. The output signal is equal to the sample signal minus
the reference signal. The information is converted to an IR spectrum
of percentage transmission against wavenumber. Wavenumber is
the reciprocal of wavelength and has units of cm−1.
The atoms in an organic molecule can absorb IR radiation. IR
radiation causes parts of a molecule to vibrate but not to break. The
wavelengths absorbed depend on the type of chemical bond and the
group of atoms at the end of these bonds. These characteristic
wavelengths can be found on page 14 of the SQA data booklet.
Molecules can vibrate by bending and stretching. Most of these
vibrations will have absorbed certain infrared wavelengths. As a
result, the very specific characteristic wavenumbers in the IR
absorption spectrum can be used to identify functional groups.
Figure 30.1 Stretching and bending of bonds
In IR spectroscopy, discs of sodium chloride or potassium bromide
(‘salt flats’) are often used to hold a thin film of liquid sample in the IR
beam. These are ionic so do not have molecules which vibrate
through bending and stretching.
The area of the spectrum below 1500 cm−1 is known as the
fingerprint region of the spectrum. This part of the spectrum is
unique to the organic molecule and has a very complicated series of
absorptions due to lots of different bending vibrations within the
molecule. It is much more difficult to pick out individual bonds in the
region below 1500 cm−1 than it is in the ‘cleaner’ region at higher
wavenumbers. Different compounds produce a different pattern of
troughs in this part of the spectrum so the fingerprint region can be
used to identify a compound.
The infrared spectra of ethanoic acid and butanoic acid are shown in
Figure 30.2.
Figure 30.2 Infrared spectra of two different carboxylic acids
Both spectra show a broad peak between around 3000 and 2500
cm−1; this indicates the presence of the hydroxylic bond. Page 14 of
the data booklet gives a range from 3500–2500 cm−1 for a hydrogen
bond (O—H) stretch in a —COOH carboxyl group. There is another
peak that is near identical in the two spectra between 1750 and1680
cm−1, indicating the presence of the carbonyl group (C==O). The
data booklet states this as a carbonyl stretch. These two peaks
identify both the compounds as carboxylic acids. Databases now
exist for most molecules, and IR spectra can be produced and
compared to the database.
Nuclear magnetic resonance
Nuclear magnetic resonance (NMR) spectroscopy is a method of
examining the 1H nuclei present in an organic molecule. When these
nuclei are placed in a strong external magnetic field, they behave
like tiny magnets and orientate themselves parallel with the field.
There are two possible spin alignments: aligning with the magnetic
field or against it.
Figure 30.3 The two possible spin alignments of 1H nuclei
When the magnetic nuclei are aligned against the magnetic field, this
is the higher-energy state of the nuclei. The quantity of energy
required to flip the nuclei between the lower- and higher-energy
states is in the radio wave frequency of the electromagnetic
spectrum. So the energy absorbed gives details regarding the
environment in which the nuclei in the molecule are found. Energy
absorbed makes the nuclei flip from the lower-energy state to the
higher-energy state. Energy is emitted when nuclei move back down
to the lower-energy state.
The chemical shift axis on the spectrum is a comparative measure of
the energy needed to flip the spin of the nuclei. It is often
represented by the Greek letter delta (δO) and it is usually measured
in parts per million (ppm). The chemical shift gives information about
the environment in which the 1H nuclei find themselves. The further
to the left, the further ‘down field’ a peak is said to be. The value of
the chemical shift is related to the type of hydrogen environment.
Chemical shift values for 1H in different environments are given on
page 16 of the data booklet.
A compound called tetramethylsilane (TMS) is used as a reference
to which all other 1H nuclei are compared. TMS has four methyl
groups attached to a central silicon atom. It produces one strong
peak at the right-hand side of the spectrum, which is set as the zero
point of chemical shift. TMS does not generally overlap with the
chemical shift of 1H nuclei in most carbon compounds.
Figure 30.4 The structure of tetramethylsilane
Low resolution 1H NMR spectroscopy
The spectra produced in low-resolution NMR gives detailed
information about the hydrogen atoms present in organic molecules.
It will tell how many possible hydrogen atom environments exist, as
well as the number of atoms in each environment.
A peak appears for each different type of hydrogen environment in a
molecule. So if a spectrum has two peaks, there are two different
environments.
The structural formula of ethanal is shown in Figure 30.5 and the
low-resolution spectrum for this molecule is given in Figure 30.6.
Figure 30.5 Ethanal
Figure 30.6 The low-resolution 1H NMR spectrum of ethanal
The three 1H nuclei in the methyl group are all in the same
environment compared to the one 1H nucleus adjacent to the
carbonyl group. The spectrum for ethanal therefore has two peaks.
The ratio of areas under the peaks gives a ratio of the number of
hydrogen atoms in each environment. The difference in heights
(measured using a ruler) can give the ratio. So for ethanal, it is seen
that the peak at chemical shift 2.3 ppm has a ratio 3 times that of the
peak at chemical shift 9.8 ppm. These are often called integration
peaks.
High-resolution 1H NMR spectroscopy
High-resolution NMR results in more detailed spectra being
produced. The major difference seen in a high-resolution spectrum is
the phenomenon known as multiplicity or coupling. The
magnetism of a hydrogen atom in one environment can affect the
magnetism of a hydrogen atom in a different environment on an
adjacent carbon atom. The NMR signals can then split the peaks into
new signals which are termed multiplets (doublets, triplets, quartets,
etc.).
Figure 30.7 An interaction with 1H atoms on neighbouring
carbons can result in the splitting of peaks into multiplets
These split signals indicate the number of hydrogen atoms on
carbon atoms adjacent to the hydrogen atoms responsible for the
signal (Figure 30.7).
The splitting of the peaks into a set of peaks is called the spin–spin
splitting pattern. The splitting pattern follows the n+1 rule. If there
are five peaks (n+1) on a spectrum, this is caused by four (= n)
hydrogen nuclei in adjacent carbon atoms. This idea is summarised
further in Table 30.1.
The high resolution NMR spectrum of ethanol shows three main
peaks (Figure 30.8). There is a triplet at chemical shift 1.2 ppm due
to the methyl protons labelled a. However, the triplet is due to the
adjacent nuclei of two hydrogens on the CH2 labelled c. There is a
singlet at 2.6 ppm for the hydrogen nucleus in the hydroxyl bond.
There is a quartet at chemical shift 3.7 ppm. This quartet is due to
the adjacent nuclei being three hydrogens, labelled a.
Figure 30.8 The high-resolution NMR spectrum of ethanol
Key links
Unit 3 Chapter 22: Haloalkanes
Unit 3 Chapter 23: Alcohols
Study questions
1 The spectrum shown in Figure 30.9 was obtained from a sample
of propanone. Explain why only one peak is observed on the
NMR spectrum.
[2 marks]
Figure 30.9
2 The infrared spectrum of a compound with the molecular formula
C4H10O is shown in Figure 30.10. Explain how the spectrum can
be used to support the idea that it is an alcohol being analysed.
[2 marks]
Figure 30.10
Chapter 31
Pharmaceutical chemistry
Key points
• Drugs are substances that alter the biochemical processes in the
body.
• Drugs that have beneficial effects are used in medicines.
• A medicine usually contains the drug plus other ingredients, such
as fillers to add bulk or sweeteners to improve the taste.
• Drugs generally work by binding to specific protein molecules.
These protein molecules can be found on the surface of a cell
(receptor) or can be specific enzyme molecules within a cell.
• Drugs that act on receptors can be classified as agonists or
antagonists:
• An agonist mimics the natural compound and binds to the
receptor molecules to produce a response similar to the natural
active compound.
• An antagonist prevents the natural compound from binding to
the receptor, and so blocks the natural response from
occurring.
• Many drugs that act on enzymes are classified as enzyme
inhibitors and act by binding to the active site of the enzyme and
blocking the reaction normally catalysed there.
• The overall shape and size of a drug is such that it interacts with
a receptor binding site or to the active site of an enzyme. The
types of interactions formed can include van der Waals forces
and/or ionic bonds.
• The structural fragment of a drug molecule that allows it to form
interactions with a receptor binding site or to an enzyme active
site normally consists of different functional groups correctly
orientated with respect to each other.
• By comparing the structures of medicines with similar
pharmacological activity, the pharmacophore can be identified.
The discovery of drugs
Drugs are substances that alter the biochemical processes of the
body. They are said to be pharmacologically active. Substances that
have a beneficial effect are called medicines. Most medicines
contain a single active ingredient which is usually an organic
compound. Other ingredients can include:
• bulking agents to bulk out the medicine if the mass of the active
ingredient is very small
• sweeteners to enhance the palatability of a medicine
• coating agents to ease swallowing
• preservatives to extend the shelf-life.
Hints & tips
Medicines are drugs which have a beneficial effect on the body.
The first drugs were discovered by chance when people identified
the healing benefits of natural remedies from plants. Willow bark has
been used for centuries to reduce pain and is still used to make the
active ingredient in aspirin.
Some drugs are still discovered by chance but in most cases they
are now discovered by research, which looks at the activity based
upon the drug’s structure.
A new drug is expensive to develop and only around 10% make it to
market. The time to market is usually around ten years from the
research stages. This can cost pharmaceutical companies millions of
pounds, with the vast majority of compounds proving unsuccessful.
Chemists still search the world for plants, flowers and berries that
might contain new medicines. On discovering a suitable compound,
they then isolate and determine the structure of the naturally
occurring drug. This can then serve as a prototype in the design of
other biologically active compounds.
The goal of medicinal chemists is to find compounds that have
beneficial effects on given diseases with minimal side-effects. A drug
must be selective, it must be transported to the correct cells in the
body and it must only react in these selected cells.
How drugs work
Many different types of cells in the body have protein molecules on
the surface. These have receptor sites on them which are of a
specific shape, size and structure with which small, biologically
active molecules can interact. This interaction occurs between
functional groups in the active molecule and functional groups in the
complex protein. These interactions involve weak forces such as
hydrogen bonding and weak electrostatic interactions.
When the correct drug molecule binds to the receptor site, it
activates the cell and triggers a biological response within the cell.
The active molecule then leaves the site without itself being
chemically changed.
Figure 31.1 Drug and receptor interaction
Drugs work by acting on a receptor site either to mimic the response
of the natural active compound or to block the effect of the natural
compound.
They can be split into two types called agonists and antagonists:
• An agonist mimics the natural compound and binds to the receptor
molecules to produce a response similar to the body’s natural
active compound.
• An antagonist prevents the natural compound from binding to the
receptor and so blocks the action of the body’s natural active
compound.
An example of an agonist medicine is morphine (Figure 31.2). Along
with all other opiates, morphine has a chemical structure similar to
endorphins, chemicals found naturally in the brain. Endorphins are
released by the brain to provide relief when the body experiences
pain or stress. When someone takes morphine, the molecule binds
to the receptor sites in the brain and mimics the function of natural
endorphins, thus providing pain relief.
Figure 31.2 The structures of morphine, codeine and heroin
An example of an antagonist is pronethalol. This chemical binds to
the specific receptors in the heart, preventing adrenaline from raising
blood pressure and dilating the blood vessels in the heart.
An enzyme inhibitor is a molecule that binds to an enzyme’s active
site and decreases its activity by blocking the reaction normally
catalysed there.
The active ingredients in a medicine can have several different types
of interactions with the receptor sites. These could be ionic
interactions, as many drugs contain acidic/alkaline groups which
ionise under biological conditions. Alternatively, the interactions
could be van der Waals forces, polar–polar interactions like
hydrogen bonding or non-specific London dispersion forces.
Structures of drugs
By comparing the structures of drugs that have similar effects on the
body, the structural fragment that is involved in the drug action can
be identified. This part of the structure of the molecule is called the
pharmacophore. The pharmacophore of morphine is shown in
Figure 31.3. When compared to the structures of other opiates,
codeine and heroin, in Figure 31.2, it can be seen they all contain
the same structural fragment.
Figure 31.3 Morphine with the pharmacophore highlighted in
red
The overall shape and size of the drug has to be just right so that it
fits a binding site, just as a key fits into a lock. The structural
fragment of a drug molecule which confers pharmacological activity
usually consists of different functional groups correctly orientated
with respect to each other.
Certain molecular features, including a molecule’s size and some
specific functional groups, can influence its toxicity.
Many drug molecules have chiral carbons and both enantiomers can
be present within a medicine. Often, one of the enantiomers has no
activity at all but it can sometimes act as an antagonist or bind to a
totally different receptor, causing side-effects. An example of this is
thalidomide. Thalidomide is a drug which was prescribed for morning
sickness in the 1950s. This drug has two optical isomers: one isomer
is the correct shape and had the desired response; the other isomer
behaved differently and caused major birth defects by interfering with
DNA.
Key link
Unit 3 Chapter 28: Stereochemistry
Study questions
1 Most medicines work by binding to receptors. Receptors are
usually
A electrophiles
B free radicals
C protein molecules
D nucleophiles.
[1 mark]
2 Omeprazole is a drug commonly used to prevent stomach ulcers.
It is described as a proton pump inhibitor as it reduces the ability
of enzymes to produce gastric acid. Explain whether this is an
example of an agonist or an antagonist.
[2 marks]
3 Lidocaine is used as a numbing cream. The maximum safe dose
of lidocaine for a child is 3 mg of lidocaine per kg of body mass.
1.0 cm3 of lidocaine solution contains 5 mg of lidocaine.
Calculate the maximum volume of lidocaine solution that could
be given to a 35 kg child.
[1 mark]
Unit 3 Glossary
Addition reaction: Joining two or more molecules together to form a
larger molecule.
Aliphatic hydrocarbon: A hydrocarbon compound containing
carbon and hydrogen atoms joined together in straight chains,
branched chains or non-aromatic rings.
Alkyl group: A group with the general formula CnH2n+1 obtained by
removing a hydrogen atom from an alkane; usually represented by
R.
Alkylation: Introduction of an alkyl group into an organic molecule.
Antibonding orbital: A molecular orbital outside the region between
the two nuclei.
Aromatic: An organic compound that contains a benzene ring.
Asymmetric carbon atom: A carbon atom in a molecule that is
attached to four different atoms and/or functional groups.
Chiral centre: A carbon atom attached to four non-identical atoms
or groups.
Condensation reaction: The joining of molecules with the release
of water as a product.
Conjugated: Molecules with double or triple bonds that are
separated by one single bond; there is delocalisation of electrons in
the pi (π) orbitals between the carbon atoms linked by the single
bond.
Covalent bond: The sharing of one or more pairs of electrons
between two atoms.
Dehydration reaction: A reaction in which water is eliminated from
a molecule.
Delocalised: Molecules or ions that have p orbitals extending over
three or more atoms have delocalised π electrons.
Displayed formula: The relative positioning of atoms and the
number of bonds between them.
Electron shells: The main energy levels of an atom where the
electrons are located.
Electronegativity: A measure of the tendency of an atom in a
molecule to attract a pair of shared electrons towards itself;
electronegativity values increase from left to right across the Periodic
Table and decrease down a group.
Electrophile: A molecule or cation that can act as an electron pair
acceptor (lone pair or π pair) or Lewis acid in a reaction with an
organic molecule.
Electrophilic addition: An addition reaction initiated by the rate-
determining attack of an electrophile on the π electrons of the
carbon–carbon double bond.
Electrophilic substitution: A substitution reaction in which the new
group introduced into the molecule is an electrophile.
Elimination reaction: When a small group of atoms breaks away
from a larger molecule with the release of a small molecule.
Empirical formula: The simplest ratio of elements present in a
substance.
Enantiomers: A pair of molecules that are non-superimposable
mirror images.
Free radical: A species with an unpaired electron. Represented in
mechanisms by a single dot.
Functional group: An atom or group of atoms (other than hydrogen)
that imparts specific physical and chemical properties to a
homologous series of organic compounds.
General formula: The simplest algebraic formula for a member of a
homologous series.
Ground state: The lowest possible energy state of an atom or
molecule.
Halide: A compound of one of the halogens, group 7.
Haloalkanes: A homologous series of organic compounds in which
one (or more) of the hydrogen atoms of an alkane have been
substituted or replaced by halogen atoms.
Halogenation: Any reaction in which a halogen atom (and no other
element) is introduced into a molecule.
Heterolytic fission: The process of breaking a covalent bond within
a molecule, leading to the formation of ions.
High-resolution NMR: NMR performed in the presence of a strong
and stable magnetic field so that spin–spin coupling can be
observed.
Homologous series: Series of organic compounds with the same
functional group and general formula.
Homolytic fission: The process of breaking a covalent bond within
a molecule, leading to the formation of free radicals.
Hybridisation: The mixing of two or more atomic orbitals to form the
equivalent number of hybrid molecular orbitals (all of identical shape
and energy) which overlap and form covalent bonds.
Hydration: The addition of a water molecule.
Hydrogenation: The addition of hydrogen.
Hydrogen bond: An unusually strong intermolecular interaction that
occurs among molecules possessing permanent dipole moments, for
example, H2O, NH3 and HF.
Hydrolysis reaction: The breakdown of a molecule by water.
Hydrophilic: Used to describe molecules or functional groups that
are soluble in water.
Hydrophobic: Used to describe molecules or functional groups that
are poorly soluble or insoluble in water.
Inductive effect: Effect of a functional group or atom in an organic
molecule which attracts sigma electrons towards itself, or repels
them, resulting in the formation of a dipole in the molecule.
Infrared spectroscopy: A type of absorption spectroscopy carried
out in the infrared region of the electromagnetic spectrum, generally
detecting bond stretching and bending.
Initiation: The first elementary step in a free radical reaction; it
involves the homolytic cleavage of a bond, typically by ultraviolet
radiation or high temperature, to generate free radicals.
Ionic bonding: A strong electrostatic force of attraction between all
the oppositely charged ions arranged into a lattice.
Kekulé structure: A localised description of the structure of
benzene in which there is a six-membered ring with alternate double
and single bonds.
London dispersion forces: The underlying interactions between all
atoms and molecules; they are the sole interactions between noble
gas atoms and between non-polar molecules. They are believed to
occur when the electrons within a molecule induce a temporary
dipole in an adjacent atom or molecule.
Low-resolution NMR: NMR performed in a non-homogeneous
magnetic field, where spin–spin coupling cannot be observed.
Markovnikov’s rule: When adding a hydrogen halide to an
asymmetric alkene, the major product is formed from hydrogen
adding to the carbon with more hydrogens, and halide adding to the
carbon with fewer hydrogens.
Mass spectrometer: An instrument (maintained under a high
vacuum) in which gaseous atoms or molecules are fragmented and
ionised and then accelerated into a magnetic field where the ions are
separated according to their mass-to-charge ratio.
Mechanism: A description in terms of bond breaking, bond making
and intermediate formation of the series of elementary steps by
which an overall chemical reaction occurs; also describes the
movement of electrons.
Molecular formula: The exact number of atoms of each element in
a molecule. The molecular formula is an exact multiple of the
empirical formula.
Molecular ion: Unipositive ion formed by an unfragmented molecule
losing one electron following electron bombardment.
Molecular orbital theory: A theory of chemical bonding based upon
the postulated existence of molecular orbitals.
Molecular orbitals: Formed in molecules when atomic orbitals
combine and merge as atoms bond together; s and π bonds are
molecular orbitals.
Multiplicity: The spin of one nucleus affects that of a chemically
different nucleus on an adjacent atom leading to doublets or triplets
in NMR spectra.
Neutralisation reaction: A neutralisation reaction is when an acid
and a base react to form water.
Nuclear magnetic resonance (NMR): Absorption of radio waves at
a precise frequency by nuclei with an odd nucleon number when in
an external magnetic field.
Nucleophilic substitution: The substitution of an atom or group of
atoms with a nucleophile as the attacking species; can occur via an
SN1 or SN2 mechanism.
Optical isomerism: Occurs when a molecule has no plane of
symmetry and can exist in left- and right-handed forms that are non-
superimposable mirror images of each other. The molecule must
possess a chiral centre; optical isomers rotate plane-polarised light.
Orbital: A region in space in which an electron may be found in an
atom or molecule; each atomic orbital can hold up to a maximum of
two electrons with opposite spins.
Oxidation: Increase in oxygen to hydrogen ratio; loss of electrons.
Oxidising agent: An oxidising agent (oxidant) is an electron
acceptor.
Physical property: A property that can be measured without
changing the chemical composition of a substance, for example,
density and melting point.
Pi bond: A bond formed by the sideways overlap of two p orbitals.
Polar covalent bond: A bond formed when electrons are shared
unequally between two atoms due to a difference in electronegativity.
One atom has a partial positive charge and the other atom has an
equal but opposite partial negative charge. The larger the difference
in electronegativities, the greater the polarity and the larger the
partial or fractional charges.
Polarimeter: A device used to study optically active substances.
Racemic mixture: An equimolar mixture of two enantiomers of the
same compound; as their rotation of plane-polarised light is equal
but opposite, the mixture is not optically active.
Rate-determining step: The slowest elementary step in a reaction
mechanism, which controls the rate of the overall reaction.
Reaction mechanism: A list of all elementary reactions that occur in
the course of an overall chemical reaction.
Reaction pathway: A sequence of reactions involving the
conversion of organic compounds.
Reduction: Decrease in oxygen to hydrogen ratio; gain of electrons.
Reflux: Process of boiling a liquid in a flask connected to a
condenser so that the condensed liquid runs back into the flask.
Saturated: Used to describe an organic molecule, for example an
alkane, that contains no carbon–carbon multiple bonds and contains
only carbon–carbon single bonds.
Sigma bond: Formed by the head-on overlap between atomic
orbitals.
Skeletal formula: A simplified organic formula where hydrogen
atoms are removed from alkyl chains. This leaves just a carbon
skeleton and associated functional groups.
SN1 (mechanism): A nucleophilic substitution in which a carbocation
intermediate is formed in the rate-determining step which then reacts
with the nucleophile.
SN2 (mechanism): A nucleophilic substitution where a concerted
reaction occurs in which the nucleophile begins to bond with the
carbon bearing the halogen as the halogen begins to leave the
molecule.
Stereoisomerism: When two double-bonded carbon atoms each
have two different atoms or groups attached to them; also known as
geometric isomerism.
Steric hindrance: The prevention or slowing down of a reaction by
atoms or functional groups blocking the access of an attacking
molecule or ion.
Structural formula: The minimal detail that shows the arrangement
of atoms in a molecule.
Substitution reaction: Atoms are added to a molecule by replacing
atoms already present.
Tetramethylsilane (TMS): The reference standard for proton (and
carbon) nuclear magnetic resonance.
Ultraviolet/visible spectroscopy: The absorption of
ultraviolet/visible light by a molecule, causing the promotion of an
electron from a ground electronic state to an excited electronic state.
Unsaturated: Used to describe a molecule, such as an alkene,
containing one or more carbon–carbon double bonds.
Van der Waals forces: The weak electrostatic forces of attraction
between molecules.
Wavenumber: Wavenumber is the reciprocal of wavelength and has
the units of cm−1 (number of cycles per cm).
Unit 3 Exam-style questions
1 Which of the following reactions would not produce propanoic
acid?
A Hydrolysis of propanenitrile
B Hydrolysis of methyl propanoate
C Oxidation of propan-2-ol
D Oxidation of propanal
[1 mark]
2 Elemental analysis of an organic compound showed it contained
70.6% carbon, 23.5% oxygen and 5.9% hydrogen by mass.
The structural formula of the compound could be:
3 A compound X containing only carbon, hydrogen and oxygen was
subjected to elemental analysis. Complete combustion of 1.76 g of
X gave 3.52 g of carbon dioxide and 1.44 g of water. No other
product was formed.
a) (i) Calculate the masses of carbon and hydrogen in the
original sample and hence deduce the mass of oxygen
present.
[2 marks]
(ii) Show, by calculation, that the empirical formula of
compound X is C2H4O.
[1 mark]
b) Given that the relative molecular mass of compound X is 88,
deduce its molecular formula.
[1 mark]
4 1-bromobutane and 2-bromo-2-methylpropane are isomers which
can be converted to their corresponding alcohols using hydroxide
ions.
a) The conversion of 2-bromo-2-methylpropane to 2-
methylpropan-2-ol involves an SN1 mechanism. Outline the
two steps in this mechanism using structural formulae.
[2 marks]
b) Explain why the conversion of 1-bromobutane to butan-1-ol is
unlikely to proceed by an SN1 mechanism.
[1 mark]
Unit 4 Researching chemistry
Chapter 32
Errors and significant figures
Key points
• The correct number of significant figures should be used in
chemistry measurements.
• The calculated result needs to be reported with the correct
number of significant figures. This will depend on the number of
significant figures in the individual measurements and on the type
of mathematical operation being carried out.
• Measurements from balances, burettes and other pieces of
laboratory equipment should be reported to the correct number of
decimal places.
• Processed results must be given to the correct number of
significant figures. The number used should be the same as the
measurement with the lowest number of significant figures.
• Calculations should not be rounded off to the correct number of
significant figures until the last step in a calculation.
• Uncertainty is an estimate attached to a measurement which
gives the range of values within which the true value is thought to
lie.
• From uncertainty values, percentage errors can be calculated.
These percentage errors can then be added to determine the
overall error in the procedure.
Significant figures
Significant figures are the meaningful digits in a number. There are
some simple rules to help you determine the number of significant
figures:
• Digits that are not zero are always significant. So, 215 cm3 has
three significant figures and 5.248 g has four significant figures.
• Zeros that lie between non-zero digits are always significant. Thus,
5005 g has four significant figures while 6.01 kg has three
significant figures.
• Zeros at the beginning of a number are never significant as all
they do is set the position of the decimal point. So 0.321 mg has
three significant figures and 0.07 mg has one significant figure.
022, 0022 and 0.0022 all have two significant figures.
• Zeros at the end of a number are always significant if the number
contains a decimal point. Thus, 280.0 cm has four significant
figures and 0.0500 mol l−1 has three significant figures.
• Zeros at the end of a number may or may not be significant if the
number contains no decimal point.
In calculations, you round the answer to a certain number of
significant figures. The rules for rounding are:
• If the next number is 5 or more, round up.
• If the next number is 4 or less, do not round up.
When combining measurements with different degrees of accuracy
and precision, the accuracy of the final answer can be no greater
than that of the least accurate measurement.
Example
1 In a titration, 20.4 cm3 of 0.25 mol l−1 sodium hydroxide reacts
with 0.21 mol l−1 nitric acid. Calculate the volume of nitric acid
required to neutralise the sodium hydroxide solution.
Answer
Step 1: Write down the number of significant figures in each
measurement.
Measurement Number of significant
figures
20.4 cm3 3
0.25 mol l−1 2
0.21 mol l−1 2
Step 2: The number of significant figures in the least accurate
measurement is two, so this is how many significant figures
should be given in the answer.
Step 3: NaOH + HNO3 → NaNO3 + H2O
Number moles of NaOH = c × V = 0.25 × 0.0204 = 0.0051
1 mole of NaOH : 1 mole of HNO3
Step 4: Round answer to the same number of significant figures
as the least accurate, so two significant figures.
Volume = 0.024 l or 24 cm3
Hints & tips
Rounding should be left until the very end of the calculation.
Uncertainty
All measurements have a degree of error associated with them.
Every time a measurement is made, that measurement can only be
carried out to a limited degree of accuracy depending on the
equipment used. This uncertainty in the measurement gives a range
of values within which the actual value lies. All glassware used in
titration, such as a burette or pipette, is manufactured with the
maximum uncertainty marked on the glassware. This can be seen in
Figure 32.1.
Figure 32.1 The uncertainty measurement on a 15 cm3 pipette
is marked ±0.03
A balance that reads to one decimal place will have an uncertainty of
±0.1 whereas a three-decimal-place balance will have greater
accuracy and an uncertainty of ±0.001. It is useful to be able to
calculate the percentage uncertainty in any result. This is often
called the percentage error and is given by the relationship
The percentage errors can then be added together to determine the
overall error in the procedure.
Example
2 What is the percentage error involved in making up a 0.1 mol l−1
solution of calcium carbonate in a 250 cm3 class B standard flask
using a two-decimal-place balance?
Answer
Moles of calcium carbonate = c × V = 0.1 × 0.25 = 0.025 moles
Mass of calcium carbonate = n × GFM = 0.025 × 100 = 2.5 g
Two measurements are made using the balance when weighing
by difference so the uncertainty value is doubled.
Total percentage error = 0.8 + 0.12 = 0.92%
Key links
Unit 4 Chapter 33: Practical skills and techniques
Unit 4 Chapter 35: Volumetric analysis
Unit 4 Chapter 36: Gravimetric analysis
Study questions
1 Calculate the number of moles in 2.2 g of sodium. Give your
answer to the correct number of significant figures.
[1 mark]
2 Write the following numbers to the stated number of significant
figures:
a) 5.378 to 3 s.f.
b) 0.0002512 to 2 s.f.
[2 marks]
3 A burette has an uncertainty of ±0.05 cm3. In a titration, the initial
burette reading was 0.05 cm3 and the final burette reading was
12.5 cm3. Calculate the percentage error in the titration.
[1 mark]
Chapter 33
Practical skills and techniques
Key points
• Weighing by difference is the most accurate way to measure the
mass of reactants and products.
• Desiccators are used to store dried samples in a dry atmosphere.
• Vacuum filtration involves carrying out a filtration under reduced
pressure and provides a faster means of separating a precipitate
from a filtrate. A Büchner, Hirsch or sintered glass funnel can be
used during vacuum filtration.
• Heating under reflux allows heat energy to be applied to a
chemical reaction mixture over an extended period of time without
volatile substances escaping. When carrying out heating under
reflux, the reaction mixture is usually placed in a round-bottomed
flask with anti-bumping granules and the flask is fitted with a
condenser. The flask is then heated using an appropriate source
of heat.
• Distillation is a technique used to purify and identify organic
compounds. Distillation can be used to purify a compound by
separating it from less volatile substances in the mixture.
• Solvent extraction involves isolating a solute from a liquid mixture
or solution by extraction using an immiscible solvent in which the
solute is soluble. The quantity of solute extracted is greater using
a number of extractions with smaller volumes of solvent than a
single extraction using a large volume of solvent.
• The technique of recrystallisation is used to purify an impure
solid. The solvent for recrystallisation is chosen so that the
compound being purified is completely soluble at high
temperatures and only sparingly soluble at lower temperatures.
• A pure compound can be identified by melting point analysis with
reference to the literature melting point value. Determination of
the melting point of a compound can give an indication of the
purity of a compound. The presence of impurities in the
compound lowers the melting point and broadens its melting
temperature range due to the disruption in intermolecular bonding
in the crystal lattice.
• Determination of a mixed melting point involves mixing a quantity
of the product with the pure compound and determining the
melting point range. The melting point value and the range of the
melting temperature is used to determine whether the product
and the pure compound are the same substance.
Weighing by difference
Being able to measure the mass of reactants and products is an
important skill. The most accurate method used is called weighing
by difference. A clean, dry weighing boat or weighing bottle is first
weighed empty and then the sample is added to it. The accurate
mass of the weighing boat and its contents is then measured and
recorded. The sample from the weighing boat is then transferred to
the container to be used in the experiment. Gentle tapping on the
base of the weighing boat will ensure that the bulk of the sample is
transferred, but it is unimportant if traces of the sample remain.
Finally, the weighing boat and any residual material are accurately
weighed and the mass recorded. The accurate mass of the sample
transferred is the difference between the two recorded masses.
Desiccator
A desiccator is a closed vessel that contains a desiccant (a drying
agent) in its base. The main desiccant is self-indicating silica gel: it is
blue when dry and turns pink when it absorbs moisture. An airtight
seal is maintained in the desiccator by lightly greasing the ground-
glass surfaces on the lid and base.
Vacuum filtration
Vacuum filtration is a technique used for separating a solid product
from a liquid. This type of filtration is carried out under reduced
pressure and is faster than other methods of filtration. The mixture of
solid and liquid is poured through a filter paper into a Büchner funnel,
a Hirsch funnel or a sintered glass funnel. The solid is trapped by the
filter and the liquid is drawn through the funnel into the flask below
by a vacuum.
Figure 33.1 Vacuum filtration apparatus
Refluxing
Refluxing is a technique used to apply heat energy to a chemical
reaction mixture over an extended period of time. The liquid reaction
mixture is placed in a pear-shaped or round-bottomed flask, along
with anti-bumping granules, with a condenser at the top. Anti-
bumping granules are added to minimise the tendency for violent
boiling, which is a safety hazard.
The flask is heated vigorously over the course of the chemical
reaction; any vapours given off are immediately returned to the
reaction vessel as liquids when they reach the condenser.
Figure 33.2 Reflux apparatus
Hints & tips
Refluxing and distillation are often done using ‘Quickfit’ glassware.
To create an airtight seal, and so the joints do not stick, a small
amount of grease should be used around the seals.
Distillation
Distillation can be used to purify a compound from a mixture of
liquids. It does this by a process of heating and cooling the liquids so
they can be separated from each other. The method uses their
different boiling points to separate the more volatile from the less
volatile material. A difference of 15 °C between the boiling points of
two liquids is necessary to make sure that the maximum amount of
desired product is isolated from the mixture. The liquid that distils
over within a certain temperature range should be collected in the
receiving flask. The temperature range will be given in the
procedure, and it will detail the specific temperature at which the
pure product boils. If the liquid product is particularly volatile, it is
good practice to place the receiving flask in an ice/water bath and to
ensure the receiver adapter on the condenser extends well into the
flask. This measure will help to minimise loss of product through
evaporation.
Hints & tips
Anti-bumping granules should be used in both distillation and
refluxing. They should be added at the start before heating is
started.
Figure 33.3 Distillation apparatus
Solvent extraction
Solvent extraction is another process which can be used to purify a
compound by using its different solubility between two solvents. This
technique is used in organic chemistry for purification.
If the desired product is present in an aqueous mixture, it can be
extracted from the mixture by the addition of a second solvent. The
choice of the second solvent, which is usually an organic non-polar
solvent, is important. It must be immiscible with water, so when the
two are mixed they form separate layers. The desired product must
not react with the organic solvent and it must be more soluble in the
organic solvent than in water. When the organic solvent is added to
the aqueous mixture, the product will move out of the aqueous layer
into the organic solvent layer. It can then be separated from the
organic solvent. A separating funnel is used for this technique.
In the first step, the aqueous layer is added to the separating funnel.
The organic solvent is then added to the funnel. With the stopper put
firmly in place, the funnel is inverted. The tap is opened to release
any pressure build-up caused by the solvent vaporising. The tap is
then closed and the mixture is inverted several more times over the
course of a few minutes. This enables more of the product to move
from the aqueous layer into the organic solvent layer due to
increased contact between the two layers. It is important to invert the
funnel and open the tap from time to time to release the pressure.
After shaking, the layers are allowed to separate. The organic
solvent is usually less dense than water so will form the top layer.
The lower aqueous layer can then be removed by turning the tap.
The tap is closed once the organic layer starts to get close to the
opening. The top organic layer is then collected in a separate flask.
Some of the product will still be present in the aqueous layer, as an
equilibrium will have been established between the two layers:
product in aqueous solvent ⇋ product in organic solvent
As not all the product has been extracted from the aqueous layer,
this layer is now mixed with a fresh volume of organic solvent in the
separating funnel. This will extract more of the product remaining in
the aqueous layer. This process is usually repeated two or three
times to maximise the yield obtained.
The organic solvent and product are then treated with an anhydrous
salt to remove any remaining water. The salt is removed by filtration.
Recrystallisation
Recrystallisation is a very important technique used to purify solids
by removing unwanted by-products. The steps involved in
recrystallisation are as follows:
• Dissolve the impure crystals in the minimum volume of hot solvent.
• Filter the hot solution to remove any insoluble impurities.
• Allow the solution to cool and crystallise.
• Filter off the crystals using vacuum filtration and wash with cold
solvent.
• Allow the crystals to dry in the oven and then cool in a desiccator.
The solvent for recrystallisation is chosen so that the compound
being purified is completely soluble at high temperatures and only
sparingly soluble at lower temperatures.
Melting-point determinations
A pure substance has a fixed melting point. An impure substance
melts over a wide range of temperatures and at a lower temperature
than a pure substance. The melting point of a substance is the
temperature range from when the solid first starts to melt, to when all
of the solid has melted. To check the purity of a solid, a melting point
can be determined. Some of the solid is put in a melting-point tube
and this is placed in the melting-point apparatus. It is heated slowly
and the temperatures at which the solid starts to melt and the
temperature at which it finishes melting are recorded. The technique
is repeated, and an average of the temperatures is taken. This
average is then compared with known melting-point values in a data
booklet.
The greater the range of melting points, the more impurities are
present. A range of less than 2 °C indicates a fairly pure substance.
Mixed melting-point determination involves mixing a small
quantity of the product with some of the pure compound (50% of
each) and determining the melting point. If the melting point turns out
to be sharp and close to the expected value, then the two
substances must be identical. If the two substances are not similar,
then the melting point of the mixture will be much lower and the
melting range much broader. This results from the fact that each
compound would act as an impurity of the other.
Determination of percentage yield
During the Higher Chemistry course, the calculation to determine
percentage yield was introduced. It is useful for a chemist to be
able to determine whether a reaction gives a high yield and so is
successful or gives a low yield, in which case some alteration of the
experimental procedure can be used prior to the reaction being
repeated. Percentage yields will always be less than 100%. Losses
can occur during a separation technique or may be due to the
transfer of the product from one flask to another. Purification by
recrystallisation can result in losses if some product remains in the
solution. Small yields may also be due to side reactions occurring,
incomplete reactions, impure reactants and the position of
equilibrium.
Example
An esterification reaction mixture containing 7.5 g of ethanoic acid,
excess methanol and concentrated sulfuric acid produced a sample
of methyl ethanoate weighing 6.5 g. Calculate the percentage yield
of the methyl ethanoate.
CH3OH + HOOCCH3 ⇋ CH3OOCCH3 + H2O
1 mole of ethanoic acid = 60
1 mole of methyl ethanoate = 74
Answer
Number of moles of ethanoic acid
1 mol HOOCCH3 : 1 mol CH3OOCCH3
Theoretical mass of CH3OOCCH3 = n × GFM = 0.125 × 74 = 9.25 g
Key link
Unit 3: Organic chemistry and instrumental analysis
Study questions
1 In theory, 1 mole of aspirin is produced for every 1 mole of
salicylic acid reacted. In an experiment, 8.1 g of aspirin was
obtained from 9.6 g of salicylic acid. Using the data given,
calculate the percentage yield of aspirin.
1 mole of aspirin = 180 g; 1 mole of salicylic acid = 138 g
[3 marks]
2 The ester propyl ethanoate can be prepared in the laboratory.
a) Name the alcohol and carboxylic acid required for this
reaction along with the catalyst used.
[1 mark]
b) Why would an electric heating mantle be used to provide heat
rather than a Bunsen burner?
[1 mark]
c) Why would this reaction be carried out under reflux?
[1 mark]
3 Describe how a mixture of butan-1-ol and butan-2-ol could be
separated. You may wish to use page 9 of the data booklet to
assist you.
[2 marks]
Chapter 34
Standard solutions and dilutions
Key points
• A solution of accurately known concentration is called a standard
solution.
• Standard solutions can be prepared by accurate dilution by
pipetting an appropriate volume of a standard solution into a
volumetric flask, making up to the graduation mark with solvent,
stoppering and inverting.
• A primary standard must be available in a high state of purity, be
stable when solid and in solution, be soluble and have a
reasonably high GFM.
• Examples of primary standards include sodium carbonate,
hydrated oxalic acid, potassium hydrogen phthalate, silver nitrate,
potassium iodate and potassium dichromate.
• Sodium hydroxide is not a primary standard as it has a relatively
low GFM, is unstable as a solid (absorbs moisture) and unstable
as a solution. Sodium hydroxide solution must be standardised
before being used in volumetric analysis.
Standard solutions
A standard solution is one where the concentration is known
accurately. It can be prepared directly from a solute if that solute is a
primary standard. To be suitable as a primary standard, a
substance must have the following properties:
• it must have a high purity
• it must be stable in air and in solution
• it must be readily soluble in a solvent (normally water)
• its solubility should be high enough that solutions of relatively high
concentrations can be prepared
• it should have a reasonably large relative formula mass in order to
minimise the uncertainty in the mass of substance weighed out.
Sodium hydroxide is not suitable as a primary standard as it has a
relatively low GFM, is unstable as a solid (it absorbs moisture) and
unstable as a solution. Sodium hydroxide solution must be
standardised before being used in volumetric analysis. There are a
limited number of primary standards available. Examples are sodium
carbonate, oxalic acid, potassium hydrogen phthalate, silver nitrate,
potassium iodate and potassium dichromate.
To make up a standard solution, the mass of the solute is first
weighed accurately and recorded. The mass of solute is dissolved in
a small volume of solvent in a beaker. The solution should then be
transferred carefully to an appropriate standard flask along with
several washings from the beaker to ensure complete transfer of
solute. More distilled water is added, making the solution up to the
graduation mark; a dropping bottle is used when the volume is close
to the graduation mark. The flask is then stoppered and inverted
several times.
The standard solution can then be diluted to make solutions with a
lower concentration. This involves the use of pipettes and standard
flasks (Figure 34.1).
Figure 34.1 A standard flask
Diluting standard solutions
Standard flasks are used when diluting one of the standard solutions
before titration is carried out. Dilution involves the following stages:
• pipetting a known volume of standard solution into a clean
standard flask
• adding deionised water to the flask until the water is just below the
graduation mark
• using a dropper to add deionised water very slowly until the
bottom of the meniscus is on the graduation mark
• putting a stopper in the standard flask and then inverting it to mix it
thoroughly.
The dilution factor is the amount the original solution is diluted by. It
is calculated by dividing the new total by the volume of the original
solution put into the mixture:
(as long as the units of volume are the same).
If a 25 cm3 sample of the solution is made up to a total volume of
250 cm3 using deionised water, then the dilution factor is 10. If a 10
cm3 sample of the solution is made up to a total volume of 250 cm3
using deionised water, then the dilution factor is 25.
When a concentrated solution is diluted the amount of solute does
not change, only the amount of solvent.
For example, you may be asked what volume of a 0.25 mol l−1
potassium carbonate solution is required, when diluted with water, to
make one litre of a potassium carbonate solution with a
concentration of 0.1 mol l−1.
First determine the number of moles of potassium carbonate you will
need in the diluted solution:
number of moles = concentration of diluted solution × volume of
diluted solution = 0.1 × 1 = 0.1 mol
Then calculate the volume of stock solution which contains this
number of moles:
The relationship C1V1 = C2V2 can also be used to calculate the
volume of stock solution needed, where C1 is the concentration of
the original solution, V1 is the volume of the original solution, C2 is
the concentration of the diluted solution and V2 is the volume of the
diluted solution.
Using the numbers from the example above:
C1V1 = C2V2
0.25V1 = 0.1 × 1
Examples
1 What volume of 0.5 mol l−1 sodium carbonate is required, when
diluted with water, to make 500 cm3 of a sodium carbonate
solution with a concentration of 0.1 mol l−1?
Answer
C1V1 = C2V2 can be used
C1 = 0.5 mol l−1, V1 = ?, C2 = 0.1 mol l−1, V2 = 0.5 l
0.5 × V1 = 0.1 × 0.5
Therefore 100 cm3 of stock solution should be pipetted into a
500 cm3 standard flask and made up to the graduation mark.
The dilution factor in this example is
2 What volume of 0.5 mol l−1 sodium sulfate is required, when
diluted with water, to make 1 litre of a solution with a sodium ion
(Na+) concentration of 0.2 mol l−1?
Answer
C1V1 = C2V2 can be used
C1 = 0.5 mol l−1, V1 = ?, C2 = 0.2 mol l−1, V2 = 1 l
0.5 × V1 = 0.2 × 1
The dilution factor in this example is
But the question asks for sodium ion concentration. The formula
for sodium sulfate is Na2SO4. There are two Na+ ions present so
we need to divide our volume by
Therefore 200 cm3 of stock solution should be pipetted into a 1 l
standard flask and made up to the graduation mark.
Units for concentrations of
solutions
Concentrations of solutions can often be expressed in units other
than mol l−1, particularly by pharmacists or in medications used by
doctors or veterinarians. Table 34.1 shows the common units that
you may encounter.
In order to convert from moles per litre (mol l−1) to grams per litre (g
l−1), multiply the concentration in mol l−1 by the gram formula mass
of the substance:
• mol l−1 → g l−1 multiply by GFM
• g l−1 → mol l−1 divide by GFM.
Example
3 Give the concentration, in g l−1, of a solution of potassium
chloride if 0.25 mol is dissolved in 250 cm3 of water.
Answer
Potassium chloride, KCl, GFM = 74.6
Concentration in mol l−1
Hints & tips
The relationship C1V1 = C2V2 can be used even when the
concentration of a solution is not given in mol l−1.
Study questions
1 Preparation of a 250 cm3 solution of 0.15 mol l−1 hydrated
sodium bicarbonate (NaHCO3.2H2O) was required for analysis.
What mass of solid is required to make this solution?
[1 mark]
2 5 cm3 of a 0.50 mol l−1 solution of hydrochloric acid was added to
a 1000 cm3 standard flask and made up to the mark with
deionised water. What will the final concentration of the
hydrochloric acid solution be?
[1 mark]
3 What volume of 0.2 mol l−1 sodium phosphate is required to
make, by dilution with water, 250 cm3 of a solution with a sodium
ion concentration of 0.06 mol l−1?
[1 mark]
4 Give the concentration of a sulfuric acid solution of 35 g l−1 in mol
l−1.
[1 mark]
Chapter 35
Volumetric analysis
Key points
• There are four main types of titrations:
• acid–base titrations
• redox titrations
• complexometric titrations
• back titrations.
• Back titrations are used to find the number of moles of a
substance by reacting it with an excess volume of a reactant of
known concentration. The resulting mixture is then titrated to work
out the number of moles of the reactant in excess. From the initial
number of moles of that reactant, the number of moles used in
the reaction can be determined. The initial number of moles of the
substance being analysed can then be calculated.
• A back titration is useful when trying to work out the quantity of
substance in a solid with a low solubility.
Titration
Volumetric analysis involves using a standard solution in a
quantitative reaction to determine the concentration of another
reactant. This is achieved by titration, a procedure whereby one of
the solutions is slowly added from a burette to a pipetted volume of
the other solution contained in a conical flask. The point at which
reaction between the two reactants is just complete is usually
detected by adding a suitable indicator which gives a permanent
colour change. A ‘rough’ titration is carried out first to determine an
approximate titre volume and to observe the expected colour
change. Subsequent titre volumes should aim to be concordant,
within 0.1 or 0.2 cm3 of each other. The average of the concordant
results is used in all calculations.
Figure 35.1 Carrying out a titration
Hints & tips
Titration results are concordant when titre volumes are within 0.1 or
0.2 cm3 of each other.
Types of titrations
There are four main types of titrations, which will be discussed in
further detail: acid–base, redox, complexometric and back.
Acid–base titrations
Acid–base titrations are neutralisation reactions. An indicator is
required and the choice of indicator depends upon the pH
equivalence point. The experimenter is looking for a permanent
colour change in the solution in the conical flask. This is called the
end point. The equivalence point is the point at which the reaction is
just complete.
Key link
Unit 2 Chapter 13: Indicators and buffers
Redox titrations
Redox titrations are based on redox reactions. Redox reactions
involve the transfer of electrons from the reducing agent to the
oxidising agent.
• Oxidising agents include potassium permanganate and sodium
dichromate; these are easily reduced.
• Reducing agents include sodium sulfite, potassium iodide and any
alkali metal; these are easily oxidised.
Potassium permanganate and sodium dichromate are widely used
since they are self-indicating. Potassium permanganate is
decolourised in redox reactions and the end point is indicated by a
very pale pink, almost colourless, colour caused by a small excess of
Mn(VII).
Sodium dichromate also acts as its own indicator as its colour
changes from orange to green.
Complexometric titrations
Complexometric titration is a form of volumetric analysis in which
the formation of a coloured complex is used to indicate the end point
of a titration. Complexometric titrations are useful for the
determination of metal ions in solution. EDTA is generally used as it
forms complexes with metal ions in a 1 : 1 ratio. Its use can be
limited, though, as appropriate indicators are not always available.
The indicator has to form a coloured complex with the metal ion.
However, the indicator must bind less well with the transition metal
than the EDTA does. As the EDTA is added to the solution, the
indicator needs to be displaced. When all the indicator attached to
the metal ions has been displaced, a colour change is observed,
indicating the end point of the reaction. The most common indicator
used is murexide, which is an excellent indicator for the titration of
calcium and nickel ions.
Back titrations
Most titrations are direct: one reagent is added directly to the other
until the end point is reached. However, a direct titration may not
always be possible. In such cases a technique known as a back
titration is used. This involves adding a known but excess amount
of one standard reagent to a known mass of the substance being
determined (the analyte). Once the reaction between the two is
complete, the excess amount of the standard reagent is determined
by titration against a second standard reagent. Moles of the excess
reagent which have reacted can then be determined.
Back titrations are used when the analyte is insoluble or has low
solubility, such as aspirin or calcium carbonate. A control experiment
can be used to validate the back titration technique. This involves the
use of a solution of known concentration.
Hints & tips
Several titration values are usually given and you may be asked to
choose which ones are concordant. Choose the ones which are
within 0.2 cm3 of each other.
Example
A 0.120 g sample of egg shell, calcium carbonate (CaCO3), was
placed in a 250 cm3 conical flask. 50 cm3 of 0.200 mol l–1
hydrochloric acid (HCl) was added to dissolve the egg shell. The
excess hydrochloric acid was then titrated with 0.250 mol l–1 NaOH.
The average NaOH titre was found to be 32.12 cm3. Calculate the
mass of calcium carbonate, in grams, present in the sample of egg
shell.
Answer
Total moles of HCl
1 mole NaOH reacts with 1 mole HCl
Moles of excess HCl left = 0.00803 mol
2HCl + CaCO3 → CaCl2 + CO2 + H2O
2 moles HCl react with 1 mole CaCO3
Moles of CaCO3
GFM of CaCO3 = 100
Mass of CaCO3 present in egg shell = n × GFM = 0.000985 × 100 =
0.0985 g
Study questions
1 Calculate the concentration of nitric acid if 11.5 cm3 of nitric acid
reacts with 25 cm3 of 0.200 mol l–1 potassium hydroxide solution.
[2 marks]
2 A tablet of aspirin weighing 1.05 g was hydrolysed completely by
25 cm3 of 0.110 mol l–1 sodium hydroxide solution.
The solution was then transferred to a 250 cm3 standard flask
and made up to the mark with water. 25 cm3 samples of this
solution were titrated with 0.010 mol l–1 hydrochloric acid. The
average titre was found to be 15 cm3.
CH3COOC6H4COOH + 2NaOH → CH3COONa +
HOC6H4COONa + H2O
NaOH + HCl → NaCl + H2O
Calculate the mass of aspirin present in the tablet.
[3 marks]
Chapter 36
Gravimetric analysis
Key points
• Gravimetric analysis can be used to determine the mass of an
element or compound in a substance.
• The substance is converted into another substance of known
chemical composition, which can be readily isolated and purified.
• The conversion can occur either through precipitation or
volatilisation.
• In precipitation conversion, the substance undergoes a
precipitation reaction. The precipitate is separated from the filtrate
and the filtrate tested to ensure the reaction has gone to
completion. The precipitate is washed, dried to constant mass
and then weighed.
• In volatilisation conversion, the substance is heated and any
volatile products (often water) are evaporated. The substance is
heated to constant mass and the final mass recorded.
Gravimetric analysis
Gravimetric analysis is a method used to analyse the mass of an
element or compound present in a substance. Gravimetric analysis
can be generalised into two types: precipitation and volatilisation.
Precipitation
Gravimetric analysis often involves the precipitation of a new
substance followed by filtration. For example, if the mass of lead
nitrate in a sample was to be analysed, it could be reacted with
potassium iodide to form a precipitate of lead iodide. The lead iodide
could then be filtered, washed, dried and finally weighed. Through
the stoichiometric equation, the concentration or mass of lead nitrate
could be determined.
Pb(NO3)2(aq) + KI(aq) → PbI2(s) + 2KNO3(aq)
When using this method, the precipitate should have a low solubility.
This is so all the product is precipitated. It should also be stable at
temperatures of 100–105 °C, so the solid can be dried in an oven.
The precipitate should have a particle size that is not too small, to
allow easy filtration.
The success of this method relies on the procedure being carried out
carefully. All material needs to be transferred from reaction vessel to
filtration. A Büchner funnel and water pump can be used to carry out
filtration as this is much faster than gravity filtration. The precipitate
should be washed with a little solvent to remove any traces of the
filtrate and then transferred to an oven to dry. A small amount of
precipitate is lost during transfer, but not during drying or weighing.
After drying the sample in an oven, the sample is transferred to a
desiccator to cool. After cooling in a desiccator, the precipitate is
weighed and the heating, cooling and weighing are repeated until the
mass is constant. This proves that all the water has been driven off.
The accuracy of this method relies on the use of reactants at high
purity, reactions going to completion (not being equilibria) and having
no side reactions. The accuracy is also dependent upon the balance
used. To obtain the final mass of product, the initial and final masses
are recorded and the difference between them taken.
Example
1 A student decided to analyse a biscuit for the presence of sodium
chloride. A sample of the biscuit was weighed at 110 g and then
liquidised with water to dissolve all the soluble substances
present. The food/water mixture was filtered and the filtrate was
reacted with excess silver nitrate (AgNO3). This produced a silver
chloride precipitate which, after drying, weighed 2.15 g.
Assuming no other products were present in the precipitate and
all the chloride ions came from sodium chloride, determine the
mass of sodium chloride and the percentage by mass present in
the biscuits.
Answer
Work out the balanced equation for the reaction between sodium
chloride and silver nitrate:
NaCl(aq) + AgNO3(aq) → AgCl(s) + NaNO3(aq)
GFM of AgCl = 143.4; GFM of NaCl = 58.5
Number of moles of AgCl
Therefore, NaCl present in sample = 0.015 mol
Mass of NaCl = n × GFM = 0.015 × 58.5 = 0.8775 g
Percentage by mass of NaCl in biscuit
Volatilisation
Another common form of gravimetric analysis involves heating to
change one substance into another. This is often by a decomposition
reaction, such as barium carbonate into barium oxide, or due to loss
of water of crystallisation from a salt (dehydration).
Water of crystallisation is the molecules of water that are found in
some salt crystals when they are crystallised out of water. Copper
sulfate has blue crystals when hydrated; the actual formula of the
crystals is CuSO4.5H2O. This formula shows that five water
molecules are associated with each ‘CuSO4’ in its ‘hydrated’
crystalline form. Anhydrous copper sulfate is a white powder and has
the formula CuSO4.
Water of crystallisation can be driven off hydrated crystals by strong
heating. This results in a weight loss, from which the quantity of
water per mole of compound can be found.
Figure 36.1 Volatilisation gravimetric apparatus
Hints & tips
The crucible and samples must not be hot when they are weighed,
otherwise their measured mass will be less than their true mass.
Cooling should take place in a desiccator so no moisture from the
air is taken up by the sample.
This technique is carried out by weighing an empty crucible, then re-
weighing it with the sample present. The masses are recorded. The
sample is heated gently for 2 minutes and then strongly for 10–15
minutes. It is allowed to cool in a desiccator and is then re-weighed.
A desiccator allows the crucible to cool down without absorbing any
more moisture.
The completion of this reaction is checked by repeated heating and
cooling followed by weighing until a constant mass is obtained.
Successive weighings, at room temperature, should be within +0.01
g of each other.
Example
2 Hydrated sodium carbonate, Na2CO3.xH2O, is also called
washing soda, It is a natural cleaner and water softener. A
sample of washing soda was heated in a crucible until no further
mass loss took place. Using the data given below, calculate the
number of water molecules in the water of crystallisation of
washing soda and therefore the formula of hydrated sodium
carbonate.
Mass of crucible = 33.65 g
Mass of crucible + hydrated sodium carbonate = 45.79 g
Mass of crucible + anhydrous sodium carbonate = 38.15 g
Answer
Mass of hydrated sodium carbonate = 45.79 – 33.65 = 12.14 g
Mass of anhydrous sodium carbonate = 38.15 – 33.65 = 4.5 g
Mass of water = mass of hydrated sodium carbonate – mass of
anhydrous sodium carbonate = 12.14 – 4.5 = 7.64 g
GFM of sodium carbonate = 106 g mol–1; GFM of water = 18 g
mol–1
Moles of sodium carbonate
Moles of water
The ratio of number of moles of sodium carbonate to the number
of moles of water = 0.0425 : 0.424 = 1 : 9.98, so 1 : 10. There
are, therefore, 10 water molecules in the water of crystallisation
of washing soda, giving the formula of hydrated sodium
carbonate as Na2CO3.10H2O.
Study questions
1 A 5.05 g sample of silver jewellery was dissolved using excess
dilute nitric acid. The sample was then treated with an excess of
sodium chloride solution, which yielded a precipitate of silver
chloride (AgCl). No other substances were precipitated. The
precipitate was filtered, washed and dried and was weighed at
2.5 g. Calculate the mass of silver present in the jewellery and
express this as a percentage of silver found in the jewellery
sample.
[3 marks]
2 Nails are made up of an iron alloy. A nail of 7.45 g was dissolved
in excess dilute hydrochloric acid. This caused all the iron to be
converted into iron(II) chloride solution. After the removal of all
other substances, the sample was evaporated to dryness,
leaving 12.8 g of solid iron(II) chloride. Calculate the mass of iron
in the nail and express the percentage of iron present in the alloy.
[3 marks]
3 A 0.966 g sample of hydrated copper(II) nitrate was heated to a
constant mass of 0.750 g of the anhydrous salt, Cu(NO3)2.
Determine the water of crystallisation present and write the
formula for hydrated copper(II) nitrate.
[3 marks]
Chapter 37
Chromatography
Key points
• Chromatography is a technique used to separate the components
present within a mixture.
• Chromatography separates substances by making use of
differences in their polarity or molecular size.
• Thin-layer chromatography (TLC) uses a fine film of silica or
aluminium oxide spread over glass, aluminium foil or plastic. A
small sample of the mixture being tested is spotted onto the base
(pencil) line of the chromatogram. A solvent dissolves the
compounds in the spot and carries the compounds up the
chromatogram. How far the compounds are carried depends on
how soluble the compounds are in the chosen solvent and how
well they adhere to the plate. A developing agent or ultraviolet
light is normally required to visualise the spots on the
chromatogram.
• Rf values can be calculated using the formula:
• Under the same conditions (temperature, solvent and saturation
levels) a compound always has the same Rf value (within
experimental error).
• The identity of a compound can be confirmed by comparing the
experimentally determined Rf value with a literature or known
value determined under the same conditions. A compound can
also be determined by making a direct comparison on a TLC
plate between the compound being tested and the pure
substance.
• TLC is used to assess the purity of substances. A pure
substance, when spotted and developed on a TLC plate, should
appear as a single spot (some impurities may not be visible by
TLC analysis). The presence of more than one spot shows that
impurities are present.
Thin-layer chromatography
Chromatography is a method of separating soluble substances by
their partition between two different phases. All types of
chromatography have a mobile phase and a stationary phase. Thin-
layer chromatography is carried out in an almost identical fashion
to paper chromatography. The thin-layer plates are usually coated
with silica gel. Silica gel is a polymer formed from silicic acid that has
many hydroxyl groups. This makes the surface of the gel very polar.
As a result, polar substances tend to be held by the gel, making their
movement much slower than that of non-polar substances.
The silica gel makes up the stationary phase while the mobile phase
is a suitable solvent or mixture of solvents. The solvent moves
through the stationary phase and carries the dissolved components
of the substance to be analysed with it. Different substances travel
along the stationary phase at different rates. This allows them to
separate as the mobile phase moves along the stationary phase.
To prepare a TLC plate, a fine pencil line called the base line is
drawn about 1.5 cm from one end. A pencil cross is put where a
sample is to be added. Using a capillary tube, some of the sample
dissolved in an appropriate solvent is spotted onto the cross. The
TLC plate is then placed in the development tank with the pencil line
just above the level of the solvent. The plate is allowed to run for 20
to 30 minutes as the solvent moves up. The solvent should not run
off the end of the chromatogram, so checking is necessary. As the
solvent moves, the substances in the spot also move up the plate
and separate. Once finished, a pencil line is drawn to show where
the solvent has reached on the chromatogram. This is called the
solvent front. The distance the solvent has moved is measured in
centimetres.
Figure 37.1 Thin-layer chromatography of black ink
A substance should always have the same Rf value with a given
solvent (as long as conditions of temperature and saturation levels
are the same), and that allows comparisons to be made.
The Rf value is calculated using the following formula:
Most chromatograms are not visible unless developed. Some spots
on chromatograms are only visible under UV light and can be
marked with a pencil so the Rf value can be determined in normal
light. A solution of iodine can be used to mark some aromatic
compounds. Amino acid spots are usually developed by spraying the
plate with ninhydrin.
Example
The TLC plate in Figure 37.2 shows a developed chromatogram
with three spots. Use the diagram and the information from the
table below to identify the amino acid labelled as ‘X’ on the
chromatogram.
Amino acid Rf value
Alanine 0.38
Arginine 0.20
Asparagine 0.50
Lysine 0.14
Answer
Figure 37.2 Thin-layer chromatography
The process of TLC works by making use of differences in the
polarity or molecular mass of molecules:
• Molecules that have more polarity will move slower through the
plate than molecules with less polarity.
• Molecules with a large molecular mass will move more slowly than
molecules with a low molecular mass.
So larger molecules with more polarity will have lower Rf values
under the same conditions (temperature, solvent and saturation
levels).
Uses of thin-layer chromatography
Thin-layer chromatography can be used to determine the purity of a
substance prepared in the lab. If the sample is a pure product this
will only show as one spot; there would be more than one spot if
impurities were present.
The identity of a compound can be confirmed by comparing Rf
values with a literature or known value determined under the same
conditions. Pure samples can be run alongside the sample being
analysed to see whether the same Rf values are obtained. Another
way to determine whether a sample is pure is to use a co-spot. This
is a 50 : 50 mixture of the sample and a known pure sample and it
can be spotted on the same cross. If the sample is pure, then only
one spot will be present on the chromatogram. If the sample is
impure, then more than one spot will be observed.
Key link
Unit 3: Organic chemistry and instrumental analysis
Study questions
1 A sample of protein was hydrolysed and the mixture of amino
acids formed was separated using thin-layer chromatography.
Which of the following is not required for TLC?
[1 mark]
A A ninhydrin developer
B A solvent
C A gaseous mobile phase
D A stationary phase
2 Melting-point determination and thin-layer chromatography are
useful techniques for assessing the purity of a compound. Which
line in the table below shows the results expected from a pure
compound which is known to have a melting point of 130 °C?
[1 mark]
Chapter 38
Colorimetry
Key points
• Colorimetry uses the relationship between colour intensity of a
solution and the concentration of the coloured species present.
• A colorimeter or a spectrophotometer can be used to measure
the absorbance of light of a series of standard solutions. This
data is then used to plot a calibration graph.
• The concentration of the solution being tested can be determined
from its absorbance value and by comparing it to the calibration
curve.
• The concentration of the coloured species in the solution being
tested must lie in the straight-line section of the calibration graph
to give an accurate value.
Colorimetry is the measuring of colour intensity. The piece of
apparatus used to measure the colour intensity of a solution is called
a colorimeter. The intensity of the colour in the solution is directly
proportional to the concentration of the coloured species. A
colorimeter works by passing a beam of light through the coloured
sample. The light that passes into the sample is called the incident
light and the light that passes through the sample is called the
transmitted light. The ratio of the incident light to the transmitted
light gives an absorbance value.
The light from the source passes through a coloured filter which
changes the light to a specific wavelength. The incident light then
passes into the sample. A solution is coloured if it absorbs some but
not all parts of the white light passing through it.
Figure 38.1 The colorimeter process
The choice of the incident light wavelength (i.e. colour) is important.
It should be the complementary colour to the solution to allow the
maximum absorbance. The complementary colours and wavelengths
can be determined from the colour wheel on page 20 of the data
booklet. Colours directly opposite each other on the colour wheel are
complementary colours. For example, if you are measuring the
intensity of a blue solution, you should use a yellow filter as the
solution is absorbing red light. The amount of red light the sample
absorbs relates directly to the concentration. If you are measuring
the intensity of a green solution, a purple filter should be used as the
solution is absorbing purple light.
Figure 38.2 A colorimeter
A calibration curve should be set up at first with known
concentrations of the reactant or product. This will enable the
colorimeter reading to be related directly to the concentration of the
sample. The standard samples are placed in a cuvette, which is a 1
cm × 1 cm vial made from glass or plastic. The transmitted light is
detected and the absorbance of the solution can be calculated using
where I0 is the initial intensity of light, before it passes through the
solution, and I is the intensity of light after it passes through the
solution.
The path length is defined as the distance that light travels through
a sample in a cuvette. A simple equation called the Beer–Lambert
law relates the absorbance to concentration and path length:
where c is the concentration in mol l–1, l is the path length in cm and
ε is the molar absorption coefficient (this is a constant dependent on
the solution at a particular wavelength).
When the absorbance is plotted against concentration of the
coloured standard solution, the resulting calibration graph should be
a straight line (Figure 38.3), indicating that the two are directly
proportional. Calibration graphs must be drawn as best-fit straight
lines.
The unknown coloured sample is then run and an absorbance value
obtained. From this value, the concentration of the unknown sample
can be obtained from the calibration graph.
Figure 38.3 A calibration graph
Hints & tips
Calibration graphs must be drawn as best-fit straight lines. These
need to show both maximum and minimum values if plotted using
computer software.
The concentrations of the standard solutions used to create the
calibration graph should cover the range expected for the
concentration of the unknown solution. If the unknown solution has
too high an absorbance value for the calibration graph drawn, a new
calibration graph using suitable concentrations of standard solutions
should be run. Alternatively, the unknown solution can be diluted
accurately and its new absorbance measured. The calibration graph
can be used to determine the concentration and this is then
multiplied by the dilution factor.
If the absorbance reading is too high for the concentrations of
standard solutions, then a curve is seen rather than a straight line
when plotted. This requires the standard solutions to undergo
accurate dilution. Absorbance readings taken from a curve are not
directly proportional to the concentration. The concentration of
coloured species in the solution being tested must always lie in the
straight-line section of the calibration graph.
Absorbance values should always be measured at least twice and
an average taken at each concentration. Values which show an
anomaly (vary greatly from the line of best fit) can be removed.
If the data points on your calibration graph are close to the line of
best fit, it can be assumed that the value obtained will have a high
degree of accuracy. If the data points are not all lying close to the
line of best fit, then the value obtained is more likely to be
inaccurate.
Hints & tips
A blank cuvette should be run on the specified wavelength using
the solvent to determine a blank or reference sample. This
establishes a base line.
Key links
Unit 1 Chapter 7: Ligands and transition metal complexes
Unit 4 Chapter 34: Standard solutions and dilutions
Study questions
1 Iron can be reacted with thiocyanate to give a blood-red coloured
complex. Describe all the steps a student would be required to
take in order to determine the concentration of iron in an iron
tablet using the technique of colorimetry.
[3 marks]
2 An unknown solution is blue in colour. What colour of filter should
be used in the colorimeter?
[1 mark]
Unit 4 Glossary
Absorption spectrum: A spectrum consisting of dark absorption
lines superimposed on a bright continuous spectrum; it shows the
absorption of radiation by a material over a range of wavelengths.
Accuracy: A measure of the agreement between a measurement
and the true or correct value.
Actual yield: The quantity of a product that is obtained from a
chemical reaction.
Anomalous data: Data with unexpected values that do not match
the relationship observed in previous measurements.
Atom economy: A ratio of the total mass of atoms in the desired
product to the total mass of atoms in the reactants.
Back titration: A titration performed when an insoluble and slowly
reacting reagent is treated with an excess of an acid or base. The
excess acid or base is then titrated with base or acid solution of a
primary standard.
Chromatography: A technique for analysing or separating mixtures
of gases, liquids or dissolved substances based upon differential
solubility in two phases.
Colorimeter: An instrument used to measure the intensity of colour
in a solution.
Complementary colours: Colours directly opposite each other on
the colour wheel. One of two coloured lights related to each other in
such a way that when blended together, they produce white light; so-
called because each colour makes up in the other what it lacks to
make it white.
Concentration: The ratio of the amount (in moles) of a substance
dissolved in a given volume of solution; expressed in mol l–1.
Dependent variable: The variable that is measured during an
investigation.
Emission spectroscopy: The study of emission spectra produced
by excited substances (often gaseous atoms or molecules).
Empirical formula: A formula for a compound which shows the
simplest whole number ratio of atoms present.
End point: The point at which the indicator changes colour suddenly
during a titration.
Equivalence point: The point in an acid–base titration where the
acid and base have been added in stoichiometric amounts; if a
suitable indicator is chosen it corresponds to the end point.
Gravimetric analysis: A method of quantitative analysis for finding
the composition and formulae of compounds based on accurate
weighing of reactants and products.
Independent variable: The variable that is manipulated or changed
during an investigation.
Line of best fit: The straight line or curve which gives the best
approximation to a given set of data.
Literature value: A value from the chemical literature of a physical
constant or experimental measurement.
Mobile phase: Liquid or gas which percolates through or along the
stationary phase during chromatography.
Parts per million (ppm): Concentration expressed as parts of solute
per million parts of solution.
Percentage error: An error expressed as a percentage of the value
measured or the true value.
Percentage uncertainty: Random uncertainty/actual measurement
× 100
Percentage yield: The actual yield as a percentage of the
theoretical yield; actual yield/theoretical yield × 100
Precision: The total amount of random error present in a
measurement; indicates how close together a series of
measurements are.
Qualitative analysis: Analysis used to identify the substance but not
the quantity of the substance.
Quantitative analysis: Analysis used to determine the quantity of
each component in a material or substance.
Redox reaction: A reaction involving reduction and oxidation which
results in one or more electrons being transferred and is
accompanied by changes in oxidation numbers.
Redox titration: A titration used to determine the concentration of a
solution of an oxidising agent or a reducing agent.
Reflux: The process of boiling a liquid in a flask connected to a
condenser so that the condensed liquid runs back into the flask.
Repeatability: The closeness of agreement between independent
results obtained with the same method, under the same conditions
(same operator, same apparatus, same laboratory and after short
intervals of time).
Retention time: The time between injection of a substance onto a
chromatography column and the appearance of the peak.
Separating funnel: A funnel with a tap at the lower end and a
stopper in the top used to separate immiscible liquids.
Solute: The solid, liquid or gas that has been dissolved to form a
solution.
Solution: A solution is formed when a solid, liquid or gas is
dissolved into a solvent.
Solvent: A liquid that dissolves solids, liquids or gases to form a
solution.
Solvent extraction: A method for separating a compound from a
mixture by causing it to move to another solvent, for example the
extraction of solute from an aqueous solution using an organic
solvent in which the solute of interest is more soluble.
Standard solution: A solution whose concentration is accurately
known and which does not change with time.
Stationary phase: One of the two phases forming a
chromatographic system, which may be a solid, a gel or a liquid.
Steam distillation: A technique used to purify organic substances
that are immiscible with water and which would decompose if
distilled. The impure substance is added to water and steam is blown
through the mixture. The distillate is a two-phase mixture of water
and the purified organic material.
Stoichiometric quantities: Refers to a reaction where amounts of
reactants are reacted together so that all are consumed at exactly
the same time.
Systematic error: An error which biases your measurements in
some predictable way; the effects cannot be reduced by repeating
the measurement and averaging.
Theoretical yield: The maximum amount or mass of a particular
product that can be formed when the limiting reactant is completely
consumed and there are no losses or side reactions.
Titration: A chemical technique in which one solution is used to
analyse another solution and find its concentration or amount.
Vacuum filtration: A mixture of solid and liquid is poured through a
filter paper in a Büchner funnel. The solid is trapped by the filter and
the liquid is drawn through the funnel into the flask below by a
vacuum.
Water of crystallisation: Water molecules contained in the crystal
lattice of inorganic salts when they crystallise from aqueous solution.
Unit 4 Exam-style questions
1 Which one of the following is not suitable for the preparation of a
primary standard in volumetric analysis?
A Anhydrous sodium carbonate
B Sodium hydroxide
C Oxalic acid
D Potassium iodate
[1 mark]
2 Sulfa drugs are compounds with antibiotic properties. Sulfa drugs
can be prepared from a solid compound called sulfanilamide.
Sulfanilamide is prepared in a six-stage synthesis. The equation
for the final step in the synthesis is shown.
a) Converting the 4-acetamidobenzenesulfonamide into
sulfanilamide involves gentle heating of the 4-
acetamidobenzenesulfonamide in a solution of water and
hydrochloric acid in a round bottom flask for at least 30
minutes. Name two pieces of equipment necessary to carry
out this reaction successfully.
[2 marks]
b) The sulfanilamide is separated from the reaction mixture and
recrystallised from boiling water.
(i) Why is recrystallisation necessary?
[1 mark]
(ii) Describe the solubility of pure sulfanilamide in cold and
hot water.
[1 mark]
(iii) Simple filtration to isolate the pure sulfanilamide is very
slow. How could the filtration be speeded up?
[1 mark]
c) Describe how a mixed melting point experiment would be
carried out and the result used to confirm that the product was
pure.
[2 marks]
d) Suggest another analytical technique which could be used to
indicate whether the final sample is pure.
[1 mark]
Appendix
Additional features of the
Advanced Higher Chemistry exam
Numeracy
The Advanced Higher Chemistry exam will contain several questions
that test your numeracy skills. Some of these questions will be on
topics you have covered in previous chapters such as calculations
involving Gibbs free energy, entropy and kinetics. Some questions
may ask you to scale up or scale down reaction quantities,
something encountered by practising scientists on a daily basis.
Other types of questions may ask you to calculate empirical
formulae.
Being able to deal with proportion is key to answering numeracy
questions. Some examples are shown below.
Example
Titanium is present in ancient pottery and can be identified using
atomic emission spectroscopy. One piece of ancient pottery gave
the results below.
The ancient pot sample was found to have a titanium content of 460
mg kg–1.
A different pot sample produced a spectral line at 498 nm with a
relative intensity of 60.
Calculate the titanium content, in mg kg–1, for this second sample.
Answer
This is a proportion question.
Step 1: State the relationship.
The ancient sample gave an intensity of 400 nm with 460 mg kg–
1.
Step 2: Determine the ratio of mg kg–1 to intensity.
1 unit of intensity relates to 460/400 = 1.15 mg kg–1
Step 3: Scale up for 60 units of intensity.
60 × 1.15 = 69 mg kg–1
The different pot sample contains 69 mg kg–1.
Example
During colorimetric analysis, 0.45 g of steel paperclips were
dissolved in nitric acid. The manganese in the resulting solution
was oxidised. The solution was made up to 250 cm3 in a standard
flask. The concentration of this solution was found to be 3.28 × 10–4
mol l–1. Calculate the percentage, by mass, of manganese in the
steel sample.
Answer
This is a simple moles-to-mass calculation.
Step 1: Determine the moles in 250 cm3.
Step 2: Determine the mass of manganese.
Mass = moles × GFM = 8.2 × 10–5 × 54.9 = 4.5018 × 10–3 g
Step 3: Calculate the percentage, by mass, of manganese in the
steel sample.
% manganese in steel paperclips = 4.5018 × 10–3/0.45 × 100 =
1%
Example
Oxalic acid is present in rhubarb leaves. The lowest lethal dose of
oxalic acid for humans is estimated to be 400 mg per kg of body
mass. The concentration of oxalic acid in rhubarb leaves is 0.75 g
per 100 g of leaves.
Calculate the mass of rhubarb leaves that a person with a mass of
55 kg would need to consume to reach the lethal dose.
Answer
This is another proportion calculation.
Step 1: State the relationship.
100 g of leaves contains 0.75 g of oxalic acid.
400 mg per kg is a lethal dose.
Step 2: Determine the maximum mass of oxalic acid for a 55 kg
person.
Lethal dosage = 55 × 400 × 10–3 = 22 g
Step 3: Determine the mass of leaves.
100 g → 0.75 g of oxalic acid
of leaves would need to be consumed to
obtain a lethal dose.
Open-ended questions
Chemistry problems encountered during research and development
rarely have one fixed answer. In your Advanced Higher Chemistry
exam you will encounter open-ended questions, where there is
usually more than one ‘correct’ answer. Such questions can be
identified by the use of the phrase ‘using your knowledge of
chemistry’. Open-ended questions will be marked according to the
following mark scheme:
3 marks: The candidate has demonstrated a good conceptual
understanding of the chemistry involved, providing a logically correct
response to the problem/situation presented.
This type of response might include a statement of principle(s)
involved, a relationship or equation, and the application of these to
respond to the problem/situation. This does not mean the answer
has to be what might be termed an ‘excellent’ answer or a ‘complete’
one.
2 marks: The candidate has demonstrated a reasonable
understanding of the chemistry involved, showing that the
problem/situation is understood.
This type of response might make some statement(s) that is/are
relevant to the problem/situation, for example, a statement of
relevant principle(s) or identification of a relevant relationship or
equation.
1 mark: The candidate has demonstrated a limited understanding of
the chemistry involved, showing that a little of the chemistry that is
relevant to the problem/situation is understood.
The candidate has made some statement(s) that is/are relevant to
the problem/situation.
0 marks: The candidate has demonstrated no understanding of the
chemistry that is relevant to the problem/situation.
The candidate has made no statement(s) that is/are relevant to the
problem/situation.
Example
A student stated that ‘Ionisation energy always increases across a
period in the Periodic Table due to the increase in nuclear charge.’
Using your knowledge of chemistry, discuss the accuracy of this
statement.
Suggested approach
You need to demonstrate your knowledge of chemistry at Advanced
Higher level. To start with you should state what is meant by
ionisation energy and illustrate this with equations. Then explain
how increasing the nuclear charge pulls electrons closer to the
nucleus and use examples of elements in a period which follow the
pattern suggested by the student. Then focus on the Advanced
Higher knowledge where you discuss the exceptions to this rule
and explain them in terms of their electronic configurations. For
example:
Boron has a lower 1st ionisation energy than beryllium.
Be: 1s2 2s2
B: 1s2 2s2 2p1
Boron’s outer electron is in a 2p orbital rather than a 2s orbital. The
2p is further from the nucleus than the 2s. The increased distance
results in a reduced attraction and so a reduced ionisation energy.
In addition, the 2p orbital is screened by the 1s2 electrons and by
the 2s2 electrons. This reduces the pull from the nucleus and so
lowers the ionisation energy.
Example
A student suggested that the conversion of methanol and propanoic
acid into methyl propanoate could be followed using analytical
techniques such as chromatography and infrared spectroscopy
allowing a chemist to know when the product ester had formed.
Using your knowledge of chemistry, explain how these
techniques could be used in this way.
Suggested approach
First start off by defining the two techniques suggested.
For infrared spectroscopy, discuss how IR can identify functional
groups. Use the data booklet to quote some of the wavenumbers
for the compounds of interest, i.e. for conversion from an alcohol
and carboxylic acid into an ester you would expect to see a
decrease/disappearance of –OH from the alcohol and acid and the
appearance of the ester functional group. This could be done by
sampling at fixed periods of time and observing the change in IR
spectrum as the reaction proceeds.
Remember this will be an equilibrium reaction so there will always
be some reactants and products. You could say how you could
improve the product yield by removing some product. The ester is
insoluble in water and will float on top of the water-soluble
reactants, so a separating funnel would allow separation of the
ester.
The product IR can be compared with reference samples/library
spectra of the product ester.
Thin-layer chromatography (TLC) of three standards, i.e. methanol,
propanoic acid and methyl propanoate, to calculate/observe their Rf
values. Describe how you determine the Rf values. Also define a
mobile phase that could be used. A different solvent may be
needed for the acid and alcohol to that of the ester due to solubility
differences.
Run plates with these three spots alongside samples from the
reaction to see the appearance of a spot corresponding to the ester
(compare with the reference). This can be done at various time
periods during the reaction to see the formation of ester product.
State how the spots are developed using spotting agents such as
iodine crystals.
Discuss the pros and cons of each technique for identification
methods.
Answers
Unit 1: Answers to Study
questions
Chapter 1 Electromagnetic spectrum
1 f = c/λ = (3.00 × 108)/(900 × 10–9) = 3.33 × 1014 Hz
2 λ = c/f = (3.00 × 108)/(5 × 109) = 0.06 m
3 E = 264 kJ mol–1
E = Lhc/1000λ
λ = Lhc/1000E = (6.02 × 1023) × (6.63 × 10–34) × (3.00 ×
108)/(1000 × 264)
= 4.54 × 10–7 m.
Question asks for answer in nm so 4.54 × 10–7 × 109 = 454 nm
4 a) Wavelength = 650 nm
b) E = Lhc/1000λ
= (6.02 × 1023) × (6.63 × 10–34) × (3.00 × 108)/(1000 × 650 ×
10–9)
= 184 kJ mol–1
Chapter 2 Shapes of atomic orbitals and quantum numbers
1 5
2 a) C
n = 4 so in the 4 shell
l = 1 means in p orbital
m = –2; this is not possible as
m = +1, 0, –1 for the p orbitals.
b) D
c) D
Chapter 3 Orbital notations
1 Potassium ion: 1s2 2s2 2p6 3s2 3p6
2 Krypton atom: 1s2 2s2 2p6 3s2 3p6 3d10 4s2 4p6
3 Gallium atom: 1s2 2s2 2p6 3s2 3p6 3d10 4s2 4p1
Oxygen atom: 1s2 2s2 2p4
4 Bromide ion: 1s2 2s2 2p6 3s2 3p6 3d10 4s2 4p6
5 C
Chapter 4 Ionisation energy
1 Oxygen: 1s2 2s2 2p4
Sodium: 1s2 2s2 2p6 3s1
Manganese: 1s2 2s2 2p6 3s2 3p6 3d5 4s2
2 a) Tungsten: d block
b) Curium: f block
c) Strontium: s block
3 a) p block
b) Atomic number 53; element is iodine
4 The first ionisation energy of boron is lower than expected as,
even with the increased nuclear charge, the added electron is in
the new p subshell which has a slightly higher energy level and is
slightly further from the nucleus. The s2 electrons also provide
some shielding for the group 3 elements. So the ionisation energy
decreases at this point.
Chapter 5 Shapes of molecules and polyatomic ions
1 a) PCl3: trigonal pyramidal shape
b) BrF4–: square planar shape
c) NH4+: tetrahedral
2 a) H2O: angular
b) SiH4: tetrahedral
c) BCl3: trigonal planar
Chapter 6 Transition metals, electronic configuration and oxidation
numbers
1 a) For the Fe2+ ion we remove two electrons from 4s2 leaving us
with: 1s2 2s2 2p6 3s2 3p6 3d6
b) 1s2 2s2 2p6 3s2 3p6 3d5
c) 1s2 2s2 2p6 3s2 3p6 3d7
2 Cr2O72– is a ion, so the overall charge is –2. There are seven
oxygen ions at –2, so that is –14. So:
–2 = (7 × –2) + 2Cr
(–14) + 2Cr = –2 so 2Cr = +12; Cr = +6 in the Cr2O72– ion
The overall charge on the CrO42– ion is –2. There are four oxygen
ions at –2 so that is –8. So:
–2 = (4 × –2) + Cr
(–8) + Cr = –2 so chromium must be Cr = +6 in the CrO42– ion.
The oxidation state for chromium is the same in both ions (+6); it
is neither a reduction nor an oxidation as the number has not
changed.
3 a) Co2O3 has an overall charge of 0. There are three oxygen ions
at –2, so this is –6. So:
0 = (3 × –2) + 2Co; 2Co =+6 so Co =+3
b) [CoCl4]2– has an overall charge of –2. There are four chloride
ions at –1, so this is –4. So:
–2 = (4 × –1) + Co; Co = +2
c) [Ag[NH3]2]+ has an overall charge of +1. There are two
ammonium ions at 0 charge, so Ag = +1
Chapter 7 Ligands and transition metal complexes
1 The ligand is a cyanide ion which is changed to cyanido in the
complex name. The complex has six of them so that is
hexacyanido. The Latin name for iron is ferrate. The oxidation
state for iron in this complex is determined to be +2 so this gives
us Roman numeral (II). So the complex is a hexacyanidoferrate(II)
ion.
2 The ligand is water so it is called aqua in the complex name. The
complex has six of them so that is hexa-aqua. Add the suffix ‘-ate’
in the name of the metal to make it nickelate. The oxidation state
for nickel in this complex is determined to be +2 so this gives us
Roman numeral (II). So the complex is a hexaaquanickelate(II)
ion.
3 K3[Fe(CN)6]
4 [Cu(Cl)4]2–
Chapter 8 Colours of transition metals and catalysts
1 D
2 D
3 Water is higher up the spectrochemical series than chloride so
splits the d orbitals more than chloride. So peak absorption shifts
to a lower wavelength when water replaces chloride.
Unit 1: Answers to Exam-style
questions
1 a) (i) +4, 4, 4+, IV, four [1] (–4, 4– would not be accepted)
(ii)
[1] (Box labels not required; noble gas start accepted in this
case; full or half-headed arrows accepted, but not vertical
lines; follow through.)
(iii) The three d orbitals are all filled singly (with parallel spins).
[1] (follow through)
b) (i) Degenerate means that the orbitals are of equal energy.
[1]
(ii) p orbital [1] (Axes not necessary; any orientation
accepted)
2 D [1]
3 D [1]
4 a) (i) An electron is excited/promoted to a higher energy level. [1]
When the electron falls to a lower energy level/ground
state, energy (corresponding to red light) is emitted. [1]
(ii) E = Lhc/1000λ
or {(6.02 × 1023) × (6.63 × 10–34) ×
(3 × 108)]/706 × 10–9 × (1000) [1]
= 170 kJ mol–1 [1]
(Alternatively, calculate the frequency first using c = fλ and
then calculate energy using E = Lhf.
i.e. f = 3 × 108/706 × 10–9 = 4.25 × 1014 s–1
b) 1, 0, 0, –½ [1] (In other words, everything is the same except
we put a negative sign in front of the ½ to show that the spin is
in the opposite direction.)
c) (i)
[1] (Full or half headed arrows would be acceptable.)
(ii) The (three degenerate) 3p orbitals are filled singly. or 3p
orbitals are filled in such a way as to maximise the number
of parallel spins. [1]
5 D [1]
Unit 2: Answers to Study
questions
Chapter 9 Chemical equilibrium and factors affecting equilibrium
1
2 C
3 Kc = [C]2/[A][B2] = (0.25)2/(0.025)(0.15)2 = 0.0625/0.0005625 =
111
4 Kc = [N2][O2]/[NO]2
5
a) There are 1 mol of nitrogen monoxide and 0.75 mol of nitrogen
in the equilibrium mixture.
b) Kc = [0.75][0.75]/[1]2 = 0.56
Chapter 10 Chemical equilibrium and the ionic product of water
1 pH = –log10(1.5) = 0.18
2 [H+] = 10(–pH) = 10(–9.3) = 5 × 10–10 mol l–1
3 Kw = [H+][OH–]
1.00 × 10–14 = [H+][1.8 ×10–4]
[H+] = 1.00 × 10–14/1.8 × 10–4 = 5.555 × 10–11 mol l–1
pH = –log10[H+] = –log10(5.555 × 10–11) = 10.3
Chapter 11 Strong and weak acids and bases
1 Kw = [H+][OH–]
5.48 × 10–14 = [H+][0.165]
[H+] = 5.48 × 10–14/0.165 = 3.321 × 10–13 mol l–1
pH = –log10[H+] = –log10[3.321 × 10–13] = 12.5
2 [H+] = 10(–pH) = 10(–0.73) = 0.186 mol l–1
3 [H+] = 2 × 0.250 = 0.5
pH = –log10[H+] = –log10(0.5) = 0.3
Chapter 12 Acid dissociation constant and salts
1 A sodium sulfite solution is a solution of a strong base and a salt
of a weak acid. When sodium sulfite is in solution, it forms an
alkaline solution. This is because the sulfite ions react with
hydrogen ions in the water equilibrium to form sulfurous acid
molecules. This means there are now fewer H+ ions than OH– ions
and so the salt solution will have a pH greater than 7.
2 Step 1: Look up pKa of methanoic acid on page 13 of the data
booklet; pKa = 3.75
Step 2: Input numbers into
= (½ 3.75) – ½log100.15
pH = 2.29
3 D
Chapter 13 Indicators and buffers
1 Step 1: Find the pKa of benzoic acid from page 13 of the data
booklet; pKa = 4.20
Step 2: Work out the concentration of sodium benzoate:
c = n/V = 0.02/0.2 = 0.1 mol l–1
Step 3: Input data into
2 Step 1: Find the pKa of butanoic acid from page 13 of the data
booklet; pKa = 4.83
Step 2: Using this equation, input numerical data
3 a) Nitric acid and ammonia could use methyl red, as strong
acid/weak base so end point pH will be in the acidic range.
b) Methanoic acid and potassium hydroxide could use
phenolphthalein, as weak acid/strong base so end point will be
in the alkaline range.
c) Sulfuric acid and sodium hydroxide could use bromothymol
blue, as strong acid/strong base so end point pH will be in the
neutral range.
Chapter 14 Reaction feasibility, standard enthalpy of formation and
entropy
1 C2H4(g) + 3O2(g) → 2CO2(g) + 2H2O(l)
ΔHreaction = ΣΔH°f(products) – ΣΔH°f(reactants)
ΔH° = {(2 × –394) + (2 × –286)} – {(+52) + (3 × 0)} = –1414 kJ
mol–1
2 4NH3(g) + 5O2(g) → 4NO(g) + 6H2O(g)
ΔHreaction = ΣΔH°f(products) – ΣΔH°f(reactants)
ΔH° = {(4 × +90) + (6 × –242)} – {(4 × –46) + (5 × 0)} = –908 kJ
mol–1
3 3Fe(s) + 2O2(g) → Fe3O4(s)
ΔS° = ΣS°(products) – ΣS°(reactants)
ΔS° = {+146} – {(3 × +27) + (2 × +205)} = –345 J K–1 mol–1
4 a) NH4NO3(s) → N2O(g) + 2H2O(g): a positive change in entropy
as disorder is increasing
b) CaO(s) + H2O(l) → Ca(OH)2(s): a negative change in entropy
as disorder is decreasing
c) 2NaHCO3(s) → Na2CO3(s) + CO2(g) + H2O(l): a positive
change in entropy as disorder is increasing
Chapter 15 Gibbs free energy
1 a) ΔG° = ΔH° – TΔS° = +100 – 900 × (175/1000) = –57.5 kJ mol–1
b) Both ΔH° and ΔS° are positive so when ΔG° = 0
ΔS° = +175/1000 = 0.160 kJ K–1 mol–1, therefore
T = ΔH°/ΔS° = 100/0.175 = 571 K
The reaction is spontaneous thermodynamically above 571 K.
2 T = ΔH°/ΔS° = 210/0.185 = 1135 K
The reaction is spontaneous thermodynamically above 1135 K.
3 ΔG° = ΔH° – TΔS°
+244 – 298 × (243/1000) = –171.6 kJ mol–1
The reaction is feasible at 298 K.
Chapter 16 Kinetics
1 B
2 Rate = k[H2]
k = Rate/[H2] = 3.20 × 10–4/0.0075 = 0.043 s–1
Unit 2: Answers to Exam-style
questions
1 A [1]
2 B [1]
3 a) H2O(l) ⇋ H+(aq) + OH–(aq)
Methanoate ions react with H+ ions from water’s equilibrium. [1]
This causes the water equilibrium to shift to the right-hand side
producing excess OH– and hence pH > 7. [1] or Produces an
excess of OH–.
b)
4 a) ΔS° = ΣS° products – ΣS° reactants
= (213.8 + 72.1) – (112) [1] = 173.9 J K–1 mol–1 [1]
b) ΔG° = ΔH° – TΔS°
Just feasible when ΔG° = 0 [1]
ΔH° – TΔS° = 0
TΔS° = ΔH°
T = 266 000/173.9 [1]
= 1529.6 K [1]
Unit 3: Answers to Study
questions
Chapter 17 Molecular orbitals and hybridisation
1 14 sigma bonds and 1 pi bond
2 sp3 hybridisation with sigma bonds between the C atoms, and
sigma bonds between hydrogen 1s orbital and carbon sp3 orbitals
Chapter 18 Molecular orbitals and colour
1 A conjugated system
2 Naphthacene
3 β-carotene
4 Green
Chapter 19 Homolytic and heterolytic fission
1
2 Carbocation
Chapter 20 Electrophiles, nucleophiles and reaction types
1 Addition
2 Oxygen has a slight negative charge so this is the nucleophile
centre.
Chapter 21 Alkenes
1 a Electrophilic addition/hydrohalogenation
b Carbocation
2
Chapter 22 Haloalkanes including SN1 and SN2 mechanisms
1 C
2 Methyloxybutane
3 Elimination
Chapter 23 Alcohols
1 D
2 Propan-2-ol
3 Nucleophilic substitution
Chapter 24 Carboxylic acids
1 A
2 D
3 C
Chapter 25 Ethers
1 Ethoxypentane
2 An alkali metal (sodium or potassium)
3 Methoxyhexane
Chapter 26 Amines
1 C
2 A
Chapter 27 Benzene
1 C
2 B
3 Reflux at 50 °C with concentrated sulfuric and nitric acid
Chapter 28 Stereochemistry (structural, geometric and optical
isomers)
1
2 Put into plane-polarised light; if no optical activity then it is a
racemic mixture.
Chapter 29 Experimental determination of structure using empirical
microanalysis and mass spectrometry
1
3 D
Chapter 30 Experimental determination of structure using NMR and
infrared
1 Propanone has the same CH3 group on both sides of the carbonyl
group so the three 1H nuclei in each of the methyl groups are all in
the same environment. Because all six 1H nuclei are in the same
environment, due to the symmetry of the molecule, only one peak
is observed.
2 3230–3550 –OH peak and compare fingerprint region to database
Chapter 31 Pharmaceutical chemistry
1 C
2 Omeprazole is an antagonist. It produces no response because it
prevents the action of the body’s natural active compound.
3 Dose of lidocaine: max = 3 × 35 = 105 mg
Volume of lidocaine: max = 105/5 = 21 cm3
Unit 3: Answers to Exam-style
questions
1 C [1]
2 C [1]
3 a) (i) 3.52 g CO2 contains 0.96 g C
1.44 g H2O contains 0.16 g H [both for 1 mark]
Therefore, 0.64 g O [1]
(ii) C = 0.96/12 = 0.08; H = 0.16/1 = 0.16; O = 0.64/16 = 0.04
Ratio = 2 : 4 : 1
Empirical formula is therefore C2H4O. [1]
b) C4H8O2 [1]
4 a)
[1 mark for correct structural formulae for both reactant and
product; 1 mark for intermediate]
b) 1-bromobutane would form a primary carbocation which is too
unstable. [1]
Unit 4: Answers to Study
questions
Chapter 32 Errors and significant figures
1 n = 2.2/23 = 0.096 mol (2 s.f.)
2 a) 5.38
b) 0.00025
3 Percentage error = {(2 × 0.05)/12.45} × 100 = 0.80%
Chapter 33 Practical skills and techniques
1 Number of moles of salicylic acid = m/GFM = 9.6/138 = 0.070 mol
Theoretical mass of aspirin = n × GFM = 0.070 × 180 = 12.5 g
% yield = actual mass/theoretical mass × 100 = 8.1/12.5 × 100 =
65%
2 a) Propanol, ethanoic acid, concentrated sulfuric acid
b) The organic compounds involved are flammable.
c) To avoid loss of the volatile reactants/products
3 By distillation: a mixture of these alcohols should be heated using
an electric heater in distillation apparatus (anti-bumping granules
would also be added). The butan-2-ol would boil first, at 100 oC.
This would be condensed and collected. The butan-1-ol would boil
second, at 118 oC. This would be condensed and collected in a
separate container.
Chapter 34 Standard solutions and dilutions
1 Number of moles required = c × V = 0.15 × 0.25 = 0.0375 mol
Mass of hydrated sodium bicarbonate (NaHCO3.2H2O)
= n × GFM = 0.0375 × 120 = 4.5 g
2 C1V1 = C2V2 can be used, where
C1 = 0.5 mol l–1 and V1 = 0.005, C2 = ? mol l–1 and V2 = 1 l
0.5 × 0.005 = C2 × 1
C2 = 0.0025/1 = 0.0025 mol l–1
3 C1V1 = C2V2 can be used, where
C1 = 0.2 mol l–1 and V1 = ?, C2 = 0.06 mol l–1 and V2 = 250 cm3
0.2 × V1 = 0.06 × 0.250
V1 = 0.075 l or 75 cm3. But the question asks for sodium ion
concentration. The formula for sodium phosphate is Na3PO4.
There are three Na+ ions present so we need to divide our volume
by 3, so 0.075 l/3 = 0.025 l or 25 cm3 stock solution should be
pipetted into a 250 cm3 standard flask.
4 H2SO4 GFM = 98.1. Concentration in mol l–1 = concentration in g
l–1/GFM = 35/98.1 = 0.357 mol l–1
Chapter 35 Volumetric analysis
1 KOH + HNO3 → KNO3 + H2O
Moles of KOH = c × V = 0.2 × 0.025 = 0.005 mol
Concentration of HNO3 = n/V = 0.005/0.0115 = 0.435 mol l–1
2 n of NaOH (initial) = c × V = 0.110 × 0.025 = 0.00275 mol
n of HCl = c × V = 0.0100 × 0.015 = 0.00015 mol
n of NaOH = 0.0015 mol
n of unreacted NaOH in standard flask = 0.00015 × 10 = 0.0015
mol
Reacted NaOH = 0.00275 – 0.0015 = 0.00125 mol
So n of aspirin = 0.000625 mol
Mass of aspirin in tablet = n × GFM = 0.000625 × 180 = 0.1125 g
Chapter 36 Gravimetric analysis
1 Moles of silver chloride = 2.5/143.4 = 0.0174 mol
Mass of silver in jewellery = 0.0174 × 107.9 = 1.88 g
% silver in jewellery =1.88/5.05 × 100 = 37%
2 Fe + HCl → FeCl2 + H2
Moles of iron(II) chloride = m/GFM = 12.8/126.8 = 0.101 mol
Mass of Fe in nail = n × GFM = 0.101 × 55.8 = 5.64 g
% Fe in alloy = 5.64/7.45 × 100 = 76%
3 Mass of water present = 0.966 – 0.750 = 0.216 g
Moles of Cu(NO3)2 = m/GFM = 0.750/187.5 = 0.004 mol
Moles of water = m/GFM = 0.216/18 = 0.012 mol
Ratio is 0.004 : 0.012 or 1 : 3 so formula is Cu(NO3)2.3H2O
Chapter 37 Chromatography
1 C
2 C
Chapter 38 Colorimetry
1 Solutions of accurate iron concentration are made up and the
absorbance measured.
A calibration graph of concentration against absorbance is made.
The absorbance of the iron tablet solution is measured and the
calibration graph is then used to determine the concentration
corresponding to this absorbance.
2 Red
Unit 4: Answers to Exam-style
questions
1 B [1]
2 a) Condenser, heating mantle [2]
b) (i) To purify (the sulfanilamide); to get rid of impurities [1]
(ii) Insoluble in cold water; soluble in hot water [1]
(iii) Vacuum filtration [1]
c) The sample is mixed with pure sulfanilamide. [1] The melting
point of the mixture will be the same. [1]
d) IR, NMR or TLC [1]