Sunteți pe pagina 1din 2267

chemguide: helping you to understand Chemistry - Main Menu

chemguide
Helping you to understand Chemistry
You can search Chemguide either using a keyword
search (courtesy of Google) or the menu system.

See the bottom of


this page for the
latest additions
and updates.

Keyword searching
Google Search

Search www

Search chemguide

Notes: If you don't find what you want, use the BACK button
on your browser to return to this page.
Remember that this is a UK site, so spellings will be English.
Take care if you want to search for words like sulphur,
sulphate, sulphuric acid or aluminium.

Searching using the menus

Atomic Structure and Bonding

http://www.chemguide.co.uk/index.html (1 of 4)30/12/2004 11:00:02

Covers basic atomic properties


(electronic structures, ionisation
energies, electron affinities, atomic
and ionic radii), bonding (including
intermolecular bonding) and
structures (ionic, molecular, giant
covalent and metallic).

chemguide: helping you to understand Chemistry - Main Menu

Inorganic Chemistry

Includes essential ideas about redox


reactions, and covers the trends in
Groups 2, 4 and 7 of the Periodic
Table. Plus: lengthy sections on the
chemistry of some important
complex ions, and of common
transition metals.

Physical Chemistry

Covers rates of reaction including


catalysis, an introduction to chemical
equilibria, redox equilibria, acid-base
equilibria (pH, buffer solutions,
indicators, etc) and phase equilibria
(including Raoult's Law and the use
of various phase diagrams).

Instrumental analysis

Explains how you can analyse


substances using machines - mass
spectrometry, infra-red spectroscopy
and NMR.

Basic Organic Chemistry

Includes help on bonding, naming


and isomerism, and a discussion of
organic acids and bases.

Properties of organic compounds

Covers the physical and chemical


properties of compounds on UK A
level chemistry syllabuses.

Organic Reaction Mechanisms

Covers all the mechanisms required


by the current UK A level chemistry
syllabuses.

About this site

Includes a contact address if you


have found any difficulties with the
site.

Chemistry Calculations

A description of the author's book on


calculations at UK A level chemistry
standard.

http://www.chemguide.co.uk/index.html (2 of 4)30/12/2004 11:00:02

chemguide: helping you to understand Chemistry - Main Menu

Textbook suggestions

Suggestions for textbooks and


revision guides covering the UK AS
and A level chemistry syllabuses,
with links to Amazon.co.uk if you
want to follow them up.

Download syllabuses

For UK A level students. Download a


copy of your current syllabus from
your Exam Board.

Test yourself

A link to Dr Phil Brown's website


where UK GCSE, AS and A level
chemistry students will find a wide
and growing range of multiple
choice, short answer and structured
questions.

An examiner's view

A link to Rod Beavon's chemistry


pages. Rod Beavon is chief
examiner for A level chemistry for
the UK exam board Edexcel. A close
look at what he has to offer is a must
for Edexcel students, but there is a
lot of good stuff whatever exam
system you are working in.

Latest additions and important updates

5/12/2004

There is now the beginnings of a section on phase


equilibria. I'm working on it at the moment and it is unlikely
to be finished before the end of January. It currently deals
with vapour pressure, phase diagrams for pure
substances and for solutions of non-volatile solutes
(including the effect of the solute on the boiling point and
freezing point of the solvent).

18/12/2004

An introduction to phase diagrams involving eutectic


mixtures is now available using the tin-lead system.

http://www.chemguide.co.uk/index.html (3 of 4)30/12/2004 11:00:02

chemguide: helping you to understand Chemistry - Main Menu

Jim Clark 2004

http://www.chemguide.co.uk/index.html (4 of 4)30/12/2004 11:00:02

Understanding Chemistry - Atomic Structure and Bonding Menu

Understanding Chemistry

ATOMIC STRUCTURE AND BONDING MENU

Basic atomic properties . . .


Includes a discussion of orbitals, electronic structures of atoms
and ions, ionisation energies, electron affinities, atomic and ionic
radii.
Bonding . . .
Includes ionic, covalent, co-ordinate (dative covalent) and metallic
bonding as well as intermolecular attractions like Van der Waals
forces and hydrogen bonding. Also includes full discussions of
electronegativity and shapes of molecules and ions.
Types of structure . . .
Describes and explains how the various types of structure (ionic,
giant covalent, metallic, and molecular) affect physical properties.

Go to Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/atommenu.html30/12/2004 11:00:04

Atomic properties menu

Understanding Chemistry

ATOMIC PROPERTIES MENU

Simple background . . .
Revises the simple knowledge you should already have about the
structure of atoms from introductory courses (e.g. GCSE).
Atomic orbitals . . .
Explains what atomic orbitals are and discusses their shapes and
relative energies. This is essential pre-reading before you go on to
any of the remaining topics in this section.
Electronic structures . . .
How to work out and write the electronic structures for atoms and
simple monatomic ions (containing only one atom - e.g. Cl- or Mg2
+) using s, p, d notation.
Ionisation energies . . .
Explains what ionisation energies are and how and why they vary
around the Periodic Table.
Electron affinities . . .
Explains what electron affinities are and how and why they vary
around the Periodic Table.
Atomic and ionic radii . . .

http://www.chemguide.co.uk/atoms/propsmenu.html (1 of 2)30/12/2004 11:00:05

Atomic properties menu

Looks at the various measures of atomic radius, and explains how


and why atomic radii vary around the Periodic Table. Also
considers how the radii of positive and negative ions differ from the
atoms they come from.

Go to atomic structure and bonding menu . . .


Go to Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/atoms/propsmenu.html (2 of 2)30/12/2004 11:00:05

a simple view of atomic structure

A SIMPLE VIEW OF ATOMIC STRUCTURE

This page revises the simple ideas about atomic structure that you will
have come across in an introductory chemistry course (for example,
GCSE). You need to be confident about this before you go on to the
more difficult ideas about the atom which under-pin A'level chemistry.

The sub-atomic particles


Protons, neutrons and electrons.

proton
neutron
electron

relative mass
1
1
1/1836

relative charge
+1
0
-1

Beyond A'level: Protons and neutrons don't in fact have


exactly the same mass - neither of them has a mass of
exactly 1 on the carbon-12 scale (the scale on which the
relative masses of atoms are measured). On the carbon-12
scale, a proton has a mass of 1.0073, and a neutron a mass
of 1.0087.

http://www.chemguide.co.uk/atoms/properties/gcse.html (1 of 7)30/12/2004 11:00:08

a simple view of atomic structure

The nucleus
The nucleus is at the centre of the atom and contains the protons and
neutrons. Protons and neutrons are collectively known as nucleons.
Virtually all the mass of the atom is concentrated in the nucleus, because
the electrons weigh so little.

Working out the numbers of protons and neutrons


No of protons = ATOMIC NUMBER of the atom
The atomic number is also given the more descriptive name of proton
number.
No of protons + no of neutrons = MASS NUMBER of the atom
The mass number is also called the nucleon number.

This information can be given simply in the form:

How many protons and neutrons has this atom got?


The atomic number counts the number of protons (9); the mass number
counts protons + neutrons (19). If there are 9 protons, there must be 10
neutrons for the total to add up to 19.

The atomic number is tied to the position of the element in the Periodic
Table and therefore the number of protons defines what sort of element
http://www.chemguide.co.uk/atoms/properties/gcse.html (2 of 7)30/12/2004 11:00:08

a simple view of atomic structure

you are talking about. So if an atom has 8 protons (atomic number = 8), it
must be oxygen. If an atom has 12 protons (atomic number = 12), it must
be magnesium.
Similarly, every chlorine atom (atomic number = 17) has 17 protons;
every uranium atom (atomic number = 92) has 92 protons.

Isotopes
The number of neutrons in an atom can vary within small limits. For
example, there are three kinds of carbon atom 12C, 13C and 14C. They all
have the same number of protons, but the number of neutrons varies.

carbon-12
carbon-13
carbon-14

protons
6
6
6

neutrons
6
7
8

mass number
12
13
14

These different atoms of carbon are called isotopes. The fact that they
have varying numbers of neutrons makes no difference whatsoever to
the chemical reactions of the carbon.
Isotopes are atoms which have the same atomic number but different
mass numbers. They have the same number of protons but different
numbers of neutrons.

The electrons
Working out the number of electrons
Atoms are electrically neutral, and the positiveness of the protons is
balanced by the negativeness of the electrons. It follows that in a neutral
atom:
no of electrons = no of protons

http://www.chemguide.co.uk/atoms/properties/gcse.html (3 of 7)30/12/2004 11:00:08

a simple view of atomic structure

So, if an oxygen atom (atomic number = 8) has 8 protons, it must also


have 8 electrons; if a chlorine atom (atomic number = 17) has 17
protons, it must also have 17 electrons.
The arrangement of the electrons
The electrons are found at considerable distances from the nucleus in a
series of levels called energy levels. Each energy level can only hold a
certain number of electrons. The first level (nearest the nucleus) will only
hold 2 electrons, the second holds 8, and the third also seems to be full
when it has 8 electrons. At GCSE you stop there because the pattern
gets more complicated after that.
These levels can be thought of as getting progressively further from the
nucleus. Electrons will always go into the lowest possible energy level
(nearest the nucleus) - provided there is space.
To work out the electronic arrangement of an atom

Look up the atomic number in the Periodic Table - making sure


that you choose the right number if two numbers are given. The
atomic number will always be the smaller one.
This tells you the number of protons, and hence the number of
electrons.
Arrange the electrons in levels, always filling up an inner level
before you go to an outer one.

e.g. to find the electronic arrangement in chlorine

The Periodic Table gives you the atomic number of 17.

Therefore there are 17 protons and 17 electrons.

The arrangement of the electrons will be 2, 8, 7 (i.e. 2 in the first


level, 8 in the second, and 7 in the third).

The electronic arrangements of the first 20 elements

http://www.chemguide.co.uk/atoms/properties/gcse.html (4 of 7)30/12/2004 11:00:08

a simple view of atomic structure

After this the pattern alters as you enter the transition series in the
Periodic Table.
Two important generalisations
If you look at the patterns in this table:

The number of electrons in the outer level is the same as the


group number. (Except with helium which has only 2 electrons.
The noble gases are also usually called group 0 - not group 8.)
This pattern extends throughout the Periodic Table for the main
groups (i.e. not including the transition elements).
So if you know that barium is in group 2, it has 2 electrons in its
outer level; iodine (group 7) has 7 electrons in its outer level; lead
(group 4) has 4 electrons in its outer level.

Noble gases have full outer levels. This generalisation will need
modifying for A'level purposes.

Dots-and-crosses diagrams
In any introductory chemistry course you will have come across the
electronic structures of hydrogen and carbon, for example, drawn as:

http://www.chemguide.co.uk/atoms/properties/gcse.html (5 of 7)30/12/2004 11:00:08

a simple view of atomic structure

Note: There are many places where you could still make use
of this model of the atom at A'level. It is, however, a
simplification and can be misleading. It gives the impression
that the electrons are circling the nucleus in orbits like planets
around the sun. As you will find when you look at the A'level
view of the atom, it is impossible to know exactly how they
are actually moving.

The circles show energy levels - representing increasing distances from


the nucleus. You could straighten the circles out and draw the electronic
structure as a simple energy diagram.
Carbon, for example, would look like this:

Thinking of the arrangement of the electrons in this way makes a useful


bridge to the A'level view.

Note: If you have come to this page as a UK GCSE student


(or a student on a similar introductory chemistry course
elsewhere) and want some more help, you may be interested
in my GCSE Chemistry book. This link will take you to a page
describing it.

http://www.chemguide.co.uk/atoms/properties/gcse.html (6 of 7)30/12/2004 11:00:08

a simple view of atomic structure

Where would you like to go now?


To the atomic properties menu . . .
To the atomic structure and bonding menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/atoms/properties/gcse.html (7 of 7)30/12/2004 11:00:08

GCSE chemistry

Understanding Chemistry

GCSE CHEMISTRY
This book covers the chemistry
content of all the UK GCSE
Chemistry syllabuses - whether
as a part of dual award science,
or as a separate science. It is
aimed at students likely to
achieve grades from A* to B.
If you are working in another
system, GCSE in the UK is an
exam taken at the end of a
(usually) two year course at the
age of 16. Anyone taking a
similar introductory chemistry
course may find the book helpful.
On this page you will find a
description of how the book is organised, together with a summary of the
contents. You will also find direct links to the book on both the Longman
and the Amazon.co.uk sites.
Education in Chemistry, May 2003
"I was impressed with this new book, . . ."
"The text is clearly laid out with excellent diagrams and
illustrations."
"This is an excellent textbook."
School Science Review (issue 307)
". . will stretch and enthuse those with some ability in chemistry."
http://www.chemguide.co.uk/gcsebook.html (1 of 5)30/12/2004 11:00:11

GCSE chemistry

"It would certainly help to bridge the gap between GCSE and AS
level."

How to get hold of the book


Schools or colleges would probably find it best to go to the Longman
GCSE Chemistry website, but this site isn't really set up for individual
purchases.
You can, of course, buy the book through normal book sellers, but if you
want to buy online, you will find a direct link to Amazon.co.uk coming up.
Non-UK students can also buy the book from Amazon.co.uk, but will
obviously have to pay a slightly higher delivery charge.

Note: If your usual source of books is Amazon.com, you


should compare the price for the book (including delivery)
from Amazon.com with the price from Amazon.co.uk - even if
you live in North America.
You may well find that it is significantly cheaper to buy from
Amazon.co.uk and have it sent by air mail across the Atlantic,
than to buy it in America!

http://www.chemguide.co.uk/gcsebook.html (2 of 5)30/12/2004 11:00:11

GCSE chemistry

Have a look at the book on the Amazon site

What the book covers


The book is organised into 6 sections plus an important appendix. Each
section is made up of a number of related chapters.
There are questions at the end of each chapter to test understanding,
and a set of GCSE-style exam questions at the end of each section.
Answers to all the questions are provided on the supporting website although these are password-protected so that only teachers can get at
them!
Section A: Particles
This covers an introduction to atomic structure and bonding (including
metallic bonding and intermolecular forces) and the relationship between
the structures of elements and compounds and their physical properties.
There is also a chapter on how to write formulae and equations, and a
final one on the factors affecting rates of reaction together with
explanations.
Section B: Some essential background chemistry
This is a lengthy section which covers the important lab-based chemistry:

Reactivity series

Acids and their reactions

Salts

Simple analysis

Periodic Table: including some history, the structure of the table,


the noble gases, Groups 1 and 7, and an introduction to transition

http://www.chemguide.co.uk/gcsebook.html (3 of 5)30/12/2004 11:00:11

GCSE chemistry

metals

Electrolysis and electrochemical cells

Energy changes in reactions

Section C: Large scale chemistry


This covers the extraction of several metals, and the chemistry of salt
and limestone. It introduces reversible reactions leading to the Haber and
Contact Processes.
Section D: Air, water and earth
Discusses the atmosphere (including its evolution and some
environmental problems), water (including hardness, water treatment,
and an introduction to colloids) and types of rock.
Section E: Organic chemistry
An introductory look at the oil industry and some simple organic
compounds (alkanes, alkenes, alcohols, carboxylic acids, and a brief
look at esters). Structural isomerism is explained where it arises.
There are also chapters on food and drugs, and enzymes.
Section F: Sums
This section deals with all the calculations involving relative atomic
masses and moles up to and including simple titration and electrolysis
calculations.
Appendices
The most important appendix explains how to maximise your score when
writing up coursework practical investigations to satisfy the requirements
of UK GCSE examiners. The fully written out investigation is available
from the website accompanying the book. (See below.)
The website

http://www.chemguide.co.uk/gcsebook.html (4 of 5)30/12/2004 11:00:11

GCSE chemistry

There is a website to accompany the book which you can find by


following this link. You may find this useful even if you don't end up
buying the book!
You will find lots of links to other other useful chemistry web sites, a fully
written up example of a coursework investigation, and a set of
worksheets. Answers to all the questions in the book are available, but
only to teachers who have purchased the book from Longman. The
answers are password-protected for obvious reasons!

Go to Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/gcsebook.html (5 of 5)30/12/2004 11:00:11

atomic orbitals

ATOMIC ORBITALS

This page explains what an atomic orbital is. It explores s and p orbitals
in some detail, including their shapes and energies. d orbitals are
described only in terms of their energy, and f orbitals only get a passing
mention.

What is an atomic orbital?


Orbitals and orbits
When the a planet moves around the sun, you can plot a definite path for
it which is called an orbit. A simple view of the atom looks similar and you
may have pictured the electrons as orbiting around the nucleus. The truth
is different, and electrons in fact inhabit regions of space known as
orbitals.
Orbits and orbitals sound similar, but they have quite different meanings.
It is essential that you understand the difference between them.
The impossibility of drawing orbits for electrons
To plot a path for something you need to know exactly where the object
is and be able to work out exactly where it's going to be an instant later.
You can't do this for electrons.
The Heisenberg Uncertainty Principle (not required at A'level) says loosely - that you can't know with certainty both where an electron is and
where it's going next. That makes it impossible to plot an orbit for an
electron around a nucleus. Is this a big problem? No. If something is
impossible, you have to accept it and find a way around it.
Hydrogen's electron - the 1s orbital

http://www.chemguide.co.uk/atoms/properties/atomorbs.html (1 of 8)30/12/2004 11:00:17

atomic orbitals

Note: In this diagram (and the orbital diagrams that follow),


the nucleus is shown very much larger than it really is. This is
just for clarity.

Suppose you had a single hydrogen atom and at a


particular instant plotted the position of the one
electron. Soon afterwards, you do the same thing,
and find that it is in a new position. You have no
idea how it got from the first place to the second.
You keep on doing this over and over again, and
gradually build up a sort of 3D map of the places that the electron is likely
to be found.
In the hydrogen case, the electron can be found anywhere within a
spherical space surrounding the nucleus. The diagram shows a crosssection through this spherical space.
95% of the time (or any other percentage you choose), the electron will
be found within a fairly easily defined region of space quite close to the
nucleus. Such a region of space is called an orbital. You can think of an
orbital as being the region of space in which the electron lives.

Note: If you wanted to be absolutely 100% sure of where the


electron is, you would have to draw an orbital the size of the
Universe!

http://www.chemguide.co.uk/atoms/properties/atomorbs.html (2 of 8)30/12/2004 11:00:17

atomic orbitals

What is the electron doing in the orbital? We don't know, we can't know,
and so we just ignore the problem! All you can say is that if an electron is
in a particular orbital it will have a particular definable energy.
Each orbital has a name.
The orbital occupied by the hydrogen electron is called a 1s orbital. The
"1" represents the fact that the orbital is in the energy level closest to the
nucleus. The "s" tells you about the shape of the orbital. s orbitals are
spherically symmetric around the nucleus - in each case, like a hollow
ball made of rather chunky material with the nucleus at its centre.
The orbital on the left is a 2s orbital. This is
similar to a 1s orbital except that the region
where there is the greatest chance of
finding the electron is further from the
nucleus - this is an orbital at the second
energy level.
If you look carefully, you will notice that
there is another region of slightly higher
electron density (where the dots are thicker)
nearer the nucleus. ("Electron density" is another way of talking about
how likely you are to find an electron at a particular place.)
2s (and 3s, 4s, etc) electrons spend some of their time closer to the
nucleus than you might expect. The effect of this is to slightly reduce the
energy of electrons in s orbitals. The nearer the nucleus the electrons
get, the lower their energy.
3s, 4s (etc) orbitals get progressively further from the nucleus.
p orbitals

http://www.chemguide.co.uk/atoms/properties/atomorbs.html (3 of 8)30/12/2004 11:00:17

atomic orbitals

Not all electrons inhabit s orbitals (in fact, very few


electrons live in s orbitals). At the first energy level,
the only orbital available to electrons is the 1s
orbital, but at the second level, as well as a 2s
orbital, there are also orbitals called 2p orbitals.
A p orbital is rather like 2 identical balloons tied
together at the nucleus. The diagram on the right is
a cross-section through that 3-dimensional region
of space. Once again, the orbital shows where
there is a 95% chance of finding a particular
electron.

Beyond A'level: If you imagine a horizontal plane through


the nucleus, with one lobe of the orbital above the plane and
the other beneath it, there is a zero probability of finding the
electron on that plane. So how does the electron get from
one lobe to the other if it can never pass through the plane of
the nucleus? For A'level chemistry you just have to accept
that it does! If you want to find out more, read about the wave
nature of electrons.

Unlike an s orbital, a p orbital points in a particular direction - the one


drawn points up and down the page.
At any one energy level it is possible to have three absolutely equivalent
p orbitals pointing mutually at right angles to each other. These are
arbitrarily given the symbols px, py and pz. This is simply for
convenience - what you might think of as the x, y or z direction changes
constantly as the atom tumbles in space.

http://www.chemguide.co.uk/atoms/properties/atomorbs.html (4 of 8)30/12/2004 11:00:17

atomic orbitals

The p orbitals at the second energy level are


called 2px, 2py and 2pz. There are similar
orbitals at subsequent levels - 3px, 3py, 3pz,
4px, 4py, 4pz and so on.
All levels except for the first level have p
orbitals. At the higher levels the lobes get
more elongated, with the most likely place to
find the electron more distant from the
nucleus.

d and f orbitals
In addition to s and p orbitals, there are two other sets of orbitals which
become available for electrons to inhabit at higher energy levels. At the
third level, there is a set of five d orbitals (with complicated shapes and
names) as well as the 3s and 3p orbitals (3px, 3py, 3pz). At the third level
there are a total of nine orbitals altogether.
At the fourth level, as well the 4s and 4p and 4d orbitals there are an
additional seven f orbitals - 16 orbitals in all. s, p, d and f orbitals are then
available at all higher energy levels as well.
For A'level purposes, you have to be aware that there are sets of five d
orbitals at levels from the third level upwards, but you will not be
expected to draw them or name them. Apart from a passing reference,
you won't come across f orbitals at all.

Fitting electrons into orbitals


You can think of an atom as a very bizarre house (like an inverted
pyramid!) - with the nucleus living on the ground floor, and then various
rooms (orbitals) on the higher floors occupied by the electrons. On the
first floor there is only 1 room (the 1s orbital); on the second floor there
are 4 rooms (the 2s, 2px, 2py and 2pz orbitals); on the third floor there
are 9 rooms (one 3s orbital, three 3p orbitals and five 3d orbitals); and so
on. But the rooms aren't very big . . . Each orbital can only hold 2
http://www.chemguide.co.uk/atoms/properties/atomorbs.html (5 of 8)30/12/2004 11:00:17

atomic orbitals

electrons.
A convenient way of showing the orbitals that the electrons live in is to
draw "electrons-in-boxes".
"Electrons-in-boxes"
Orbitals can be represented as boxes with the electrons in them shown
as arrows. Often an up-arrow and a down-arrow are used to show that
the electrons are in some way different.

Beyond A'level: The need to have all electrons in an atom


different comes out of quantum theory. If they live in different
orbitals, that's fine - but if they are both in the same orbital
there has to be some subtle distinction between them.
Quantum theory allocates them a property known as "spin" which is what the arrows are intended to suggest.

A 1s orbital holding 2 electrons would be drawn as shown


on the right, but it can be written even more quickly as 1s2.
This is read as "one s two" - not as "one s squared".
You mustn't confuse the two numbers in this notation:

The order of filling orbitals


Electrons fill low energy orbitals (closer to the nucleus) before they fill
higher energy ones. Where there is a choice between orbitals of equal
energy, they fill the orbitals singly as far as possible.
This filling of orbitals singly where possible is known as Hund's rule. It
only applies where the orbitals have exactly the same energies (as with p
orbitals, for example), and helps to minimise the repulsions between
electrons and so makes the atom more stable.

http://www.chemguide.co.uk/atoms/properties/atomorbs.html (6 of 8)30/12/2004 11:00:17

atomic orbitals

The diagram (not to scale) summarises the energies of the orbitals up to


the 4p level.

Notice that the s orbital always has a slightly lower energy than the p
orbitals at the same energy level, so the s orbital always fills with
electrons before the corresponding p orbitals.
The real oddity is the position of the 3d orbitals. They are at a slightly
higher level than the 4s - and so it is the 4s orbital which will fill first,
followed by all the 3d orbitals and then the 4p orbitals. Similar confusion
occurs at higher levels, with so much overlap between the energy levels
that the 4f orbitals don't fill until after the 6s, for example.
For A'level purposes you simply have to remember that the 4s orbital fills
before the 3d orbitals. The same thing happens at the next level as well the 5s orbital fills before the 4d orbitals. All the other complications are
beyond A'level.
Knowing the order of filling is central to understanding how to write
electronic structures. Follow the link below to find out how to do this.

Where would you like to go now?


To look at how to write electronic structures . . .
To the atomic properties menu . . .
http://www.chemguide.co.uk/atoms/properties/atomorbs.html (7 of 8)30/12/2004 11:00:17

atomic orbitals

To the atomic structure and bonding menu . . .


To Main Menu . . .

Jim Clark 2000 (modified 2004)

http://www.chemguide.co.uk/atoms/properties/atomorbs.html (8 of 8)30/12/2004 11:00:17

electronic structures of atoms

ELECTRONIC STRUCTURES

This page explores how you write electronic structures for atoms using s,
p, and d notation. It assumes that you know about simple atomic orbitals
- at least as far as the way they are named, and their relative energies. If
you want to look at the electronic structures of simple monatomic ions
(such as Cl-, Ca2+ and Cr3+), you will find a link at the bottom of the page.

Important! If you haven't already read the page on atomic


orbitals you should follow this link before you go any further.

The electronic structures of atoms


Relating orbital filling to the Periodic Table

Most A'level syllabuses stop at krypton when it comes to writing


electronic structures, but it is possible that you could be asked for
structures for elements up as far as barium. After barium you have to
worry about f orbitals as well as s, p and d orbitals - and that's a problem
beyond A'level. It is important that you look through past exam papers as
well as your syllabus so that you can judge how hard the questions are
likely to get.
This page looks in detail at the elements in the shortened version of the
Periodic Table above, and then shows how you could work out the
structures of some bigger atoms.

http://www.chemguide.co.uk/atoms/properties/elstructs.html (1 of 9)30/12/2004 11:00:22

electronic structures of atoms

Important! You must have a copy of your syllabus and


copies of recent exam papers. If you haven't got them, follow
this link to find out how to get hold of them.

The first period


Hydrogen has its only electron in the 1s orbital - 1s1, and at helium the
first level is completely full - 1s2.
The second period
Now we need to start filling the second level, and hence start the second
period. Lithium's electron goes into the 2s orbital because that has a
lower energy than the 2p orbitals. Lithium has an electronic structure of
1s22s1. Beryllium adds a second electron to this same level - 1s22s2.
Now the 2p levels start to fill. These levels all have the same energy, and
so the electrons go in singly at first.
B

1s22s22px1

1s22s22px12py1

1s22s22px12py12pz1
Note: The orbitals where something new is happening are
shown in bold type. You wouldn't normally write them any
differently from the other orbitals.

http://www.chemguide.co.uk/atoms/properties/elstructs.html (2 of 9)30/12/2004 11:00:22

electronic structures of atoms

The next electrons to go in will have to pair up with those already there.
O

1s22s22px22py12pz1

1s22s22px22py22pz1

Ne

1s22s22px22py22pz2

You can see that it is going to get progressively tedious to write the full
electronic structures of atoms as the number of electrons increases.
There are two ways around this, and you must be familiar with both.
Shortcut 1: All the various p electrons can be lumped together. For
example, fluorine could be written as 1s22s22p5, and neon as 1s22s22p6.
This is what is normally done if the electrons are in an inner layer. If the
electrons are in the bonding level (those on the outside of the atom), they
are sometimes written in shorthand, sometimes in full. Don't worry about
this. Be prepared to meet either version, but if you are asked for the
electronic structure of something in an exam, write it out in full showing
all the px, py and pz orbitals in the outer level separately.
For example, although we haven't yet met the electronic structure of
chlorine, you could write it as 1s22s22p63s23px23py23pz1.
Notice that the 2p electrons are all lumped together whereas the 3p ones
are shown in full. The logic is that the 3p electrons will be involved in
bonding because they are on the outside of the atom, whereas the 2p
electrons are buried deep in the atom and aren't really of any interest.
Shortcut 2: You can lump all the inner electrons together using, for
example, the symbol [Ne]. In this context, [Ne] means the electronic
structure of neon - in other words: 1s22s22px22py22pz2 You wouldn't do
this with helium because it takes longer to write [He] than it does 1s2.
On this basis the structure of chlorine would be written [Ne]
3s23px23py23pz1.
The third period

http://www.chemguide.co.uk/atoms/properties/elstructs.html (3 of 9)30/12/2004 11:00:22

electronic structures of atoms

At neon, all the second level orbitals are full, and so after this we have to
start the third period with sodium. The pattern of filling is now exactly the
same as in the previous period, except that everything is now happening
at the 3-level.
For example:
short version
Mg

1s22s22p63s2

[Ne]3s2

1s22s22p63s23px23py13pz1

[Ne]3s23px23py13pz1

Ar

1s22s22p63s23px23py23pz2

[Ne]3s23px23py23pz2

Note: Check that you can do these. Cover the text and then
work out these structures for yourself. Then do all the rest of
this period. When you've finished, check your answers
against the corresponding elements from the previous period.
Your answers should be the same except a level further out.

The beginning of the fourth period


At this point the 3-level orbitals aren't all full - the 3d levels haven't been
used yet. But if you refer back to the energies of the orbitals, you will see
that the next lowest energy orbital is the 4s - so that fills next.
K

1s22s22p63s23p64s1

Ca

1s22s22p63s23p64s2

There is strong evidence for this in the similarities in the chemistry of


elements like sodium (1s22s22p63s1) and potassium
(1s22s22p63s23p64s1)
The outer electron governs their properties and that electron is in the
same sort of orbital in both of the elements. That wouldn't be true if the
outer electron in potassium was 3d1.
s- and p-block elements

http://www.chemguide.co.uk/atoms/properties/elstructs.html (4 of 9)30/12/2004 11:00:22

electronic structures of atoms

The elements in group 1 of the Periodic Table all have an outer


electronic structure of ns1 (where n is a number between 2 and 7). All
group 2 elements have an outer electronic structure of ns2. Elements in
groups 1 and 2 are described as s-block elements.
Elements from group 3 across to the noble gases all have their outer
electrons in p orbitals. These are then described as p-block elements.
d-block elements

Remember that the 4s orbital has a lower energy than the 3d orbitals and
so fills first. Once the 3d orbitals have filled up, the next electrons go into
the 4p orbitals as you would expect.
d-block elements are elements in which the last electron to be added to
the atom is in a d orbital. The first series of these contains the elements
from scandium to zinc, which at GCSE you probably called transition
elements or transition metals. The terms "transition element" and "d-

http://www.chemguide.co.uk/atoms/properties/elstructs.html (5 of 9)30/12/2004 11:00:22

electronic structures of atoms

block element" don't quite have the same meaning, but it doesn't matter
in the present context.

If you are interested: A transition element is defined as one


which has partially filled d orbitals either in the element or
any of its compounds. Zinc (at the right-hand end of the dblock) always has a completely full 3d level (3d10) and so
doesn't count as a transition element.

d electrons are almost always described as, for example, d5 or d8 - and


not written as separate orbitals. Remember that there are five d orbitals,
and that the electrons will inhabit them singly as far as possible. Up to 5
electrons will occupy orbitals on their own. After that they will have to pair
up.
d5 means

d8 means

Notice in what follows that all the 3-level orbitals are written together,
even though the 3d electrons are added to the atom after the 4s.
Sc

1s22s22p63s23p63d14s2

Ti

1s22s22p63s23p63d24s2

1s22s22p63s23p63d34s2

Cr

1s22s22p63s23p63d54s1

Whoops! Chromium breaks the sequence. In chromium, the electrons in


the 3d and 4s orbitals rearrange so that there is one electron in each
orbital. It would be convenient if the sequence was tidy - but it's not!
Mn

1s22s22p63s23p63d54s2

Fe

1s22s22p63s23p63d64s2

(back to being tidy again)

http://www.chemguide.co.uk/atoms/properties/elstructs.html (6 of 9)30/12/2004 11:00:22

electronic structures of atoms

Co

1s22s22p63s23p63d74s2

Ni

1s22s22p63s23p63d84s2

Cu

1s22s22p63s23p63d104s1

Zn

1s22s22p63s23p63d104s2

(another awkward one!)

And at zinc the process of filling the d orbitals is complete.


Filling the rest of period 4
The next orbitals to be used are the 4p, and these fill in exactly the same
way as the 2p or 3p. We are back now with the p-block elements from
gallium to krypton. Bromine, for example, is
1s22s22p63s23p63d104s24px24py24pz1.

Useful exercise: Work out the electronic structures of all the


elements from gallium to krypton. You can check your
answers by comparing them with the elements directly above
them in the Periodic Table. For example, gallium will have the
same sort of arrangement of its outer level electrons as boron
or aluminium - except that gallium's outer electrons will be in
the 4-level.

Summary
Writing the electronic structure of an element from hydrogen to
krypton

Use the Periodic Table to find the atomic number, and hence
number of electrons.
Fill up orbitals in the order 1s, 2s, 2p, 3s, 3p, 4s, 3d, 4p - until you
run out of electrons. The 3d is the awkward one - remember that
specially. Fill p and d orbitals singly as far as possible before
pairing electrons up.
Remember that chromium and copper have electronic structures
which break the pattern in the first row of the d-block.

http://www.chemguide.co.uk/atoms/properties/elstructs.html (7 of 9)30/12/2004 11:00:22

electronic structures of atoms

Writing the electronic structure of big s- or p-block elements

Note: We are deliberately excluding the d-block elements


apart from the first row that we've already looked at in detail.
The pattern of awkward structures isn't the same in the other
rows. This isn't an A'level problem.

First work out the number of outer electrons. This is quite likely all you
will be asked to do anyway.
The number of outer electrons is the same as the group number. (The
noble gases are a bit of a problem here, because they are normally
called group 0 rather then group 8. Helium has 2 outer electrons; the rest
have 8.) All elements in group 3, for example, have 3 electrons in their
outer level. Fit these electrons into s and p orbitals as necessary. Which
level orbitals? Count the periods in the Periodic Table (not forgetting the
one with H and He in it).
Iodine is in group 7 and so has 7 outer electrons. It is in the fifth period
and so its electrons will be in 5s and 5p orbitals. Iodine has the outer
structure 5s25px25py25pz1.
What about the inner electrons if you need to work them out as well? The
1, 2 and 3 levels will all be full, and so will the 4s, 4p and 4d. The 4f
levels don't fill until after anything you will be asked about at A'level. Just
forget about them! That gives the full structure:
1s22s22p63s23p63d104s24p64d105s25px25py25pz1.
When you've finished, count all the electrons to make sure that they
come to the same as the atomic number. Don't forget to make this check
- it's easy to miss an orbital out when it gets this complicated.
Barium is in group 2 and so has 2 outer electrons. It is in the sixth
period. Barium has the outer structure 6s2.
Including all the inner levels: 1s22s22p63s23p63d104s24p64d105s25p66s2.
http://www.chemguide.co.uk/atoms/properties/elstructs.html (8 of 9)30/12/2004 11:00:22

electronic structures of atoms

It would be easy to include 5d10 as well by mistake, but the d level


always fills after the next s level - so 5d fills after 6s just as 3d fills after
4s. As long as you counted the number of electrons you could easily spot
this mistake because you would have 10 too many.

Note: Don't worry too much about these complicated


structures. You need to know how to work them out in
principle, but your examiners are much more likely to ask you
for something simple like sulphur or iron.

Where would you like to go now?


To working out electronic structures for ions . . .
To the atomic properties menu . . .
To the atomic structure and bonding menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/atoms/properties/elstructs.html (9 of 9)30/12/2004 11:00:22

Understanding Chemistry - A' level syllabuses

Understanding Chemistry

UK A and AS LEVEL CHEMISTRY SYLLABUSES


I assume that you want to get the best grade you possibly can with the
minimum of effort! Getting a good A level grade is rather like playing a
game with your examiners - in which they make up the rules (and
occasionally change them). You aren't going to win unless you know
those rules.
Before you do anything else:

Get a copy of your syllabus if you haven't already got one. Details
of how to do this are given below.
Syllabuses are often quite difficult to interpret, so you need to
know exactly what questions your examiners are asking, and how
they are marking them.
Explore your Exam Board web site. They all offer free downloads
of specimen papers (including mark schemes), but you might have
to pay for recent exam papers and mark schemes, and other
support material. If they don't offer these free, find out how to order
them and invest a small amount of money in your future!
If you want the best possible grade, you should be working with
exam papers all the way through your course. Leaving looking at
exam papers until your last minute revision is too late.
Be careful, though! Syllabuses change and so do examiners.
Make sure that the question papers and mark schemes you get
relate to your current syllabus and are as recent as possible. A
new chief examiner can make a lot of difference to the style of a
question paper.

How to download a copy of your syllabus

http://www.chemguide.co.uk/syllabuses.html (1 of 3)30/12/2004 11:00:23

Understanding Chemistry - A' level syllabuses

Finding your way to the right syllabus


The following links take you to the front pages of each of the Exam Board
web sites and you will then have to find your own way to your syllabus.
This is because these sites are liable to change.
Be aware that the syllabuses are known as specifications. You want
GCE Advanced and Advanced Subsidiary (A and AS) Chemistry.
Finding the syllabuses is very straightforward - finding other information
may take you longer!
The Exam Boards:

OCR
This includes both the standard OCR syllabus and the Salters
syllabus.

Edexcel
This includes both the standard Edexcel syllabus and the Nuffield
syllabus.

AQA
AQA have free downloadable versions of all their recent exam
papers and mark schemes. Once you get to the chemistry page,
look for it under "Assessment Material". You can also get
Examiners' Reports (another link from the chemistry page). These
are essential if you want to avoid common mistakes.

WJEC
This link should take you directly to the correct chemistry page to
download a syllabus. At the time of writing, you will have to pay if
you want past papers or mark schemes.

Problems reading the downloaded syllabus?


The syllabuses are available only in pdf format. You need software such
as Adobe Acrobat Reader to access it. You have almost certainly got this
http://www.chemguide.co.uk/syllabuses.html (2 of 3)30/12/2004 11:00:23

Understanding Chemistry - A' level syllabuses

(or the equivalent) software on your computer, but if your computer is


old, you may not have the latest version.
If your downloaded syllabus won't open, it may be that the syllabus was
created in a newer version of the Reader than you've got. You will have
to download a new version of Reader.
Each of the Exam Board web sites provides a link to Adobe, but these
links are often easy to miss.
Use this link:

www.adobe.com
This will take you to Adobe's front page where you will find a link
enabling you to download the Reader. Be warned that this is a
seriously large bit of software and could take a long time to
download on a dial-up connection.

Go to Main Menu . . .

Jim Clark 2000 (modified 2004)

http://www.chemguide.co.uk/syllabuses.html (3 of 3)30/12/2004 11:00:23

electronic structures of ions

ELECTRONIC STRUCTURES OF IONS

This page explores how you write electronic structures for simple
monatomic ions (ions containing only one atom) using s, p, and d
notation. It assumes that you already understand how to write electronic
structures for atoms.

Important! If you have come straight to this page via a


search engine, you should read the page on electronic
structures of atoms before you go any further.

Working out the electronic structures of ions


Ions are atoms (or groups of atoms) which carry an electric charge
because they have either gained or lost one or more electrons. If an
atom gains electrons it acquires a negative charge. If it loses electrons, it
becomes positively charged.
The electronic structure of s- and p-block ions
Write the electronic structure for the neutral atom, and then add (for a
negative ion) or subtract electrons (for a positive ion).
To write the electronic structure for Cl -:
Cl

1s22s22p63s23px23py23pz1

Cl-

1s22s22p63s23px23py23pz2

but Cl- has one more electron

To write the electronic structure for O2-:


O

1s22s22px22py12pz1

but O2- has two more electrons

http://www.chemguide.co.uk/atoms/properties/ionstruct.html (1 of 4)30/12/2004 11:00:26

electronic structures of ions

O2-

1s22s22px22py22pz2

To write the electronic structure for Na+:


Na

1s22s22p63s1

Na+

1s22s22p6

but Na+ has one less electron

To write the electronic structure for Ca2+:


Ca

1s22s22p63s23p64s2

Ca2+

1s22s22p63s23p6

but Ca2+ has two less electrons

The electronic structure of d-block ions


Here you are faced with one of the most irritating facts in A'level
chemistry! You will recall that the first transition series (from scandium to
zinc) is the result of the 3d orbitals being filled after the 4s orbital.
However, once the electrons are established in their orbitals, the energy
order changes - and in all the chemistry of the transition elements, the 4s
orbital behaves as the outermost, highest energy orbital. The reversed
order of the 3d and 4s orbitals only applies to building the atom up in the
first place. In all other respects, the 4s electrons are always the electrons
you need to think about first.
You must remember this:

When d-block elements form ions, the 4s electrons are lost first.

Provided you remember that, working out the structure of a d-block ion is
no different from working out the structure of, say, a sodium ion.

http://www.chemguide.co.uk/atoms/properties/ionstruct.html (2 of 4)30/12/2004 11:00:26

electronic structures of ions

To write the electronic structure for Cr3+:


Cr

1s22s22p63s23p63d54s1

Cr3+

1s22s22p63s23p63d3

The 4s electron is lost first followed by two of the 3d electrons.

To write the electronic structure for Zn2+:


Zn

1s22s22p63s23p63d104s2

Zn2+

1s22s22p63s23p63d10

This time there is no need to use any of the 3d electrons.

To write the electronic structure for Fe3+:


Fe

1s22s22p63s23p63d64s2

Fe3+

1s22s22p63s23p63d5

The 4s electrons are lost first followed by one of the 3d electrons.


The rule is quite simple. Take the 4s electrons off first, and then as many
3d electrons as necessary to produce the correct positive charge.

http://www.chemguide.co.uk/atoms/properties/ionstruct.html (3 of 4)30/12/2004 11:00:26

electronic structures of ions

Note: You may well have the impression from GCSE that
ions have to have noble gas structures. It's not true! Most
(but not all) ions formed by s- and p-block elements do have
noble gas structures, but if you look at the d-block ions we've
used as examples, not one of them has a noble gas structure
- yet they are all perfectly valid ions. Getting away from a
reliance on the concept of noble gas structures is one of the
difficult mental leaps that you have to make at the beginning
of A'level chemistry.

Where would you like to go now?


To the atomic properties menu . . .
To the atomic structure and bonding menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/atoms/properties/ionstruct.html (4 of 4)30/12/2004 11:00:26

first ionisation energy

IONISATION ENERGY

This page explains what first ionisation energy is, and then looks at the
way it varies around the Periodic Table - across periods and down
groups. It assumes that you know about simple atomic orbitals, and can
write electronic structures for simple atoms. You will find a link at the
bottom of the page to a similar description of successive ionisation
energies (second, third and so on).

Important! If you aren't reasonable happy about atomic


orbitals and electronic structures you should follow these
links before you go any further.

Defining first ionisation energy


Definition
The first ionisation energy is the energy required to remove the most
loosely held electron from one mole of gaseous atoms to produce 1 mole
of gaseous ions each with a charge of 1+.
This is more easily seen in symbol terms.

It is the energy needed to carry out this change per mole of X.

Worried about moles? Don't be! For now, just take it as a


measure of a particular amount of a substance. It isn't worth
worrying about at the moment.

http://www.chemguide.co.uk/atoms/properties/ies.html (1 of 12)30/12/2004 11:00:33

first ionisation energy

Things to notice about the equation


The state symbols - (g) - are essential. When you are talking about
ionisation energies, everything must be present in the gas state.
Ionisation energies are measured in kJ mol-1 (kilojoules per mole). They
vary in size from 381 (which you would consider very low) up to 2370
(which is very high).
All elements have a first ionisation energy - even atoms which don't form
positive ions in test tubes. The reason that helium (1st I.E. = 2370 kJ mol1) doesn't normally form a positive ion is because of the huge amount of
energy that would be needed to remove one of its electrons.

Patterns of first ionisation energies in the Periodic Table


The first 20 elements

First ionisation energy shows periodicity. That means that it varies in a


repetitive way as you move through the Periodic Table. For example,
look at the pattern from Li to Ne, and then compare it with the identical
pattern from Na to Ar.
These variations in first ionisation energy can all be explained in terms of
the structures of the atoms involved.

Factors affecting the size of ionisation energy


http://www.chemguide.co.uk/atoms/properties/ies.html (2 of 12)30/12/2004 11:00:33

first ionisation energy

Ionisation energy is a measure of the energy needed to pull a particular


electron away from the attraction of the nucleus. A high value of
ionisation energy shows a high attraction between the electron and the
nucleus.
The size of that attraction will be governed by:
The charge on the nucleus.
The more protons there are in the nucleus, the more positively charged
the nucleus is, and the more strongly electrons are attracted to it.
The distance of the electron from the nucleus.
Attraction falls off very rapidly with distance. An electron close to the
nucleus will be much more strongly attracted than one further away.
The number of electrons between the outer electrons and the
nucleus.
Consider a sodium atom, with the electronic structure 2,8,1. (There's no
reason why you can't use this notation if it's useful!)
If the outer electron looks in towards the nucleus, it doesn't see the
nucleus sharply. Between it and the nucleus there are the two layers of
electrons in the first and second levels. The 11 protons in the sodium's
nucleus have their effect cut down by the 10 inner electrons. The outer
electron therefore only feels a net pull of approximately 1+ from the
centre. This lessening of the pull of the nucleus by inner electrons is
known as screening or shielding.

Warning! Electrons don't, of course, "look in" towards the


nucleus - and they don't "see" anything either! But there's no
reason why you can't imagine it in these terms if it helps you
to visualise what's happening. Just don't use these terms in
an exam! You may get an examiner who is upset by this sort
of loose language.

http://www.chemguide.co.uk/atoms/properties/ies.html (3 of 12)30/12/2004 11:00:33

first ionisation energy

Whether the electron is on its own in an orbital or paired with


another electron.
Two electrons in the same orbital experience a bit of repulsion from each
other. This offsets the attraction of the nucleus, so that paired electrons
are removed rather more easily than you might expect.

Explaining the pattern in the first few elements


Hydrogen has an electronic structure of 1s1. It is a very small atom, and
the single electron is close to the nucleus and therefore strongly
attracted. There are no electrons screening it from the nucleus and so
the ionisation energy is high (1310 kJ mol-1).
Helium has a structure 1s2. The electron is being removed from the
same orbital as in hydrogen's case. It is close to the nucleus and
unscreened. The value of the ionisation energy (2370 kJ mol-1) is much
higher than hydrogen, because the nucleus now has 2 protons attracting
the electrons instead of 1.
Lithium is 1s22s1. Its outer electron is in the second energy level, much
more distant from the nucleus. You might argue that that would be offset
by the additional proton in the nucleus, but the electron doesn't feel the
full pull of the nucleus - it is screened by the 1s2 electrons.

http://www.chemguide.co.uk/atoms/properties/ies.html (4 of 12)30/12/2004 11:00:33

first ionisation energy

You can think of the electron as feeling a net 1+ pull from the centre (3
protons offset by the two 1s2 electrons).
If you compare lithium with hydrogen (instead of with helium), the
hydrogen's electron also feels a 1+ pull from the nucleus, but the
distance is much greater with lithium. Lithium's first ionisation energy
drops to 519 kJ mol-1 whereas hydrogen's is 1310 kJ mol-1.

The patterns in periods 2 and 3


Talking through the next 17 atoms one at a time would take ages. We
can do it much more neatly by explaining the main trends in these
periods, and then accounting for the exceptions to these trends.
The first thing to realise is that the patterns in the two periods are
identical - the difference being that the ionisation energies in period 3 are
all lower than those in period 2.

http://www.chemguide.co.uk/atoms/properties/ies.html (5 of 12)30/12/2004 11:00:33

first ionisation energy

Explaining the general trend across periods 2 and 3


The general trend is for ionisation energies to increase across a period.
In the whole of period 2, the outer electrons are in 2-level orbitals - 2s or
2p. These are all the same sort of distances from the nucleus, and are
screened by the same 1s2 electrons.
The major difference is the increasing number of protons in the nucleus
as you go from lithium to neon. That causes greater attraction between
the nucleus and the electrons and so increases the ionisation energies.
In fact the increasing nuclear charge also drags the outer electrons in
closer to the nucleus. That increases ionisation energies still more as you
go across the period.

Note: Factors affecting atomic radius are covered on a


separate page.

http://www.chemguide.co.uk/atoms/properties/ies.html (6 of 12)30/12/2004 11:00:33

first ionisation energy

In period 3, the trend is exactly the same. This time, all the electrons
being removed are in the third level and are screened by the 1s22s22p6
electrons. They all have the same sort of environment, but there is an
increasing nuclear charge.
Why the drop between groups 2 and 3 (Be-B and Mg-Al)?
The explanation lies with the structures of boron and aluminium. The
outer electron is removed more easily from these atoms than the general
trend in their period would suggest.
Be

1s22s2

1st I.E. = 900 kJ mol-1

1s22s22px1

1st I.E. = 799 kJ mol-1

You might expect the boron value to be more than the beryllium value
because of the extra proton. Offsetting that is the fact that boron's outer
electron is in a 2p orbital rather than a 2s. 2p orbitals have a slightly
higher energy than the 2s orbital, and the electron is, on average, to be
found further from the nucleus. This has two effects.

The increased distance results in a reduced attraction and so a


reduced ionisation energy.
The 2p orbital is screened not only by the 1s2 electrons but, to
some extent, by the 2s2 electrons as well. That also reduces the
pull from the nucleus and so lowers the ionisation energy.

The explanation for the drop between magnesium and aluminium is the
same, except that everything is happening at the 3-level rather than the 2level.
Mg

1s22s22p63s2

1st I.E. = 736 kJ mol-1

Al

1s22s22p63s23px1

1st I.E. = 577 kJ mol-1

The 3p electron in aluminium is slightly more distant from the nucleus


than the 3s, and partially screened by the 3s2 electrons as well as the
inner electrons. Both of these factors offset the effect of the extra proton.

http://www.chemguide.co.uk/atoms/properties/ies.html (7 of 12)30/12/2004 11:00:33

first ionisation energy

Warning! You might possibly come across a text book which


describes the drop between group 2 and group 3 by saying
that a full s2 orbital is in some way especially stable and that
makes the electron more difficult to remove. In other words,
that the fluctuation is because the group 2 value for ionisation
energy is abnormally high. This is quite simply wrong! The
reason for the fluctuation is because the group 3 value is
lower than you might expect for the reasons we've looked at.

Why the drop between groups 5 and 6 (N-O and P-S)?


Once again, you might expect the ionisation energy of the group 6
element to be higher than that of group 5 because of the extra proton.
What is offsetting it this time?
N

1s22s22px12py12pz1

1st I.E. = 1400 kJ mol-1

1s22s22px22py12pz1

1st I.E. = 1310 kJ mol-1

The screening is identical (from the 1s2 and, to some extent, from the 2s2
electrons), and the electron is being removed from an identical orbital.
The difference is that in the oxygen case the electron being removed is
one of the 2px2 pair. The repulsion between the two electrons in the
same orbital means that the electron is easier to remove than it would
otherwise be.
The drop in ionisation energy at sulphur is accounted for in the same way.

Trends in ionisation energy down a group


As you go down a group in the Periodic Table ionisation energies
generally fall. You have already seen evidence of this in the fact that the
ionisation energies in period 3 are all less than those in period 2.
Taking Group 1 as a typical example:

http://www.chemguide.co.uk/atoms/properties/ies.html (8 of 12)30/12/2004 11:00:33

first ionisation energy

Why is the sodium value less than that of lithium?


There are 11 protons in a sodium atom but only 3 in a lithium atom, so
the nuclear charge is much greater. You might have expected a much
larger ionisation energy in sodium, but offsetting the nuclear charge is a
greater distance from the nucleus and more screening.
Li

1s22s1

1st I.E. = 519 kJ mol-1

Na

1s22s22p63s1

1st I.E. = 494 kJ mol-1

Lithium's outer electron is in the second level, and only has the 1s2
electrons to screen it. The 2s1 electron feels the pull of 3 protons
screened by 2 electrons - a net pull from the centre of 1+.
The sodium's outer electron is in the third level, and is screened from the
11 protons in the nucleus by a total of 10 inner electrons. The 3s1
electron also feels a net pull of 1+ from the centre of the atom. In other
words, the effect of the extra protons is compensated for by the effect of
the extra screening electrons. The only factor left is the extra distance
between the outer electron and the nucleus in sodium's case. That
lowers the ionisation energy.
Similar explanations hold as you go down the rest of this group - or,
indeed, any other group.

Trends in ionisation energy in a transition series

http://www.chemguide.co.uk/atoms/properties/ies.html (9 of 12)30/12/2004 11:00:33

first ionisation energy

Apart from zinc at the end, the other ionisation energies are all much the
same.
All of these elements have an electronic structure [Ar]3dn4s2 (or 4s1 in
the cases of chromium and copper). The electron being lost always
comes from the 4s orbital.

Note: Confusingly, once the orbitals have electrons in them,


the 4s orbital has a higher energy than the 3d - quite the
opposite of their order when the atoms are being filled with
electrons. That means that it is a 4s electron which is lost
from the atom when it forms an ion. It also means that the 3d
orbitals are slightly closer to the nucleus than the 4s - and so
offer some screening.
You will find this commented on in the page about electronic
structures of ions.

As you go from one atom to the next in the series, the number of protons
in the nucleus increases, but so also does the number of 3d electrons.
The 3d electrons have some screening effect, and the extra proton and
the extra 3d electron more or less cancel each other out as far as
attraction from the centre of the atom is concerned.
The rise at zinc is easy to explain.
Cu

[Ar]3d104s1

1st I.E. = 745 kJ mol-1

Zn

[Ar]3d104s2

1st I.E. = 908 kJ mol-1

In each case, the electron is coming from the same orbital, with identical
http://www.chemguide.co.uk/atoms/properties/ies.html (10 of 12)30/12/2004 11:00:33

first ionisation energy

screening, but the zinc has one extra proton in the nucleus and so the
attraction is greater.

Ionisation energies and reactivity


The lower the ionisation energy, the more easily this change happens:

You can explain the increase in reactivity of the Group 1 metals (Li, Na,
K, Rb, Cs) as you go down the group in terms of the fall in ionisation
energy. Whatever these metals react with, they have to form positive
ions in the process, and so the lower the ionisation energy, the more
easily those ions will form.
The danger with this approach is that the formation of the positive ion is
only one stage in a multi-step process.
For example, you wouldn't be starting with gaseous atoms; nor would
you end up with gaseous positive ions - you would end up with ions in a
solid or in solution. The energy changes in these processes also vary
from element to element. Ideally you need to consider the whole picture
and not just one small part of it.
However, the ionisation energies of the elements are going to be major
contributing factors towards the activation energy of the reactions.
Remember that activation energy is the minimum energy needed before
a reaction will take place. The lower the activation energy, the faster the
reaction will be - irrespective of what the overall energy changes in the
reaction are.
The fall in ionisation energy as you go down a group will lead to lower
activation energies and therefore faster reactions.

http://www.chemguide.co.uk/atoms/properties/ies.html (11 of 12)30/12/2004 11:00:33

first ionisation energy

Note: You will find a page discussing this in more detail in


the inorganic section of this site dealing with the reactions of
Group 2 metals with water.

Where would you like to go now?


To look at second (and successive) ionisation energies . . .
To the atomic properties menu . . .
To the atomic structure and bonding menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/atoms/properties/ies.html (12 of 12)30/12/2004 11:00:33

atomic and ionic radius

ATOMIC AND IONIC RADIUS

This page explains the various measures of atomic radius, and then looks
at the way it varies around the Periodic Table - across periods and down
groups. It assumes that you understand electronic structures for simple
atoms written in s, p, d notation.

Important! If you aren't reasonable happy about electronic


structures you should follow this link before you go any further.

ATOMIC RADIUS
Measures of atomic radius
Unlike a ball, an atom doesn't have a fixed radius. The radius of an atom
can only be found by measuring the distance between the nuclei of two
touching atoms, and then halving that distance.

As you can see from the diagrams, the same atom could be found to have
a different radius depending on what was around it.
The left hand diagram shows bonded atoms. The atoms are pulled closely
together and so the measured radius is less than if they are just touching.
This is what you would get if you had metal atoms in a metallic structure,
or atoms covalently bonded to each other. The type of atomic radius being
measured here is called the metallic radius or the covalent radius
depending on the bonding.

http://www.chemguide.co.uk/atoms/properties/atradius.html (1 of 6)30/12/2004 11:00:37

atomic and ionic radius

The right hand diagram shows what happens if the atoms are just
touching. The attractive forces are much less, and the atoms are
essentially "unsquashed". This measure of atomic radius is called the van
der Waals radius after the weak attractions present in this situation.

Note: If you want to explore these various types of bonding


this link will take you to the bonding menu.

Trends in atomic radius in the Periodic Table


The exact pattern you get depends on which measure of atomic radius
you use - but the trends are still valid.
The following diagram uses metallic radii for metallic elements, covalent
radii for elements that form covalent bonds, and van der Waals radii for
those (like the noble gases) which don't form bonds.
Trends in atomic radius in Periods 2 and 3

Trends in atomic radius down a group


It is fairly obvious that the atoms get bigger as you go down groups. The
reason is equally obvious - you are adding extra layers of electrons.
Trends in atomic radius across periods
You have to ignore the noble gas at the end of each period. Because
neon and argon don't form bonds, you can only measure their van der
Waals radius - a case where the atom is pretty well "unsquashed". All the
other atoms are being measured where their atomic radius is being
lessened by strong attractions. You aren't comparing like with like if you

http://www.chemguide.co.uk/atoms/properties/atradius.html (2 of 6)30/12/2004 11:00:37

atomic and ionic radius

include the noble gases.

Leaving the noble gases out, atoms get smaller


as you go across a period.

If you think about it, the metallic or covalent radius is going to be a


measure of the distance from the nucleus to the electrons which make up
the bond. (Look back to the left-hand side of the first diagram on this page
if you aren't sure, and picture the bonding electrons as being half way
between the two nuclei.)
From lithium to fluorine, those electrons are all in the 2-level, being
screened by the 1s2 electrons. The increasing number of protons in the
nucleus as you go across the period pulls the electrons in more tightly.
The amount of screening is constant for all of these elements.

Note: You might possibly wonder why you don't get extra
screening from the 2s2 electrons in the cases of the elements
from boron to fluorine where the bonding involves the p
electrons.
In each of these cases, before bonding happens, the existing s
and p orbitals are reorganised (hybridised) into new orbitals of
equal energy. When these atoms are bonded, there aren't any
2s electrons as such.
If you don't know about hybridisation, just ignore this comment
- you won't need it for UK A level purposes anyway.

http://www.chemguide.co.uk/atoms/properties/atradius.html (3 of 6)30/12/2004 11:00:37

atomic and ionic radius

In the period from sodium to chlorine, the same thing happens. The size of
the atom is controlled by the 3-level bonding electrons being pulled closer
to the nucleus by increasing numbers of protons - in each case, screened
by the 1- and 2-level electrons.

Trends in the transition elements

Although there is a slight contraction at the beginning of the series, the


atoms are all much the same size.
The size is determined by the 4s electrons. The pull of the increasing
number of protons in the nucleus is more or less offset by the extra
screening due to the increasing number of 3d electrons.

Note: Confusingly, once the orbitals have electrons in them,


the 4s orbital has a higher energy than the 3d - quite the
opposite of their order when the atoms are being filled with
electrons. That means that it is the 4s electrons which can be
thought of as being on the outside of the atom, and so
determine its size. It also means that the 3d orbitals are
slightly closer to the nucleus than the 4s - and so offer some
screening.
You will find this commented on in the page about electronic
structures of ions.

http://www.chemguide.co.uk/atoms/properties/atradius.html (4 of 6)30/12/2004 11:00:37

atomic and ionic radius

IONIC RADIUS
Ions aren't the same size as the atoms they come from. Compare the
sizes of sodium and chloride ions with the sizes of sodium and chlorine
atoms.

Positive ions
Positive ions are smaller than the atoms they come from. Sodium is 2,8,1;
Na+ is 2,8. You've lost a whole layer of electrons, and the remaining 10
electrons are being pulled in by the full force of 11 protons.
Negative ions
Negative ions are bigger than the atoms they come from. Chlorine is
2,8,7; Cl- is 2,8,8. Although the electrons are still all in the 3-level, the
extra repulsion produced by the incoming electron causes the atom to
expand. There are still only 17 protons, but they are now having to hold 18
electrons.

Where would you like to go now?


To the atomic properties menu . . .
To the atomic structure and bonding menu . . .
To Main Menu . . .

http://www.chemguide.co.uk/atoms/properties/atradius.html (5 of 6)30/12/2004 11:00:37

atomic and ionic radius

Jim Clark 2000 (modified 2004)

http://www.chemguide.co.uk/atoms/properties/atradius.html (6 of 6)30/12/2004 11:00:37

Bonding menu

Understanding Chemistry

BONDING MENU

Ionic bonding . . .
Includes a simple view of ionic bonding and the way you need to
modify this for A'level purposes.
Covalent bonding . . .
Includes a simple view of covalent bonding (single and double)
and the modifications needed for A'level purposes.
Co-ordinate (dative covalent) bonding . . .
Explains what co-ordinate (dative covalent) bonding is, and looks
at a wide range of examples.
Electronegativity . . .
Explains what electronegativity is and how it varies around the
Periodic Table. Describes and explains how electronegativity
differences determine the type of bond formed. Looks at polar
bonds and molecules.
Shapes of simple molecules and ions . . .
Explains how to work out the shapes of a wide range of simple
molecules and ions.
Metallic bonding . . .
A simple explanation of the forces holding metals together.
http://www.chemguide.co.uk/atoms/bondingmenu.html (1 of 2)30/12/2004 11:00:38

Bonding menu

van der Waals forces . . .


A description of van der Waals forces (temporary fluctuating dipole
and dipole-dipole interactions) causing attractions between
individual molecules.
Hydrogen bonding . . .
An explanation of how hydrogen bonding arises and its effect on
boiling points.

Bonding in organic compounds . . .


This leads you to the bonding menu in the organic section of this
site in case you are only interested in bonding in organic
compounds.

Go to atomic structure and bonding menu . . .


Go to Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/atoms/bondingmenu.html (2 of 2)30/12/2004 11:00:38

ionic (electrovalent) bonding

IONIC (ELECTROVALENT) BONDING

This page explains what ionic (electrovalent) bonding is. It starts with a
simple picture of the formation of ions, and then modifies it slightly for
A'level purposes.

A simple view of ionic bonding


The importance of noble gas structures
At a simple level (like GCSE) a lot of importance is attached to the
electronic structures of noble gases like neon or argon which have eight
electrons in their outer energy levels (or two in the case of helium).
These noble gas structures are thought of as being in some way a
"desirable" thing for an atom to have.
You may well have been left with the strong impression that when other
atoms react, they try to organise things such that their outer levels are
either completely full or completely empty.

Note: The central role given to noble gas structures is very


much an over-simplification. We shall have to spend some
time later on demolishing the concept!

http://www.chemguide.co.uk/atoms/bonding/ionic.html (1 of 7)30/12/2004 11:00:42

ionic (electrovalent) bonding

Ionic bonding in sodium chloride


Sodium (2,8,1) has 1 electron more than a stable noble gas structure
(2,8). If it gave away that electron it would become more stable.
Chlorine (2,8,7) has 1 electron short of a stable noble gas structure
(2,8,8). If it could gain an electron from somewhere it too would become
more stable.
The answer is obvious. If a sodium atom gives an electron to a chlorine
atom, both become more stable.

The sodium has lost an electron, so it no longer has equal numbers of


electrons and protons. Because it has one more proton than electron, it
has a charge of 1+. If electrons are lost from an atom, positive ions are
formed.
Positive ions are sometimes called cations.
The chlorine has gained an electron, so it now has one more electron
than proton. It therefore has a charge of 1-. If electrons are gained by an
atom, negative ions are formed.
A negative ion is sometimes called an anion.
The nature of the bond
The sodium ions and chloride ions are held together by the strong
electrostatic attractions between the positive and negative charges.
The formula of sodium chloride
You need one sodium atom to provide the extra electron for one chlorine
atom, so they combine together 1:1. The formula is therefore NaCl.

http://www.chemguide.co.uk/atoms/bonding/ionic.html (2 of 7)30/12/2004 11:00:42

ionic (electrovalent) bonding

Some other examples of ionic bonding


magnesium oxide

Again, noble gas structures are formed, and the magnesium oxide is held
together by very strong attractions between the ions. The ionic bonding is
stronger than in sodium chloride because this time you have 2+ ions
attracting 2- ions. The greater the charge, the greater the attraction.
The formula of magnesium oxide is MgO.
calcium chloride

This time you need two chlorines to use up the two outer electrons in the
calcium. The formula of calcium chloride is therefore CaCl2.
potassium oxide

Again, noble gas structures are formed. It takes two potassiums to


supply the electrons the oxygen needs. The formula of potassium oxide
is K2O.
http://www.chemguide.co.uk/atoms/bonding/ionic.html (3 of 7)30/12/2004 11:00:42

ionic (electrovalent) bonding

THE A'LEVEL VIEW OF IONIC BONDING

Electrons are transferred from one atom to another resulting in the


formation of positive and negative ions.
The electrostatic attractions between the positive and negative
ions hold the compound together.

So what's new? At heart - nothing. What needs modifying is the view that
there is something magic about noble gas structures. There are far more
ions which don't have noble gas structures than there are which do.
Some common ions which don't have noble gas structures
You may have come across some of the following ions in a basic course
like GCSE. They are all perfectly stable , but not one of them has a noble
gas structure.
Fe3+

[Ar]3d5

Cu2+

[Ar]3d9

Zn2+

[Ar]3d10

Ag+

[Kr]4d10

Pb2+

[Xe]4f145d106s2

Noble gases (apart from helium) have an outer electronic structure


ns2np6.

Note: If you aren't happy about writing electronic structures


using of s, p and d notation, follow this link before you go on.
Return to this page via the menus or by using the BACK
button on your browser.

http://www.chemguide.co.uk/atoms/bonding/ionic.html (4 of 7)30/12/2004 11:00:42

ionic (electrovalent) bonding

Apart from some elements at the beginning of a transition series


(scandium forming Sc3+ with an argon structure, for example), all
transition elements and any metals following a transition series (like tin
and lead in Group 4, for example) will have structures like those above.
That means that the only elements to form positive ions with noble gas
structures (apart from odd ones like scandium) are those in groups 1 and
2 of the Periodic Table and aluminium in group 3 (boron in group 3
doesn't form ions).
Negative ions are tidier! Those elements in Groups 5, 6 and 7 which form
simple negative ions all have noble gas structures.
If elements aren't aiming for noble gas structures when they form ions,
what decides how many electrons are transferred? The answer lies in the
energetics of the process by which the compound is made.

Warning! From here to the bottom of this page goes beyond


anything you are likely to need for A'level purposes. It is
included for interest only.

What determines what the charge is on an ion?


Elements combine to make the compound which is as stable as possible
- the one in which the greatest amount of energy is evolved in its making.
The more charges a positive ion has, the greater the attraction towards
its accompanying negative ion. The greater the attraction, the more
energy is released when the ions come together.
That means that elements forming positive ions will tend to give away as
many electrons as possible. But there's a down-side to this.
Energy is needed to remove electrons from atoms. This is called
ionisation energy. The more electrons you remove, the greater the total
ionisation energy becomes. Eventually the total ionisation energy needed
becomes so great that the energy released when the attractions are set
up between positive and negative ions isn't large enough to cover it.

http://www.chemguide.co.uk/atoms/bonding/ionic.html (5 of 7)30/12/2004 11:00:42

ionic (electrovalent) bonding

The element forms the ion which makes the compound most stable - the
one in which most energy is released over-all.
For example, why is calcium chloride CaCl2 rather than CaCl or CaCl3?
If one mole of CaCl (containing Ca+ ions) is made from its elements, it is
possible to estimate that about 171 kJ of heat is evolved.
However, making CaCl2 (containing Ca2+ ions) releases more heat. You
get 795 kJ. That extra amount of heat evolved makes the compound
more stable, which is why you get CaCl2 rather than CaCl.
What about CaCl3 (containing Ca3+ ions)? To make one mole of this, you
can estimate that you would have to put in 1341 kJ. This makes this
compound completely non-viable. Why is so much heat needed to make
CaCl3? It is because the third ionisation energy (the energy needed to
remove the third electron) is extremely high (4940 kJ mol-1) because the
electron is being removed from the 3-level rather than the 4-level.
Because it is much closer to the nucleus than the first two electrons
removed, it is going to be held much more strongly.

Note: It would pay you to read about ionisation energies if


you really want to understand this.
You could also go to a standard text book and investigate
Born-Haber Cycles.

http://www.chemguide.co.uk/atoms/bonding/ionic.html (6 of 7)30/12/2004 11:00:42

ionic (electrovalent) bonding

A similar sort of argument applies to the negative ion. For example,


oxygen forms an O2- ion rather than an O- ion or an O3- ion, because
compounds containing the O2- ion turn out to be the most energetically
stable.

Where would you like to go now?


To explore the physical properties of ionic compounds . . .
To the bonding menu . . .
To the atomic structure and bonding menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/atoms/bonding/ionic.html (7 of 7)30/12/2004 11:00:42

ionic structures

IONIC STRUCTURES

This page explains the relationship between the arrangement of the ions
in a typical ionic solid like sodium chloride and its physical properties melting point, boiling point, brittleness, solubility and electrical behaviour.

Note: If you need to revise how ionic bonding arises, then


you might like to follow this link. It isn't important for
understanding this page, however.

The structure of a typical ionic solid - sodium chloride


How the ions are arranged in sodium chloride
Sodium chloride is taken as a typical ionic compound. Compounds like
this consist of a giant (endlessly repeating) lattice of ions. So sodium
chloride (and any other ionic compound) is described as having a giant
ionic structure.
You should be clear that giant in this context doesn't just mean very
large. It means that you can't state exactly how many ions there are.
There could be billions of sodium ions and chloride ions packed together,
or trillions, or whatever - it simply depends how big the crystal is. That is
different from, say, a water molecule which always contains exactly 2
hydrogen atoms and one oxygen atom - never more and never less.
A small representative bit of a sodium chloride lattice looks like this:

http://www.chemguide.co.uk/atoms/structures/ionicstruct.html (1 of 7)30/12/2004 11:00:48

ionic structures

If you look at the diagram carefully, you will see that the sodium ions and
chloride ions alternate with each other in each of the three dimensions.
This diagram is easy enough to draw with a computer, but extremely
difficult to draw convincingly by hand. We normally draw an "exploded"
version which looks like this:

Only those ions joined by lines are actually touching each other. The
sodium ion in the centre is being touched by 6 chloride ions. By chance
we might just as well have centred the diagram around a chloride ion that, of course, would be touched by 6 sodium ions. Sodium chloride is
described as being 6:6-co-ordinated.
You must remember that this diagram represents only a tiny part of the
whole sodium chloride crystal. The pattern repeats in this way over
countless ions.

How to draw this structure


Draw a perfect square:

http://www.chemguide.co.uk/atoms/structures/ionicstruct.html (2 of 7)30/12/2004 11:00:48

ionic structures

Now draw an identical square behind this one and offset a bit. You might
have to practice a bit to get the placement of the two squares right. If you
get it wrong, the ions get all tangled up with each other in your final
diagram.

Turn this into a perfect cube by joining the squares together:

Now the tricky bit! Subdivide this big cube into 8 small cubes by joining
the mid point of each edge to the mid point of the edge opposite it. To
complete the process you will also have to join the mid point of each face
(easily found once you've joined the edges) to the mid point of the
opposite face.

Now all you have to do is put the ions in. Use different colours or different
sizes for the two different ions, and don't forget a key. It doesn't matter
whether you end up with a sodium ion or a chloride ion in the centre of
http://www.chemguide.co.uk/atoms/structures/ionicstruct.html (3 of 7)30/12/2004 11:00:48

ionic structures

the cube - all that matters is that they alternate in all three dimensions.

You should be able to draw a perfectly adequate free-hand sketch of this


in under two minutes - less than one minute if you're not too fussy!

Why is sodium chloride 6:6-co-ordinated?


The more attraction there is between the positive and negative ions, the
more energy is released. The more energy that is released, the more
energetically stable the structure becomes.
That means that to gain maximum stability, you need the maximum
number of attractions. So why does each ion surround itself with 6 ions of
the opposite charge?
That represents the maximum number of chloride ions that you can fit
around a central sodium ion before the chloride ions start touching each
other. If they start touching, you introduce repulsions into the crystal
which makes it less stable.

Note: If the positive ion is big enough, you can fit 8 chloride
ions around it. For example, caesium ions are significantly
bigger than sodium ions, and so caesium chloride is 8:8-coordinated. It can gain stability from the extra attractions
without any problems because of repulsion due to ions with
the same charge touching each other.
The structure of caesium chloride isn't on any current A'level
syllabuses.

http://www.chemguide.co.uk/atoms/structures/ionicstruct.html (4 of 7)30/12/2004 11:00:48

ionic structures

The physical properties of sodium chloride


Sodium chloride has a high melting and boiling point
There are strong electrostatic attractions between the positive and
negative ions, and it takes a lot of heat energy to overcome them. Ionic
substances all have high melting and boiling points. Differences between
ionic substances will depend on things like:

The number of charges on the ions


Magnesium oxide has exactly the same structure as sodium
chloride, but a much higher melting and boiling point. The 2+ and
2- ions attract each other more strongly than 1+ attracts 1-.

The sizes of the ions


If the ions are smaller they get closer together and so the
electrostatic attractions are greater. Rubidium iodide, for example,
melts and boils at slightly lower temperatures than sodium
chloride, because both rubidium and iodide ions are bigger than
sodium and chloride ions. The attractions are less between the
bigger ions and so less heat energy is needed to separate them.

Sodium chloride crystals are brittle


Brittleness is again typical of ionic substances. Imagine what happens to
the crystal if a stress is applied which shifts the ion layers slightly.

Ions of the same charge are brought side-by-side and so the crystal
http://www.chemguide.co.uk/atoms/structures/ionicstruct.html (5 of 7)30/12/2004 11:00:48

ionic structures

repels itself to pieces!

Sodium chloride is soluble in water


Many ionic solids are soluble in water - although not all. It depends on
whether there are big enough attractions between the water molecules
and the ions to overcome the attractions between the ions themselves.
Positive ions are attracted to the lone pairs on water molecules and coordinate (dative covalent) bonds may form. Water molecules form
hydrogen bonds with negative ions.

Note: The bonding in hydrated metal ions is covered in the


page on co-ordinate bonding. The bonding between negative
ions like chloride ions and water molecules is covered in the
page on hydrogen bonding.

Sodium chloride is insoluble in organic solvents


This is also typical of ionic solids. The attractions between the solvent
molecules and the ions aren't big enough to overcome the attractions
holding the crystal together.

The electrical behaviour of sodium chloride


Solid sodium chloride doesn't conduct electricity, because there are no
electrons which are free to move. Molten sodium chloride undergoes
electrolysis, which involves conduction of electricity because of the
movement of the ions. In the process, sodium and chlorine are produced.
This is a chemical change rather than a physical process.

http://www.chemguide.co.uk/atoms/structures/ionicstruct.html (6 of 7)30/12/2004 11:00:48

ionic structures

Where would you like to go now?


To the structures menu . . .
To the atomic structure and bonding menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/atoms/structures/ionicstruct.html (7 of 7)30/12/2004 11:00:48

co-ordinate (dative covalent) bonding

CO-ORDINATE (DATIVE COVALENT) BONDING

This page explains what co-ordinate (also called dative covalent)


bonding is. You need to have a reasonable understanding of simple
covalent bonding before you start.

Important! If you are uncertain about covalent bonding


follow this link before you go on with this page.

Co-ordinate (dative covalent) bonding


A covalent bond is formed by two atoms sharing a pair of electrons. The
atoms are held together because the electron pair is attracted by both of
the nuclei.
In the formation of a simple covalent bond, each atom supplies one
electron to the bond - but that doesn't have to be the case. A co-ordinate
bond (also called a dative covalent bond) is a covalent bond (a shared
pair of electrons) in which both electrons come from the same atom.
For the rest of this page, we shall use the term co-ordinate bond - but if
you prefer to call it a dative covalent bond, that's not a problem!

The reaction between ammonia and hydrogen chloride


If these colourless gases are allowed to mix, a thick white smoke of solid
ammonium chloride is formed.

Ammonium ions, NH4+, are formed by the transfer of a hydrogen ion from
the hydrogen chloride to the lone pair of electrons on the ammonia
molecule.
http://www.chemguide.co.uk/atoms/bonding/dative.html (1 of 9)30/12/2004 11:00:59

co-ordinate (dative covalent) bonding

When the ammonium ion, NH4+, is formed, the fourth hydrogen is


attached by a dative covalent bond, because only the hydrogen's nucleus
is transferred from the chlorine to the nitrogen. The hydrogen's electron
is left behind on the chlorine to form a negative chloride ion.
Once the ammonium ion has been formed it is impossible to tell any
difference between the dative covalent and the ordinary covalent bonds.
Although the electrons are shown differently in the diagram, there is no
difference between them in reality.
Representing co-ordinate bonds
In simple diagrams, a co-ordinate bond is shown by an arrow. The arrow
points from the atom donating the lone pair to the atom accepting it.

Dissolving hydrogen chloride in water to make hydrochloric acid


Something similar happens. A hydrogen ion (H+) is transferred from the
chlorine to one of the lone pairs on the oxygen atom.
http://www.chemguide.co.uk/atoms/bonding/dative.html (2 of 9)30/12/2004 11:00:59

co-ordinate (dative covalent) bonding

The H3O+ ion is variously called the hydroxonium ion, the hydronium ion
or the oxonium ion.
In an introductory chemistry course (such as GCSE), whenever you have
talked about hydrogen ions (for example in acids), you have actually
been talking about the hydroxonium ion. A raw hydrogen ion is simply a
proton, and is far too reactive to exist on its own in a test tube.
If you write the hydrogen ion as H+(aq), the "(aq)" represents the water
molecule that the hydrogen ion is attached to. When it reacts with
something (an alkali, for example), the hydrogen ion simply becomes
detached from the water molecule again.
Note that once the co-ordinate bond has been set up, all the hydrogens
attached to the oxygen are exactly equivalent. When a hydrogen ion
breaks away again, it could be any of the three.

The reaction between ammonia and boron trifluoride, BF3


If you have recently read the page on covalent bonding, you may
remember boron trifluoride as a compound which doesn't have a noble
gas structure around the boron atom. The boron only has 3 pairs of
electrons in its bonding level, whereas there would be room for 4 pairs.
BF3 is described as being electron deficient.
The lone pair on the nitrogen of an ammonia molecule can be used to
overcome that deficiency, and a compound is formed involving a coordinate bond.
http://www.chemguide.co.uk/atoms/bonding/dative.html (3 of 9)30/12/2004 11:00:59

co-ordinate (dative covalent) bonding

Using lines to represent the bonds, this could be drawn more simply as:

The second diagram shows another way that you might find co-ordinate
bonds drawn. The nitrogen end of the bond has become positive
because the electron pair has moved away from the nitrogen towards the
boron - which has therefore become negative. We shan't use this method
again - it's more confusing than just using an arrow.

The structure of aluminium chloride


Aluminium chloride sublimes (turns straight from a
solid to a gas) at 178C. If it contained ions it would
have a very high melting and boiling point because of
the strong attractions between the positive and
negative ions. The implication is that it must be
covalent. The dots-and-crosses diagram shows only
the outer electrons.
AlCl3, like BF3, is electron deficient. There is likely to be a similarity,
because aluminium and boron are in the same group of the Periodic
Table, as are fluorine and chlorine.

http://www.chemguide.co.uk/atoms/bonding/dative.html (4 of 9)30/12/2004 11:00:59

co-ordinate (dative covalent) bonding

Measurements of the relative formula mass of aluminium chloride show


that its formula in the solid is not AlCl3, but Al2Cl6. It exists as a dimer
(two molecules joined together). The bonding between the two molecules
is co-ordinate, using lone pairs on the chlorine atoms. Each chlorine
atom has 3 lone pairs, but only the two important ones are shown in the
line diagram.

Note: The uninteresting electrons on the chlorines have


been faded in colour to make the co-ordinate bonds show up
better. There's nothing special about those two particular lone
pairs - they just happen to be the ones pointing in the right
direction.

Energy is released when the two co-ordinate bonds are formed, and so
the dimer is more stable than two separate AlCl3 molecules.

The bonding in hydrated metal ions


Water molecules are strongly attracted to ions in solution - the water
molecules clustering around the positive or negative ions. In many cases,
the attractions are so great that formal bonds are made, and this is true
of almost all positive metal ions. Ions with water molecules attached are
described as hydrated ions.

http://www.chemguide.co.uk/atoms/bonding/dative.html (5 of 9)30/12/2004 11:00:59

co-ordinate (dative covalent) bonding

Although aluminium chloride is covalent, when it dissolves in water, ions


are produced. Six water molecules bond to the aluminium to give an ion
with the formula Al(H2O)63+. It's called the hexaaquaaluminium ion which translates as six ("hexa") water molecules ("aqua") wrapped
around an aluminium ion.
The bonding in this (and the similar ions formed by the
great majority of other metals) is co-ordinate (dative
covalent) using lone pairs on the water molecules.

Aluminium is 1s22s22p63s23px1. When it forms an Al3+ ion it loses the 3level electrons to leave 1s22s22p6.
That means that all the 3-level orbitals are now empty. The aluminium reorganises (hybridises) six of these (the 3s, three 3p, and two 3d) to
produce six new orbitals all with the same energy. These six hybrid
orbitals accept lone pairs from six water molecules.
You might wonder why it chooses to use six orbitals rather than four or
eight or whatever. Six is the maximum number of water molecules it is
possible to fit around an aluminium ion (and most other metal ions). By
making the maximum number of bonds, it releases most energy and so
becomes most energetically stable.

Only one lone pair is shown on each water molecule. The other lone pair
is pointing away from the aluminium and so isn't involved in the bonding.
The resulting ion looks like this:

http://www.chemguide.co.uk/atoms/bonding/dative.html (6 of 9)30/12/2004 11:00:59

co-ordinate (dative covalent) bonding

Because of the movement of electrons towards the centre of the ion, the
3+ charge is no longer located entirely on the aluminium, but is now
spread over the whole of the ion.

Note: Dotted arrows represent lone pairs coming from water


molecules behind the plane of the screen or paper. Wedge
shaped arrows represent bonds from water molecules in front
of the plane of the screen or paper.

Two more molecules

Note: Only one current A'level syllabus wants these two.


Check yours! If you haven't got a copy of your syllabus, follow
this link to find out how to get one.

http://www.chemguide.co.uk/atoms/bonding/dative.html (7 of 9)30/12/2004 11:00:59

co-ordinate (dative covalent) bonding

Carbon monoxide, CO
Carbon monoxide can be thought of as having two ordinary covalent
bonds between the carbon and the oxygen plus a co-ordinate bond using
a lone pair on the oxygen atom.

Nitric acid, HNO3


In this case, one of the oxygen atoms can be thought of as attaching to
the nitrogen via a co-ordinate bond using the lone pair on the nitrogen
atom.

In fact this structure is misleading because it suggests that the two


oxygen atoms on the right-hand side of the diagram are joined to the
nitrogen in different ways. Both bonds are actually identical in length and
strength, and so the arrangement of the electrons must be identical.
There is no way of showing this using a dots-and-crosses picture. The
bonding involves delocalisation.

http://www.chemguide.co.uk/atoms/bonding/dative.html (8 of 9)30/12/2004 11:00:59

co-ordinate (dative covalent) bonding

If you are interested: The bonding is rather similar to the


bonding in the ethanoate ion (although without the negative
charge). You will find thisdescribed on a page about the
acidity of organic acids.

Where would you like to go now?


To the bonding menu . . .
To the atomic structure and bonding menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/atoms/bonding/dative.html (9 of 9)30/12/2004 11:00:59

covalent bonding - single bonds

COVALENT BONDING - SINGLE BONDS

This page explains what covalent bonding is. It starts with a simple
picture of the single covalent bond, and then modifies it slightly for A'level
purposes. It also takes a more sophisticated view (beyond A'level) if you
are interested. You will find a link to a page on double covalent bonds at
the bottom of the page.

A simple view of covalent bonding


The importance of noble gas structures
At a simple level (like GCSE) a lot of importance is attached to the
electronic structures of noble gases like neon or argon which have eight
electrons in their outer energy levels (or two in the case of helium).
These noble gas structures are thought of as being in some way a
"desirable" thing for an atom to have.
You may well have been left with the strong impression that when other
atoms react, they try to achieve noble gas structures.
As well as achieving noble gas structures by transferring electrons from
one atom to another as in ionic bonding, it is also possible for atoms to
reach these stable structures by sharing electrons to give covalent bonds.
Some very simple covalent molecules
Chlorine
For example, two chlorine atoms could both achieve stable structures by
sharing their single unpaired electron as in the diagram.

http://www.chemguide.co.uk/atoms/bonding/covalent.html (1 of 13)30/12/2004 11:01:09

covalent bonding - single bonds

The fact that one chlorine has been drawn with electrons marked as
crosses and the other as dots is simply to show where all the electrons
come from. In reality there is no difference between them.
The two chlorine atoms are said to be joined by a covalent bond. The
reason that the two chlorine atoms stick together is that the shared pair
of electrons is attracted to the nucleus of both chlorine atoms.
Hydrogen

Hydrogen atoms only need two electrons in their outer level to reach the
noble gas structure of helium. Once again, the covalent bond holds the
two atoms together because the pair of electrons is attracted to both
nuclei.
Hydrogen chloride

The hydrogen has a helium structure, and the chlorine an argon structure.

Covalent bonding at A'level


Cases where there isn't any difference from the simple view
If you stick closely to modern A'level syllabuses, there is little need to
http://www.chemguide.co.uk/atoms/bonding/covalent.html (2 of 13)30/12/2004 11:01:09

covalent bonding - single bonds

move far from the simple (GCSE) view. The only thing which must be
changed is the over-reliance on the concept of noble gas structures.
Most of the simple molecules you draw do in fact have all their atoms
with noble gas structures.
For example:

Even with a more complicated molecule like PCl3, there's no problem. In


this case, only the outer electrons are shown for simplicity. Each atom in
this structure has inner layers of electrons of 2,8. Again, everything
present has a noble gas structure.

Cases where the simple view throws up problems


Boron trifluoride, BF3

http://www.chemguide.co.uk/atoms/bonding/covalent.html (3 of 13)30/12/2004 11:01:09

covalent bonding - single bonds

A boron atom only has 3 electrons in its outer level, and there is no
possibility of it reaching a noble gas structure by simple sharing of
electrons. Is this a problem? No. The boron has formed the maximum
number of bonds that it can in the circumstances, and this is a perfectly
valid structure.
Energy is released whenever a covalent bond is formed. Because energy
is being lost from the system, it becomes more stable after every
covalent bond is made. It follows, therefore, that an atom will tend to
make as many covalent bonds as possible. In the case of boron in BF3,
three bonds is the maximum possible because boron only has 3
electrons to share.

Note: You might perhaps wonder why boron doesn't form


ionic bonds with fluorine instead. Boron doesn't form ions
because the total energy needed to remove three electrons
to form a B3+ ion is simply too great to be recoverable when
attractions are set up between the boron and fluoride ions.

Phosphorus(V) chloride, PCl5


In the case of phosphorus 5 covalent bonds are possible - as in PCl5.
Phosphorus forms two chlorides - PCl3 and PCl5. When phosphorus
burns in chlorine both are formed - the majority product depending on
how much chlorine is available. We've already looked at the structure of
PCl3.
The diagram of PCl5 (like the previous diagram of PCl3) shows only the
outer electrons.

http://www.chemguide.co.uk/atoms/bonding/covalent.html (4 of 13)30/12/2004 11:01:09

covalent bonding - single bonds

Notice that the phosphorus now has 5 pairs of electrons in the outer level
- certainly not a noble gas structure. You would have been content to
draw PCl3 at GCSE, but PCl5 would have looked very worrying.
Why does phosphorus sometimes break away from a noble gas structure
and form five bonds? In order to answer that question, we need to
explore territory beyond the limits of current A'level syllabuses. Don't be
put off by this! It isn't particularly difficult, and is extremely useful if you
are going to understand the bonding in some important organic
compounds.

A more sophisticated view of covalent bonding


The bonding in methane, CH4

Warning! If you aren't happy with describing electron


arrangements in s and p notation, and with the shapes of s
and p orbitals, you need to read about orbitals before you go
on.
Use the BACK button on your browser to return quickly to this
point.

http://www.chemguide.co.uk/atoms/bonding/covalent.html (5 of 13)30/12/2004 11:01:09

covalent bonding - single bonds

What is wrong with the dots-and-crosses picture of bonding in


methane?
We are starting with methane because it is the simplest case which
illustrates the sort of processes involved. You will remember that the dotsand-crossed picture of methane looks like this.

There is a serious mis-match between this structure and the modern


electronic structure of carbon, 1s22s22px12py1. The modern structure
shows that there are only 2 unpaired electrons for hydrogens to share
with, instead of the 4 which the simple view requires.
You can see this more readily using the
electrons-in-boxes notation. Only the 2level electrons are shown. The 1s2
electrons are too deep inside the atom to
be involved in bonding. The only electrons
directly available for sharing are the 2p
electrons. Why then isn't methane CH2?
Promotion of an electron

http://www.chemguide.co.uk/atoms/bonding/covalent.html (6 of 13)30/12/2004 11:01:09

covalent bonding - single bonds

When bonds are formed, energy is


released and the system becomes more
stable. If carbon forms 4 bonds rather than
2, twice as much energy is released and so
the resulting molecule becomes even more
stable.
There is only a small energy gap between
the 2s and 2p orbitals, and so it pays the
carbon to provide a small amount of energy
to promote an electron from the 2s to the
empty 2p to give 4 unpaired electrons. The
extra energy released when the bonds form
more than compensates for the initial input.

Note: People sometimes worry that the promoted electron is


drawn as an up-arrow, whereas it started as a down-arrow. It
simply makes the diagram look tidier - nothing very
sophisticated is going on!

Now that we've got 4 unpaired electrons ready for bonding, another
problem arises. In methane all the carbon-hydrogen bonds are identical,
but our electrons are in two different kinds of orbitals. You aren't going to
get four identical bonds unless you start from four identical orbitals.
Hybridisation
The electrons rearrange themselves again
in a process called hybridisation. This
reorganises the electrons into four identical
hybrid orbitals called sp3 hybrids (because
they are made from one s orbital and three
p orbitals). You should read "sp3" as "s p
three" - not as "s p cubed".

http://www.chemguide.co.uk/atoms/bonding/covalent.html (7 of 13)30/12/2004 11:01:09

covalent bonding - single bonds

sp3 hybrid orbitals look a bit like half a p orbital,


and they arrange themselves in space so that
they are as far apart as possible. You can picture
the nucleus as being at the centre of a
tetrahedron (a triangularly based pyramid) with
the orbitals pointing to the corners. For clarity,
the nucleus is drawn far larger than it really is.

What happens when the bonds are formed?


Remember that hydrogen's electron is in a 1s orbital - a spherically
symmetric region of space surrounding the nucleus where there is some
fixed chance (say 95%) of finding the electron. When a covalent bond is
formed, the atomic orbitals (the orbitals in the individual atoms) merge to
produce a new molecular orbital which contains the electron pair which
creates the bond.

Four molecular orbitals are formed, looking rather like the original sp3
hybrids, but with a hydrogen nucleus embedded in each lobe. Each
orbital holds the 2 electrons that we've previously drawn as a dot and a
cross.
The principles involved - promotion of electrons if necessary, then
hybridisation, followed by the formation of molecular orbitals - can be
applied to any covalently-bound molecule.

http://www.chemguide.co.uk/atoms/bonding/covalent.html (8 of 13)30/12/2004 11:01:09

covalent bonding - single bonds

Note: You will find this bit on methane repeated in the


organic section of this site. That article on methane goes on
to look at the formation of carbon-carbon single bonds in
ethane.

The bonding in the phosphorus chlorides, PCl3 and PCl5


What's wrong with the simple view of PCl3?
This diagram only shows the outer (bonding) electrons.

Nothing is wrong with this! (Although it doesn't account for the shape of
the molecule properly.) If you were going to take a more modern look at
it, the argument would go like this:
Phosphorus has the electronic structure 1s22s22p63s23px13py13pz1. If
we look only at the outer electrons as "electrons-in-boxes":

There are 3 unpaired electrons that can be used to form bonds with 3
chlorine atoms. The four 3-level orbitals hybridise to produce 4
equivalent sp3 hybrids just like in carbon - except that one of these hybrid

http://www.chemguide.co.uk/atoms/bonding/covalent.html (9 of 13)30/12/2004 11:01:09

covalent bonding - single bonds

orbitals contains a lone pair of electrons.

Each of the 3 chlorines then forms a covalent bond by merging the


atomic orbital containing its unpaired electron with one of the phosphorus
unpaired electrons to make 3 molecular orbitals.
You might wonder whether all this is worth the bother! Probably not! It is
worth it with PCl5, though.
What's wrong with the simple view of PCl5?
You will remember that the dots-and-crosses picture of PCl5 looks
awkward because the phosphorus doesn't end up with a noble gas
structure. This diagram also shows only the outer electrons.

In this case, a more modern view makes things look better by


abandoning any pretence of worrying about noble gas structures.
If the phosphorus is going to form PCl5 it has first to generate 5 unpaired
electrons. It does this by promoting one of the electrons in the 3s orbital
to the next available higher energy orbital.
Which higher energy orbital? It uses one of the 3d orbitals. You might
http://www.chemguide.co.uk/atoms/bonding/covalent.html (10 of 13)30/12/2004 11:01:09

covalent bonding - single bonds

have expected it to use the 4s orbital because this is the orbital that fills
before the 3d when atoms are being built from scratch. Not so! Apart
from when you are building the atoms in the first place, the 3d always
counts as the lower energy orbital.

This leaves the phosphorus with this arrangement of its electrons:

The 3-level electrons now rearrange (hybridise) themselves to give 5


hybrid orbitals, all of equal energy. They would be called sp3d hybrids
because that's what they are made from.

The electrons in each of these orbitals would then share space with
electrons from five chlorines to make five new molecular orbitals - and
hence five covalent bonds.
Why does phosphorus form these extra two bonds? It puts in an amount
of energy to promote an electron, which is more than paid back when the
new bonds form. Put simply, it is energetically profitable for the
phosphorus to form the extra bonds.
The advantage of thinking of it in this way is that it completely ignores the

http://www.chemguide.co.uk/atoms/bonding/covalent.html (11 of 13)30/12/2004 11:01:09

covalent bonding - single bonds

question of whether you've got a noble gas structure, and so you don't
worry about it.

A non-existent compound - NCl5


Nitrogen is in the same Group of the Periodic Table as phosphorus, and
you might expect it to form a similar range of compounds. In fact, it
doesn't. For example, the compound NCl3 exists, but there is no such
thing as NCl5.
Nitrogen is 1s22s22px12py12pz1. The reason that NCl5 doesn't exist is
that in order to form five bonds, the nitrogen would have to promote one
of its 2s electrons. The problem is that there aren't any 2d orbitals to
promote an electron into - and the energy gap to the next level (the 3s) is
far too great.
In this case, then, the energy released when the extra bonds are made
isn't enough to compensate for the energy needed to promote an
electron - and so that promotion doesn't happen.
Atoms will form as many bonds as possible provided it is energetically
profitable.

Where would you like to go now?


To explore double covalent bonding . . .
To the bonding menu . . .
To the atomic structure and bonding menu . . .
To Main Menu . . .

http://www.chemguide.co.uk/atoms/bonding/covalent.html (12 of 13)30/12/2004 11:01:09

covalent bonding - single bonds

Jim Clark 2000

http://www.chemguide.co.uk/atoms/bonding/covalent.html (13 of 13)30/12/2004 11:01:09

bonding in methane - sp3 hybridisation

BONDING IN METHANE AND ETHANE

Warning! If you aren't happy with describing electron


arrangements in s and p notation, and with the shapes of s
and p orbitals, you really should read about orbitals.
Use the BACK button on your browser to return quickly to this
point.

Methane, CH4
The simple view of the bonding in methane
You will be familiar with drawing methane
using dots and crosses diagrams, but it is
worth looking at its structure a bit more
closely.
There is a serious mis-match between this
structure and the modern electronic structure
of carbon, 1s22s22px12py1. The modern
structure shows that there are only 2 unpaired electrons for hydrogens to
share with, instead of the 4 which the simple view requires.
You can see this more readily using the
electrons-in-boxes notation. Only the 2level electrons are shown. The 1s2
electrons are too deep inside the atom to
be involved in bonding. The only electrons
directly available for sharing are the 2p
electrons. Why then isn't methane CH2?
Promotion of an electron

http://www.chemguide.co.uk/basicorg/bonding/methane.html (1 of 6)30/12/2004 11:01:15

bonding in methane - sp3 hybridisation

When bonds are formed, energy is


released and the system becomes more
stable. If carbon forms 4 bonds rather than
2, twice as much energy is released and so
the resulting molecule becomes even more
stable.
There is only a small energy gap between
the 2s and 2p orbitals, and so it pays the
carbon to provide a small amount of energy
to promote an electron from the 2s to the
empty 2p to give 4 unpaired electrons. The
extra energy released when the bonds form
more than compensates for the initial input.

Note: People sometimes worry that the promoted electron is


drawn as an up-arrow, whereas it started as a down-arrow. It
simply makes the diagram look tidier - nothing very
sophisticated is going on!

Now that we've got 4 unpaired electrons ready for bonding, another
problem arises. In methane all the carbon-hydrogen bonds are identical,
but our electrons are in two different kinds of orbitals. You aren't going to
get four identical bonds unless you start from four identical orbitals.
Hybridisation
The electrons rearrange themselves again
in a process called hybridisation. This
reorganises the electrons into four identical
hybrid orbitals called sp3 hybrids (because
they are made from one s orbital and three
p orbitals). You should read "sp3" as "s p
three" - not as "s p cubed".

http://www.chemguide.co.uk/basicorg/bonding/methane.html (2 of 6)30/12/2004 11:01:15

bonding in methane - sp3 hybridisation

sp3 hybrid orbitals look a bit like half a p orbital,


and they arrange themselves in space so that
they are as far apart as possible. You can picture
the nucleus as being at the centre of a
tetrahedron (a triangularly based pyramid) with
the orbitals pointing to the corners. For clarity,
the nucleus is drawn far larger than it really is.

What happens when the bonds are formed?


Remember that hydrogen's electron is in a 1s orbital - a spherically
symmetric region of space surrounding the nucleus where there is some
fixed chance (say 95%) of finding the electron. When a covalent bond is
formed, the atomic orbitals (the orbitals in the individual atoms) merge to
produce a new molecular orbital which contains the electron pair which
creates the bond.

Four molecular orbitals are formed, looking rather like the original sp3
hybrids, but with a hydrogen nucleus embedded in each lobe. Each
orbital holds the 2 electrons that we've previously drawn as a dot and a
cross.
The principles involved - promotion of electrons if necessary, then
hybridisation, followed by the formation of molecular orbitals - can be
applied to any covalently-bound molecule.

http://www.chemguide.co.uk/basicorg/bonding/methane.html (3 of 6)30/12/2004 11:01:15

bonding in methane - sp3 hybridisation

Ethane, C2H6
The formation of molecular orbitals in ethane
Ethane isn't particularly important in its own right, but is included because
it is a simple example of how a carbon-carbon single bond is formed.
Each carbon atom in the ethane promotes an electron and then forms
sp3 hybrids exactly as we've described in methane. So just before
bonding, the atoms look like this:

The hydrogens bond with the two carbons to produce molecular orbitals
just as they did with methane. The two carbon atoms bond by merging
their remaining sp3 hybrid orbitals end-to-end to make a new molecular
orbital. The bond formed by this end-to-end overlap is called a sigma
bond. The bonds between the carbons and hydrogens are also sigma
bonds.

http://www.chemguide.co.uk/basicorg/bonding/methane.html (4 of 6)30/12/2004 11:01:15

bonding in methane - sp3 hybridisation

In any sigma bond, the most likely place to find the pair of electrons is on
a line between the two nuclei.
Free rotation about the carbon-carbon single bond
The two ends of this molecule can spin quite freely about the sigma bond
so that there are, in a sense, an infinite number of possibilities for the
shape of an ethane molecule. Some possible shapes are:

In each case, the left hand CH3 group has been kept in a constant
position so that you can see the effect of spinning the right hand one.
Other alkanes
All other alkanes will be bonded in the same way:

The carbon atoms will each promote an electron and then


hybridise to give sp3 hybrid orbitals.
The carbon atoms will join to each other by forming sigma bonds
by the end-to-end overlap of their sp3 hybrid orbitals.
Hydrogen atoms will join on wherever they are needed by
overlapping their 1s1 orbitals with sp3 hybrid orbitals on the carbon
atoms.

http://www.chemguide.co.uk/basicorg/bonding/methane.html (5 of 6)30/12/2004 11:01:15

bonding in methane - sp3 hybridisation

Where would you like to go now?


To the organic bonding menu. . .
To menu of basic organic chemistry. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/basicorg/bonding/methane.html (6 of 6)30/12/2004 11:01:15

electronic structure and atomic orbitals

ELECTRONIC STRUCTURE AND ATOMIC


ORBITALS

A simple view
In any introductory chemistry course you will have come across the
electronic structures of hydrogen and carbon drawn as:

Note: There are many places where you could still make use
of this model of the atom at A' level. It is, however, a
simplification and can be misleading. It gives the impression
that the electrons are circling the nucleus in orbits like planets
around the sun. As you will see in a moment, it is impossible
to know exactly how they are actually moving.

The circles show energy levels - representing increasing distances from


the nucleus. You could straighten the circles out and draw the electronic
structure as a simple energy diagram.

http://www.chemguide.co.uk/basicorg/bonding/orbitals.html (1 of 8)30/12/2004 11:01:21

electronic structure and atomic orbitals

Atomic orbitals
Orbits and orbitals sound similar, but they have quite different meanings.
It is essential that you understand the difference between them.
The impossibility of drawing orbits for electrons
To plot a path for something you need to know exactly where the object
is and be able to work out exactly where it's going to be an instant later.
You can't do this for electrons.

Note: In order to plot a plane's course, it is no use knowing


its exact location in mid-Atlantic if you don't know its direction
or speed. Equally it's no use knowing that it is travelling at
500 mph due west if you have no idea whether it is near
Iceland or the Azores at that particular moment.

The Heisenberg Uncertainty Principle (not required at A'level) says loosely - that you can't know with certainty both where an electron is and
where it's going next. That makes it impossible to plot an orbit for an
electron around a nucleus. Is this a big problem? No. If something is
impossible, you have to accept it and find a way around it.
Hydrogen's electron - the 1s orbital

Note: In this diagram (and the orbital diagrams that follow),


the nucleus is shown very much larger than it really is. This is
just for clarity.

http://www.chemguide.co.uk/basicorg/bonding/orbitals.html (2 of 8)30/12/2004 11:01:21

electronic structure and atomic orbitals

Suppose you had a single hydrogen atom and at a


particular instant plotted the position of the one
electron. Soon afterwards, you do the same thing,
and find that it is in a new position. You have no
idea how it got from the first place to the second.
You keep on doing this over and over again, and
gradually build up a sort of 3D map of the places that the electron is likely
to be found.
In the hydrogen case, the electron can be found anywhere within a
spherical space surrounding the nucleus. The diagram shows a crosssection through this spherical space.
95% of the time (or any other percentage you choose), the electron will
be found within a fairly easily defined region of space quite close to the
nucleus. Such a region of space is called an orbital. You can think of an
orbital as being the region of space in which the electron lives.

Note: If you wanted to be absolutely 100% sure of where the


electron is, you would have to draw an orbital the size of the
Universe!

What is the electron doing in the orbital? We don't know, we can't know,
and so we just ignore the problem! All you can say is that if an electron is
in a particular orbital it will have a particular definable energy.
Each orbital has a name.
The orbital occupied by the hydrogen electron is called a 1s orbital. The
"1" represents the fact that the orbital is in the energy level closest to the
nucleus. The "s" tells you about the shape of the orbital. s orbitals are
spherically symmetric around the nucleus - in each case, like a hollow
ball made of rather chunky material with the nucleus at its centre.

http://www.chemguide.co.uk/basicorg/bonding/orbitals.html (3 of 8)30/12/2004 11:01:21

electronic structure and atomic orbitals

The orbital on the left is a 2s orbital. This is


similar to a 1s orbital except that the region
where there is the greatest chance of
finding the electron is further from the
nucleus - this is an orbital at the second
energy level.
If you look carefully, you will notice that
there is another region of slightly higher
electron density (where the dots are thicker)
nearer the nucleus. ("Electron density" is another way of talking about
how likely you are to find an electron at a particular place.)
2s (and 3s, 4s, etc) electrons spend some of their time closer to the
nucleus than you might expect. The effect of this is to slightly reduce the
energy of electrons in s orbitals. The nearer the nucleus the electrons
get, the lower their energy.
3s, 4s (etc) orbitals get progressively further from the nucleus.
p orbitals
Not all electrons inhabit s orbitals (in fact, very few
electrons live in s orbitals). At the first energy level,
the only orbital available to electrons is the 1s
orbital, but at the second level, as well as a 2s
orbital, there are also orbitals called 2p orbitals.
A p orbital is rather like 2 identical balloons tied
together at the nucleus. The diagram on the right is
a cross-section through that 3-dimensional region
of space. Once again, the orbital shows where
there is a 95% chance of finding a particular
electron.

http://www.chemguide.co.uk/basicorg/bonding/orbitals.html (4 of 8)30/12/2004 11:01:21

electronic structure and atomic orbitals

Beyond A'level: If you imagine a horizontal plane through


the nucleus, with one lobe of the orbital above the plane and
the other beneath it, there is a zero probability of finding the
electron on that plane. So how does the electron get from
one lobe to the other if it can never pass through the plane of
the nucleus? For A'level chemistry you just have to accept
that it does! If you want to find out more, read about the wave
nature of electrons.

Unlike an s orbital, a p orbital points in a particular direction - the one


drawn points up and down the page.
At any one energy level it is possible to have three absolutely equivalent
p orbitals pointing mutually at right angles to each other. These are
arbitrarily given the symbols px, py and pz. This is simply for
convenience - what you might think of as the x, y or z direction changes
constantly as the atom tumbles in space.
The p orbitals at the second energy level are
called 2px, 2py and 2pz. There are similar
orbitals at subsequent levels - 3px, 3py, 3pz,
4px, 4py, 4pz and so on.
All levels except for the first level have p
orbitals. At the higher levels the lobes get
more elongated, with the most likely place to
find the electron more distant from the
nucleus.

Fitting electrons into orbitals


Because for the moment we are only interested in the electronic
structures of hydrogen and carbon, we don't need to concern ourselves
with what happens beyond the second energy level.
Remember:

http://www.chemguide.co.uk/basicorg/bonding/orbitals.html (5 of 8)30/12/2004 11:01:21

electronic structure and atomic orbitals

At the first level there is only one orbital - the 1s orbital.


At the second level there are four orbitals - the 2s, 2px, 2py and
2pz orbitals.
Each orbital can hold either 1 or 2 electrons, but no more.
"Electrons-in-boxes"
Orbitals can be represented as boxes with the electrons in them shown
as arrows. Often an up-arrow and a down-arrow are used to show that
the electrons are in some way different.

Beyond A'level: The need to have all electrons in an atom


different comes out of quantum theory. If they live in different
orbitals, that's fine - but if they are both in the same orbital
there has to be some subtle distinction between them.
Quantum theory allocates them a property known as "spin" which is what the arrows are intended to suggest.

A 1s orbital holding 2 electrons would be drawn as shown


on the right, but it can be written even more quickly as 1s2.
This is read as "one s two" - not as "one s squared".
You mustn't confuse the two numbers in this notation:

The order of filling orbitals


Electrons fill low energy orbitals (closer to the nucleus) before they fill
higher energy ones. Where there is a choice between orbitals of equal
energy, they fill the orbitals singly as far as possible.
The diagram (not to scale) summarises the energies of the various
orbitals in the first and second levels.
http://www.chemguide.co.uk/basicorg/bonding/orbitals.html (6 of 8)30/12/2004 11:01:21

electronic structure and atomic orbitals

Notice that the 2s orbital has a slightly lower energy than the 2p orbitals.
That means that the 2s orbital will fill with electrons before the 2p
orbitals. All the 2p orbitals have exactly the same energy.

The electronic structure of hydrogen


Hydrogen only has one electron and that will go into the orbital with the
lowest energy - the 1s orbital.
Hydrogen has an electronic structure of 1s1. We have already described
this orbital earlier.

The electronic structure of carbon


Carbon has six electrons. Two of them will be found in the 1s orbital
close to the nucleus. The next two will go into the 2s orbital. The
remaining ones will be in two separate 2p orbitals. This is because the p
orbitals all have the same energy and the electrons prefer to be on their
own if that's the case.

http://www.chemguide.co.uk/basicorg/bonding/orbitals.html (7 of 8)30/12/2004 11:01:21

electronic structure and atomic orbitals

Note: People sometimes wonder why the electrons choose


to go into the 2px and 2py orbitals rather than the 2pz. They
don't! All of the 2p orbitals are exactly equivalent, and the
names we give them are entirely arbitrary. It just looks tidier if
we call the orbitals the electrons occupy the 2px and 2py.

The electronic structure of carbon is normally written 1s22s22px12py1.

Where would you like to go now?


To the organic bonding menu. . .
To menu of basic organic chemistry. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/basicorg/bonding/orbitals.html (8 of 8)30/12/2004 11:01:21

Organic Bonding Menu

Understanding Chemistry

ORGANIC BONDING MENU

Electronic structure and orbitals . . .


An introduction to the arrangement of electrons in atoms - leading
to the modern electronic structures of carbon and hydrogen.
Bonding in methane . . .
Covers bonding in methane and ethane, including a simple look at
hybridisation.
Bonding in ethene . . .
Covers bonding in ethene, including a simple look at hybridisation.
Bonding in benzene - the Kekul structure . . .
A description of the Kekul structure for benzene and the reasons
(including hydrogenation energies) why it isn't satisfactory.
Bonding in benzene - a modern orbital view . . .
Covers a modern view of the bonding in benzene, including a
simple look at hybridisation.
Bonding in carbonyl compounds . . .
Describes the carbon-oxygen double bond in methanal (including
a simple look at hybridisation), but applies equally to other
aldehydes and ketones like ethanal and propanone.

http://www.chemguide.co.uk/basicorg/bondmenu.html (1 of 2)30/12/2004 11:01:22

Organic Bonding Menu

Electronegativity . . .
An introduction to electronegativity as it applies to organic
chemistry, including its causes. Bond polarity.

Go to menu of basic organic chemistry. . .


Go to Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/basicorg/bondmenu.html (2 of 2)30/12/2004 11:01:22

Understanding Chemistry - Basic Organic Chemistry Menu

Understanding Chemistry

BASIC ORGANIC CHEMISTRY MENU

Bonding in organic compounds . . .


Includes basic electronic structure, bonding in methane, ethene,
benzene and carbonyl compounds, and ideas about
electronegativity and bond polarity.
Organic chemistry conventions . . .
Includes how to name and draw organic compounds, and the use
of curly arrows in reaction mechanisms.
Isomerism in organic compounds . . .
Includes structural isomerism and stereoisomerism (both
geometric and optical).
Organic acids and bases . . .
Includes the acid strengths of carboxylic acids, phenols and
alcohols, and the base strengths of primary amines.

Go to Main Menu . . .

Jim Clark 2000


http://www.chemguide.co.uk/orgmenu.html (1 of 2)30/12/2004 11:01:26

Understanding Chemistry - Basic Organic Chemistry Menu

http://www.chemguide.co.uk/orgmenu.html (2 of 2)30/12/2004 11:01:26

Organic Conventions Menu

Understanding Chemistry

ORGANIC CONVENTIONS MENU

How to draw organic molecules . . .


Explains the various conventions used in drawing organic
molecules.
An introduction to naming organic molecules . . .
A guide to understanding the names of organic compounds,
including alkanes, cycloalkanes,alkenes, simple halogen
compounds, alcohols, aldehydes and ketones.
More organic names . . .
Explains the naming of carboxylic acids and their salts, esters, acyl
chlorides, acid anhydrides, amides, nitriles, amines and amino
acids. It assumes that you have already read the introductory page.
Naming aromatic compounds . . .
Looks at the special problems involved in naming compounds
containing benzene rings. It assumes that you are familiar with the
naming of simple chain compounds.
The use of curly arrows . . .
How to use curly arrows to show the movement of electron pairs or
single electrons in reaction mechanisms.

http://www.chemguide.co.uk/basicorg/convmenu.html (1 of 2)30/12/2004 11:01:27

Organic Conventions Menu

If you are interested in testing your ability to write names for organic
compounds, you might like to explore these links to Dr Phil Brown's
website:
Multiple choice tests on organic names . . .
"Type in the name" tests on organic names . . .

Go to menu of basic organic chemistry. . .


Go to Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/basicorg/convmenu.html (2 of 2)30/12/2004 11:01:27

How to draw organic molecules

DRAWING ORGANIC MOLECULES

This page explains the various ways that organic molecules can be
represented on paper or on screen - including molecular formulae, and
various forms of structural formulae.

Molecular formulae
A molecular formula simply counts the numbers of each sort of atom
present in the molecule, but tells you nothing about the way they are
joined together.
For example, the molecular formula of butane is C4H10, and the
molecular formula of ethanol is C2H6O.
Molecular formulae are very rarely used in organic chemistry, because
they don't give any useful information about the bonding in the molecule.
About the only place where you might come across them is in equations
for the combustion of simple hydrocarbons, for example:

In cases like this, the bonding in the organic molecule isn't important.

Structural formulae
A structural formula shows how the various atoms are bonded. There are
various ways of drawing this and you will need to be familiar with all of
them.
Displayed formulae
A displayed formula shows all the bonds in the molecule as individual

http://www.chemguide.co.uk/basicorg/conventions/draw.html (1 of 10)30/12/2004 11:01:38

How to draw organic molecules

lines. You need to remember that each line represents a pair of shared
electrons.
For example, this is a model of methane together with its displayed
formula:

Notice that the way the methane is drawn bears no resemblance to the
actual shape of the molecule. Methane isn't flat with 90 bond angles.
This mismatch between what you draw and what the molecule actually
looks like can lead to problems if you aren't careful.
For example, consider the simple molecule with the molecular formula
CH2Cl2. You might think that there were two different ways of arranging
these atoms if you drew a displayed formula.

The chlorines could be opposite each other or at right angles to each


other. But these two structures are actually exactly the same. Look at
how they appear as models.

One structure is in reality a simple rotation of the other one.

http://www.chemguide.co.uk/basicorg/conventions/draw.html (2 of 10)30/12/2004 11:01:38

How to draw organic molecules

Note: This is all much easier to understand if you have


actually got some models to play with. If your school or
college hasn't given you the opportunity to play around with
molecular models in the early stages of your organic
chemistry course, you might consider getting hold of a cheap
set. The models made by molymod are both cheap and easy
to use. An introductory organic set is more than adequate.
Find them at www.molymod.com.
Alternatively , get hold of some coloured Plasticene and
some used matches and make your own. It's cheaper, but
distinctly messier!

Consider a slightly more complicated molecule, C2H5Cl. The displayed


formula could be written as either of these:

But, again these are exactly the same. Look at the models.

The commonest way to draw structural formulae


For anything other than the most simple molecules, drawing a fully
displayed formula is a bit of a bother - especially all the carbon-hydrogen
bonds. You can simplify the formula by writing, for example, CH3 or CH2
http://www.chemguide.co.uk/basicorg/conventions/draw.html (3 of 10)30/12/2004 11:01:38

How to draw organic molecules

instead of showing all these bonds.


So for example, ethanoic acid would be shown in a fully displayed form
and a simplified form as:

You could even condense it further to CH3COOH, and would probably do


this if you had to write a simple chemical equation involving ethanoic
acid. You do, however, lose something by condensing the acid group in
this way, because you can't immediately see how the bonding works.
You still have to be careful in drawing structures in this way. Remember
from above that these two structures both represent the same molecule:

The next three structures all represent butane.

All of these are just versions of four carbon atoms joined up in a line. The
only difference is that there has been some rotation about some of the
carbon-carbon bonds. You can see this in a couple of models.

http://www.chemguide.co.uk/basicorg/conventions/draw.html (4 of 10)30/12/2004 11:01:38

How to draw organic molecules

Not one of the structural formulae accurately represents the shape of


butane. The convention is that we draw it with all the carbon atoms in a
straight line - as in the first of the structures above.
This is even more important when you start to have branched chains of
carbon atoms. The following structures again all represent the same
molecule - 2-methylbutane.

The two structures on the left are fairly obviously the same - all we've
done is flip the molecule over. The other one isn't so obvious until you
look at the structure in detail. There are four carbons joined up in a row,
with a CH3 group attached to the next-to-end one. That's exactly the
same as the other two structures. If you had a model, the only difference
between these three diagrams is that you have rotated some of the
bonds and turned the model around a bit.
To overcome this possible confusion, the convention is that you always
look for the longest possible chain of carbon atoms, and then draw it
horizontally. Anything else is simply hung off that chain.
It doesn't matter in the least whether you draw any side groups pointing
up or down. All of the following represent exactly the same molecule.

http://www.chemguide.co.uk/basicorg/conventions/draw.html (5 of 10)30/12/2004 11:01:38

How to draw organic molecules

If you made a model of one of them, you could turn it into any other one
simply by rotating one or more of the carbon-carbon bonds.
How to draw structural formulae in 3-dimensions
There are occasions when it is important to be able to show the precise 3D arrangement in parts of some molecules. To do this, the bonds are
shown using conventional symbols:

For example, you might want to show the 3-D arrangement of the groups
around the carbon which has the -OH group in butan-2-ol.
Butan-2-ol has the structural formula:

Using conventional bond notation, you could draw it as, for example:

http://www.chemguide.co.uk/basicorg/conventions/draw.html (6 of 10)30/12/2004 11:01:38

How to draw organic molecules

The only difference between these is a slight rotation of the bond


between the centre two carbon atoms. This is shown in the two models
below. Look carefully at them - particularly at what has happened to the
lone hydrogen atom. In the left-hand model, it is tucked behind the
carbon atom. In the right-hand model, it is in the same plane. The
change is very slight.

It doesn't matter in the least which of the two arrangements you draw.
You could easily invent other ones as well. Choose one of them and get
into the habit of drawing 3-dimensional structures that way. My own habit
(used elsewhere on this site) is to draw two bonds going back into the
paper and one coming out - as in the left-hand diagram above.
Notice that no attempt was made to show the whole molecule in 3dimensions in the structural formula diagrams. The CH2CH3 group was
left in a simple form. Keep diagrams simple - trying to show too much
detail makes the whole thing amazingly difficult to understand!
Skeletal formulae
In a skeletal formula, all the hydrogen atoms are removed from carbon
chains, leaving just a carbon skeleton with functional groups attached to
it.
For example, we've just been talking about butan-2-ol. The normal
structural formula and the skeletal formula look like this:

http://www.chemguide.co.uk/basicorg/conventions/draw.html (7 of 10)30/12/2004 11:01:38

How to draw organic molecules

In a skeletal diagram of this sort

there is a carbon atom at each junction between bonds in a chain


and at the end of each bond (unless there is something else there
already - like the -OH group in the example);
there are enough hydrogen atoms attached to each carbon to
make the total number of bonds on that carbon up to 4.

Beware! Diagrams of this sort take practice to interpret correctly - and


may well not be acceptable to your examiners (see below).
There are, however, some very common cases where they are frequently
used. These cases involve rings of carbon atoms which are surprisingly
awkward to draw tidily in a normal structural formula.
Cyclohexane, C6H12, is a ring of carbon atoms each with two hydrogens
attached. This is what it looks like in both a structural formula and a
skeletal formula.

And this is cyclohexene, which is similar but contains a double bond:

http://www.chemguide.co.uk/basicorg/conventions/draw.html (8 of 10)30/12/2004 11:01:38

How to draw organic molecules

But the commonest of all is the benzene ring, C6H6, which has a special
symbol of its own.

Note: Explaining exactly what this structure means needs


more space than is available here. It is explained in full in two
pages on the structure of benzene elsewhere in this site. It
would probably be better not to follow this link unless you are
actively interested in benzene chemistry at the moment - it
will lead you off into quite deep water!

Deciding which sort of formula to use


There's no easy, all-embracing answer to this problem. It depends more
than anything else on experience - a feeling that a particular way of
writing a formula is best for the situation you are dealing with.
Don't worry about this - as you do more and more organic chemistry, you
will probably find it will come naturally. You'll get so used to writing
formulae in reaction mechanisms, or for the structures for isomers, or in
simple chemical equations, that you won't even think about it.
There are, however, a few guidelines that you should follow.
What does your syllabus say?
Different examiners will have different preferences. Check first with your
syllabus. If you've down-loaded a copy of your syllabus from your Exam
Board's web site, it is easy to check what they say they want. Use the
"find" function on your Adobe Acrobat Reader to search the organic
section(s) of the syllabus for the word "formula".
You should also check recent exam papers and (particulary) mark
schemes to find out what sort of formula the examiners really prefer in
http://www.chemguide.co.uk/basicorg/conventions/draw.html (9 of 10)30/12/2004 11:01:38

How to draw organic molecules

given situations. You could also look at any support material published
by your Board.

Note: If you haven't got a copy of your syllabus and recent


exam papers, follow this link to find out how to get them.

What if you still aren't sure?


Draw the most detailed formula that you can fit into the space available.
If in doubt, draw a fully displayed formula. You would never lose marks
for giving too much detail.
Apart from the most trivial cases (for example, burning hydrocarbons),
never use a molecular formula. Always show the detail around the
important part(s) of a molecule. For example, the important part of an
ethene molecule is the carbon-carbon double bond - so write (at the very
least) CH2=CH2 and not C2H4.
Where a particular way of drawing a structure is important, this will
always be pointed out where it arises elsewhere on this site.

Where would you like to go now?


To the organic conventions menu. . .
To menu of basic organic chemistry. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/basicorg/conventions/draw.html (10 of 10)30/12/2004 11:01:38

bonding in benzene - the Kekul structure

BONDING IN BENZENE
The Kekul structure for benzene, C6H6
What is the Kekul structure?
Kekul was the first to suggest a sensible structure
for benzene. The carbons are arranged in a
hexagon, and he suggested alternating double and
single bonds between them. Each carbon atom
has a hydrogen attached to it.
This diagram is often simplified by leaving out all
the carbon and hydrogen atoms!

In diagrams of this sort, there is a carbon atom at each


corner. You have to count the bonds leaving each
carbon to work out how many hydrogens there are
attached to it.
In this case, each carbon has three bonds leaving it. Because carbon
atoms form four bonds, that means you are a bond missing - and that
must be attached to a hydrogen atom.
Problems with the Kekul structure
Although the Kekul structure was a good attempt in its time, there are
serious problems with it . . .
Problems with the chemistry
Because of the three double bonds, you might expect benzene to have
reactions like ethene - only more so!
Ethene undergoes addition reactions in which one of the two bonds
joining the carbon atoms breaks, and the electrons are used to bond with
additional atoms.

http://www.chemguide.co.uk/basicorg/bonding/benzene1.html (1 of 7)30/12/2004 11:01:44

bonding in benzene - the Kekul structure

Benzene rarely does this. Instead, it usually undergoes substitution


reactions in which one of the hydrogen atoms is replaced by something
new.

Note: Follow these links to get details about the addition


reactions of ethene, or the substitution reactions of benzene.

Problems with the shape


Benzene is a planar molecule (all the atoms lie in one plane), and that
would also be true of the Kekul structure. The problem is that C-C
single and double bonds are different lengths.
C-C
C=C

0.154 nm
0.134 nm

Note: "nm" means "nanometre", which is 10-9 metre.

That would mean that the hexagon would be irregular if it had the Kekul
structure, with alternating shorter and longer sides. In real benzene all
the bonds are exactly the same - intermediate in length between C-C and
C=C at 0.139 nm. Real benzene is a perfectly regular hexagon.
Problems with the stability of benzene
Real benzene is a lot more stable than the Kekul structure would give it
credit for. Every time you do a thermochemistry calculation based on the
Kekul structure, you get an answer which is wrong by about 150 kJ mol1. This is most easily shown using enthalpy changes of hydrogenation.

http://www.chemguide.co.uk/basicorg/bonding/benzene1.html (2 of 7)30/12/2004 11:01:44

bonding in benzene - the Kekul structure

Help! It doesn't matter whether you've done any


thermochemistry sums recently or not. This is all so simple
that you could understand it even if you had never done any!

Hydrogenation is the addition of hydrogen to something. If, for example,


you hydrogenate ethene you get ethane:
CH2=CH2 + H2

CH3CH3

In order to do a fair comparison with benzene (a ring


structure) we're going to compare it with cyclohexene.
Cyclohexene, C6H10, is a ring of six carbon atoms
containing just one C=C.

Note: If you are a bit shaky on names: cyclohexene: hex


means six carbons, cyclo means in a ring, ene means with a
C=C bond.

When hydrogen is added to this, cyclohexane, C6H12, is formed. The


"CH" groups become CH2 and the double bond is replaced by a single
one.

Note: cyclohexane: six carbons in a ring, but the ane


ending means NO C=C bond.

http://www.chemguide.co.uk/basicorg/bonding/benzene1.html (3 of 7)30/12/2004 11:01:44

bonding in benzene - the Kekul structure

The structures of cyclohexene and cyclohexane are usually simplified in


the same way that the Kekul structure for benzene is simplified - by
leaving out all the carbons and hydrogens.

In the cyclohexane case, for example, there is a carbon atom at each


corner, and enough hydrogens to make the total bonds on each carbon
atom up to four. In this case, then, each corner represents CH2.
The hydrogenation equation could be written:

The enthalpy change during this reaction is -120 kJ mol-1. In other words,
when 1 mole of cyclohexene reacts, 120 kJ of heat energy is evolved.

Help! "Enthalpy change" can be translated as "heat evolved


or absorbed". The negative sign shows that heat is evolved.

Where does this heat energy come from? When the reaction happens,
bonds are broken (C=C and H-H) and this costs energy. Other bonds
have to be made, and this releases energy.
Because the bonds made are stronger than those broken, more energy
is released than was used to break the original bonds and so there is a
net evolution of heat energy.
If the ring had two double bonds in it initially (cyclohexa-1,3-diene),
exactly twice as many bonds would have to be broken and exactly twice
as many made. In other words, you would expect the enthalpy change of
hydrogenation of cyclohexa-1,3-diene to be exactly twice that of
http://www.chemguide.co.uk/basicorg/bonding/benzene1.html (4 of 7)30/12/2004 11:01:44

bonding in benzene - the Kekul structure

cyclohexene - that is, -240 kJ mol-1.

Note: The name (cyclohexa-1,3-diene) is unimportant. Don't


worry about it unless you want to!

In fact, the enthalpy change is -232 kJ mol-1 - which isn't far off what we
are predicting.

Note: Thermochemistry sums often throw up discrepancies


of this sort of magnitude, and you couldn't be sure whether
there was any significance in it.

Applying the same argument to the Kekul structure for benzene (what
might be called cyclohexa-1,3,5-triene), you would expect an enthalpy
change of -360 kJ mol-1, because there are exactly three times as many
bonds being broken and made as in the cyclohexene case.

In fact what you get is -208 kJ mol-1 - not even within distance of the
predicted value!
This is very much easier to see on an enthalpy diagram. Notice that in
each case heat energy is released, and in each case the product is the
same (cyclohexane). That means that all the reactions "fall down" to the
same end point.

http://www.chemguide.co.uk/basicorg/bonding/benzene1.html (5 of 7)30/12/2004 11:01:44

bonding in benzene - the Kekul structure

Heavy lines, solid arrows and bold numbers represent real changes.
Predicted changes are shown by dotted lines and italics.
The most important point to notice is that real benzene is much lower
down the diagram than the Kekul form predicts. The lower down a
substance is, the more energetically stable it is.
This means that real benzene is about 150 kJ mol-1 more stable than the
Kekul structure gives it credit for. This increase in stability of benzene is
known as the delocalisation energy or resonance energy of benzene.
The first term (delocalisation energy) is the more commonly used.

Note: If you look at the diagram closely, you will see that
cyclohexa-1,3-diene is also a shade more stable than
expected. There is a tiny amount of delocalisation energy
involved here as well.

http://www.chemguide.co.uk/basicorg/bonding/benzene1.html (6 of 7)30/12/2004 11:01:44

bonding in benzene - the Kekul structure

Why is benzene so much more stable than the Kekul structure


suggests? To explain that needs a separate article! Follow the first link
below.

Where would you like to go now?


To read about the modern view of the structure of benzene. . .
To the organic bonding menu. . .
To menu of basic organic chemistry. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/basicorg/bonding/benzene1.html (7 of 7)30/12/2004 11:01:44

electrophilic addition reactions menu

Understanding Chemistry

ELECTROPHILIC ADDITION MECHANISMS MENU

Addition to symmetrical alkenes


Covers addition to symmetrical alkenes like ethene and cyclohexene. A
symmetrical alkene has the same groups attached to both ends of the
carbon-carbon double bond.
What is electrophilic addition? . . .
An explanation of the terms addition and electrophile, together
with a general mechanism for these reactions.
The reaction with hydrogen halides . . .
The mechanism for the reaction between ethene (and
cyclohexene) and hydrogen halides (like hydrogen bromide).
The reaction with sulphuric acid . . .
The mechanism for the reaction between ethene (and
cyclohexene) and sulphuric acid.
The reaction with bromine . . .
The mechanism for the reaction between ethene (and
cyclohexene) and bromine.

Addition to unsymmetrical alkenes


http://www.chemguide.co.uk/mechanisms/eladdmenu.html (1 of 3)30/12/2004 11:01:45

electrophilic addition reactions menu

Covers addition to unsymmetrical alkenes like propene. An


unsymmetrical alkene has different groups attached to each end of the
carbon-carbon double bond.

Warning! Don't even think about reading articles in this


section until you are sure you understand the corresponding
reaction(s) above!

Carbocations (carbonium ions) and their stability . . .


Essential pre-reading before you tackle anything else in this
section.
Why unsymmetric alkenes are a problem . . .
Explains the reasons behind Markovnikov's Rule, and gives a
general mechanism for these more awkward reactions. This is also
essential reading before you look at specific reactions.
The reaction with hydrogen halides . . .
The mechanism for the reaction between propene and hydrogen
halides (like hydrogen bromide).
The reaction with sulphuric acid . . .
The mechanism for the reaction between propene and sulphuric
acid.
The reaction with bromine . . .
The mechanism for the reaction between propene and bromine.

http://www.chemguide.co.uk/mechanisms/eladdmenu.html (2 of 3)30/12/2004 11:01:45

electrophilic addition reactions menu

Go to menu of other types of mechanism. . .


Go to Main Menu . . .

You might also be interested in:


properties and reactions of alkenes . . .
A survey of all the physical and chemical properties of alkenes
required by UK A level syllabuses.

Jim Clark 2000 (modified 2004)

http://www.chemguide.co.uk/mechanisms/eladdmenu.html (3 of 3)30/12/2004 11:01:45

What is electrophilic addition?

ELECTROPHILIC ADDITION

Background
Electrophilic addition happens in many of the reactions of compounds
containing carbon-carbon double bonds - the alkenes.
The structure of ethene
We are going to start by looking at ethene, because it is the simplest
molecule containing a carbon-carbon double bond. What is true of C=C
in ethene will be equally true of C=C in more complicated alkenes.
Ethene, C2H4, is often modelled as shown on
the right. The double bond between the carbon
atoms is, of course, two pairs of shared
electrons. What the diagram doesn't show is
that the two pairs aren't the same as each other.
One of the pairs of electrons is held on the line between the two carbon
nuclei as you would expect, but the other is held in a molecular orbital
above and below the plane of the molecule. A molecular orbital is a
region of space within the molecule where there is a high probability of
finding a particular pair of electrons.
In this diagram, the line between the two
carbon atoms represents a normal bond - the
pair of shared electrons lies in a molecular
orbital on the line between the two nuclei
where you would expect them to be. This
sort of bond is called a sigma bond.
The other pair of electrons is found
somewhere in the shaded part above and
below the plane of the molecule. This bond is called a pi bond. The
electrons in the pi bond are free to move around anywhere in this shaded
region and can move freely from one half to the other.

http://www.chemguide.co.uk/mechanisms/eladd/whatis.html (1 of 8)30/12/2004 11:01:53

What is electrophilic addition?

Note: This diagram shows a side view of an ethene


molecule. The dotted lines to two of the hydrogens show
bonds going back into the screen or paper away from you.
The wedge shapes show bonds coming out towards you.

The pi electrons are not as fully under the control of the carbon nuclei as
the electrons in the sigma bond and, because they lie exposed above
and below the rest of the molecule, they are relatively open to attack by
other things.

Note: Check your syllabus to see if you need to know how a


pi bond is formed. Haven't got a syllabus? Find out how to
get one by following this link.
If you do need to know about the bonding in ethene in detail,
follow this link as well.

Electrophiles
An electrophile is something which is attracted to electron-rich regions in
other molecules or ions. Because it is attracted to a negative region, an
electrophile must be something which carries either a full positive charge,
or has a slight positive charge on it somewhere.

Note: The ending ". . phile" means a liking for. For example,
a francophile is someone who likes the French; an anglophile
is someone who likes the English.

http://www.chemguide.co.uk/mechanisms/eladd/whatis.html (2 of 8)30/12/2004 11:01:53

What is electrophilic addition?

Ethene and the other alkenes are attacked by


electrophiles. The electrophile is normally the
slightly positive ( +) end of a molecule like hydrogen bromide, HBr.

Note: If you aren't sure about why some bonds are polar,
read the page on electronegativity.
Use the BACK button on your browser to return to this page.

Electrophiles are strongly attracted to the exposed electrons in the pi


bond and reactions happen because of that initial attraction - as you will
see shortly.
You might wonder why fully positive ions like sodium, Na+, don't react
with ethene. Although these ions may well be attracted to the pi bond,
there is no possibility of the process going any further to form bonds
between sodium and carbon, because sodium forms ionic bonds,
whereas carbon normally forms covalent ones.
Addition reactions
In a sense, the pi bond is an unnecessary bond. The structure would
hold together perfectly well with a single bond rather than a double bond.
The pi bond often breaks and the electrons in it are used to join other
atoms (or groups of atoms) onto the ethene molecule. In other words,
ethene undergoes addition reactions.
For example, using a general molecule X-Y . . .

Summary: electrophilic addition reactions

http://www.chemguide.co.uk/mechanisms/eladd/whatis.html (3 of 8)30/12/2004 11:01:53

What is electrophilic addition?

An addition reaction is a reaction in which two molecules join together to


make a bigger one. Nothing is lost in the process. All the atoms in the
original molecules are found in the bigger one.
An electrophilic addition reaction is an addition reaction which happens
because what we think of as the "important" molecule is attacked by an
electrophile. The "important" molecule has a region of high electron
density which is attacked by something carrying some degree of positive
charge.

Note: When we talk about reactions of alkenes like ethene,


we think of the ethene as being attacked by other molecules
such as hydrogen bromide. Because ethene is the molecule
we are focusing on, we quite arbitrarily think of it as the
central molecule and hydrogen bromide as its attacker.
There's no real justification for this, of course, apart from the
fact that it helps to put things in some sort of logical pattern.
In reality, the molecules just collide and may react if they
have enough energy and if they are lined up correctly.

Understanding the electrophilic addition mechanism


The mechanism for the reaction between ethene and a molecule X-Y
It is very unlikely that any two different atoms
joined together will have the same
electronegativity. We are going to assume that
Y is more electronegative than X, so that the
pair of electrons is pulled slightly towards the Y
end of the bond. That means that the X atom carries a slight positive
charge.

http://www.chemguide.co.uk/mechanisms/eladd/whatis.html (4 of 8)30/12/2004 11:01:53

What is electrophilic addition?

Note: Once again, if you aren't sure about electronegativity


and bond polarity follow this link before you read on.
Use the BACK button on your browser to return to this page.

The slightly positive X atom is an electrophile and is attracted to the


exposed pi bond in the ethene. Now imagine what happens as they
approach each other.

You are now much more likely to find the electrons in the half of the pi
bond nearest the XY. As the process continues, the two electrons in the
pi bond move even further towards the X until a covalent bond is made.
The electrons in the X-Y bond are pushed entirely onto the Y to give a
negative Y- ion.

http://www.chemguide.co.uk/mechanisms/eladd/whatis.html (5 of 8)30/12/2004 11:01:53

What is electrophilic addition?

Help! Why does the carbon atom have a positive charge?


The pi bond was originally made using an electron from each
carbon atom, but both of these electrons have now been
used to make a bond to the X atom. This leaves the righthand carbon atom an electron short - hence positively
charged.
Note also that we are only showing one of the pairs of
electrons around the Y- ion. There will be other lone pairs as
well, but we are only actually interested in the one we've
drawn.

Important term
An ion in which the positive charge is carried on a
carbon atom is called a carbocation or a
carbonium ion (an older term).

http://www.chemguide.co.uk/mechanisms/eladd/whatis.html (6 of 8)30/12/2004 11:01:53

What is electrophilic addition?

In the final stage of the reaction the


electrons in the lone pair on the Y- ion are
strongly attracted towards the positive
carbon atom. They move towards it and
form a co-ordinate (dative covalent) bond
between the Y and the carbon.

Help! A co-ordinate (dative covalent) bond is simply a


covalent bond in which both shared electrons originate from
the same atom. The bond formed between the X and the
other carbon atom was also a co-ordinate bond. Once a coordinate bond has been formed there is no difference
whatsoever between it and any other covalent bond.

How to write this mechanism in an exam


The movements of the various electron pairs are shown using curly
arrows.

Help! If you aren't sure about the use of curly arrows in


mechanisms, you must follow this link before you go on.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/mechanisms/eladd/whatis.html (7 of 8)30/12/2004 11:01:53

What is electrophilic addition?

Don't leave this page until you are sure that you understand how this
relates to the electron pair movements drawn in the previous diagrams.

Where would you like to go now?


To menu of electrophilic addition reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/eladd/whatis.html (8 of 8)30/12/2004 11:01:53

bonding in ethene - sp2 hybridisation

BONDING IN ETHENE

Important! You will find this much easier to understand if


you first read the article about the bonding in methane.
You may also find it useful to read the article on orbitals if you
aren't sure about simple orbital theory.

Ethene, C2H4
The simple view of the bonding in ethene
At a simple level, you will have drawn ethene
showing two bonds between the carbon atoms.
Each line in this diagram represents one pair of
shared electrons.
Ethene is actually much more interesting than this.
An orbital view of the bonding in ethene
Ethene is built from hydrogen atoms (1s1) and carbon atoms
(1s22s22px12py1).
The carbon atom doesn't have enough unpaired electrons to form the
required number of bonds, so it needs to promote one of the 2s2 pair into
the empty 2pz orbital. This is exactly the same as happens whenever
carbon forms bonds - whatever else it ends up joined to.

http://www.chemguide.co.uk/basicorg/bonding/ethene.html (1 of 5)30/12/2004 11:01:58

bonding in ethene - sp2 hybridisation

Important! If this isn't really clear to you, you must go and


read the article about the bonding in methane.

Now there's a difference, because each carbon is only joining to three


other atoms rather than four - as in methane or ethane. When the carbon
atoms hybridise their outer orbitals before forming bonds, this time they
only hybridise three of the orbitals rather than all four. They use the 2s
electron and two of the 2p electrons, but leave the other 2p electron
unchanged.
The new orbitals formed are called sp2
hybrids, because they are made by an s
orbital and two p orbitals reorganising
themselves. sp2 orbitals look rather like
sp3 orbitals that you have already come
across in the bonding in methane,
except that they are shorter and fatter.
The three sp2 hybrid orbitals arrange
themselves as far apart as possible which is at 120 to each other in a plane. The remaining p orbital is at
right angles to them.
The two carbon atoms and four hydrogen atoms would look like this
before they joined together:

The various atomic orbitals which are pointing


towards each other now merge to give
molecular orbitals, each containing a bonding
pair of electrons. These are sigma bonds just like those formed by end-to-end overlap
of atomic orbitals in, say, ethane.

http://www.chemguide.co.uk/basicorg/bonding/ethene.html (2 of 5)30/12/2004 11:01:58

bonding in ethene - sp2 hybridisation

The p orbitals on each carbon aren't pointing towards each other, and so
we'll leave those for a moment. In the diagram, the black dots represent
the nuclei of the atoms.
Notice that the p orbitals are so close that they are overlapping sideways.
This sideways overlap also creates a
molecular orbital, but of a different kind. In
this one the electrons aren't held on the line
between the two nuclei, but above and below
the plane of the molecule. A bond formed in
this way is called a pi bond.

For clarity, the sigma bonds are shown using lines - each line
representing one pair of shared electrons. The various sorts of line show
the directions the bonds point in. An ordinary line represents a bond in
the plane of the screen (or the paper if you've printed it), a broken line is
a bond going back away from you, and a wedge shows a bond coming
out towards you.

Note: The really interesting bond in ethene is the pi bond. In


almost all cases where you will draw the structure of ethene,
the sigma bonds will be shown as lines.

Be clear about what a pi bond is. It is a region of space in which you can
find the two electrons which make up the bond. Those two electrons can
live anywhere within that space. It would be quite misleading to think of
one living in the top and the other in the bottom.

http://www.chemguide.co.uk/basicorg/bonding/ethene.html (3 of 5)30/12/2004 11:01:58

bonding in ethene - sp2 hybridisation

Beyond A'level: This is another example of the curious


behaviour of electrons. How do the electrons get from one
half of the pi bond to the other if they are never found in
between? It's an unanswerable question if you think of
electrons as particles.

Even if your syllabus doesn't expect you to know how a pi bond is


formed, it will expect you to know that it exists. The pi bond dominates
the chemistry of ethene. It is very vulnerable to attack - a very negative
region of space above and below the plane of the molecule. It is also
somewhat distant from the control of the nuclei and so is a weaker bond
than the sigma bond joining the two carbons.

Important! Check your syllabus! Find out whether you


actually need to know how a pi bond is formed. Don't forget
to look in the bonding section of your syllabus as well as
under ethene. If you don't need to know it, there's no point in
learning it! You will, however, need to know that a pi bond
exists - that the two bonds between the carbon atoms in
ethene aren't both the same.
If you haven't got a copy of your syllabus, find out how to
download one

All double bonds (whatever atoms they might be joining) will consist of a
sigma bond and a pi bond.

Where would you like to go now?


To the organic bonding menu. . .
To menu of basic organic chemistry. . .

http://www.chemguide.co.uk/basicorg/bonding/ethene.html (4 of 5)30/12/2004 11:01:58

bonding in ethene - sp2 hybridisation

To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/basicorg/bonding/ethene.html (5 of 5)30/12/2004 11:01:58

electronegativity - polar bonds in organic compounds

ELECTRONEGATIVITY

This page deals with electronegativity in an organic chemistry context. If


you want a wider view of electronegativity, there is a link at the bottom of
the page.

What is electronegativity?
Electronegativity is a measure of the tendency of an atom to attract a
bonding pair of electrons. The Pauling scale is the most commonly used.
Fluorine (the most electronegative element) is given a value of 4.0, and
values range down to caesium and francium which are the least
electronegative at 0.7.
What happens if two atoms of equal electronegativity bond
together?
The most obvious example of this is the bond between two carbon
atoms. Both atoms will attract the bonding pair to exactly the same
extent. That means that on average the electron pair will be found half
way between the two nuclei, and you could draw a picture of the bond
like this:

It is important to realise that this is an average picture. The electrons are


actually in a sigma orbital, and are moving constantly within that orbital.

http://www.chemguide.co.uk/basicorg/bonding/eneg.html (1 of 8)30/12/2004 11:02:05

electronegativity - polar bonds in organic compounds

Help! A sigma orbital is a molecular orbital formed by end-toend overlap between two atomic orbitals. If you aren't happy
about this, read the articles on orbitals and the bonding in
methane and ethane.

The carbon-fluorine bond


Fluorine is much more electronegative than carbon. The actual values on
the Pauling scale are
carbon
fluorine

2.5
4.0

That means that fluorine attracts the bonding pair much more strongly
than carbon does. The bond - on average - will look like this:

Why is fluorine more electronegative than carbon?


A simple dots-and-crosses diagram of a C-F bond is perfectly adequate
to explain it.

The bonding pair is in the second energy level of both carbon and
fluorine, so in the absence of any other effect, the distance of the pair
from both nuclei would be the same.
The electron pair is shielded from the full force of both nuclei by the 1s
electrons - again there is nothing to pull it closer to one atom than the

http://www.chemguide.co.uk/basicorg/bonding/eneg.html (2 of 8)30/12/2004 11:02:05

electronegativity - polar bonds in organic compounds

other.
BUT, the fluorine nucleus has 9 protons whereas the carbon nucleus has
only 6.
Allowing for the shielding effect of the 1s electrons, the bonding pair feels
a net pull of about 4+ from the carbon, but about 7+ from the fluorine. It is
this extra nuclear charge which pulls the bonding pair (on average)
closer to the fluorine than the carbon.

Help! You have to imagine what the bonding pair "sees" if it


looks in towards the nucleus. In the carbon case, it sees 6
positive protons, and 2 negative electrons. That means that
there will be a net pull from the carbon of about 4+. The
shielding wouldn't actually be quite as high as 2-, because
the 1s electrons spend some of their time on the far side of
the carbon nucleus - and so aren't always between the
bonding pair and the nucleus.
Incidentally, thinking about electrons looking towards the
nucleus may be helpful in picturing what is going on, but
avoid using terms like this in exams.

The carbon-chlorine bond


The electronegativities are:
carbon
chlorine

2.5
3.0

The bonding pair of electrons will be dragged towards the chlorine but
not as much as in the fluorine case. Chlorine isn't as electronegative as
fluorine.
Why isn't chlorine as electronegative as fluorine?
Chlorine is a bigger atom than fluorine.

http://www.chemguide.co.uk/basicorg/bonding/eneg.html (3 of 8)30/12/2004 11:02:05

electronegativity - polar bonds in organic compounds

fluorine: 1s22s22px22py22pz1
chlorine: 1s22s22px22py22pz23s23px23py23pz1

Help! If you aren't happy about this, read the article on


orbitals. Use the BACK button on your browser to get back to
here again.

In the chlorine case, the bonding pair will be shielded by all the 1-level
and 2-level electrons. The 17 protons on the nucleus will be shielded by
a total of 10 electrons, giving a net pull from the chlorine of about 7+.
That is the same as the pull from the fluorine, but with chlorine the
bonding pair starts off further away from the nucleus because it is in the
3-level. Since it is further away, it feels the pull from the nucleus less
strongly.

Bond polarity and inductive effects


Polar bonds
Think about the carbon-fluorine bond again. Because the bonding pair is
pulled towards the fluorine end of the bond, that end is left rather more
negative than it would otherwise be. The carbon end is left rather short of
electrons and so becomes slightly positive.

The symbols + and - mean "slightly positive" and "slightly negative".


You read + as "delta plus" or "delta positive".
We describe a bond having one end slightly positive and the other end
slightly negative as being polar.
Inductive effects
http://www.chemguide.co.uk/basicorg/bonding/eneg.html (4 of 8)30/12/2004 11:02:05

electronegativity - polar bonds in organic compounds

An atom like fluorine which can pull the bonding pair away from the atom
it is attached to is said to have a negative inductive effect.
Most atoms that you will come across have a negative inductive effect
when they are attached to a carbon atom, because they are mostly more
electronegative than carbon.
You will come across some groups of atoms which have a slight positive
inductive effect - they "push" electrons towards the carbon they are
attached to, making it slightly negative.
Inductive effects are sometimes given symbols: -I (a negative inductive
effect) and +I (a positive inductive effect).

Note: You should be aware of terms like "negative inductive


effect", but don't get bogged down in them. Provided that you
understand what happens when electronegative atoms like
fluorine or chlorine are attached to carbon atoms in terms of
the polarity of the bonds, that's really all you need for most
purposes.

Some important examples of polar bonds


Hydrogen bromide (and other hydrogen halides)

Bromine (and the other halogens) are all more electronegative than
hydrogen, and so all the hydrogen halides have polar bonds with the
hydrogen end slightly positive and the halogen end slightly negative.

http://www.chemguide.co.uk/basicorg/bonding/eneg.html (5 of 8)30/12/2004 11:02:05

electronegativity - polar bonds in organic compounds

Help! Halogen: a member of group VII of the Periodic Table


- fluorine, chlorine, bromine and iodine.
Halide: a compound of one of these - e.g. hydrogen chloride,
hydrogen bromide, etc.

The polarity of these molecules is important in their reactions with


alkenes.

Note: These reactions are explored in the section dealing


with the addition of hydrogen halides to alkenes.

The carbon-bromine bond in halogenoalkanes

Note: You may come across halogenoalkanes under the


names "haloalkanes" or "alkyl halides".

Bromine is more electronegative than carbon and so the bond is


polarised in the way that we have already described with C-F and C-Cl.

The polarity of the carbon-halogen bonds is important in the reactions of


the halogenoalkanes.

http://www.chemguide.co.uk/basicorg/bonding/eneg.html (6 of 8)30/12/2004 11:02:05

electronegativity - polar bonds in organic compounds

Note: This link will take you to the nucleophilic substitution


reactions of the halogenoalkanes in which this polarity is
important.

The carbon-oxygen double bond


An orbital model of the C=O bond in methanal, HCHO, looks like this:

Note: If you aren't sure about this, read the article on


bonding in the carbonyl group (C=O).

The very electronegative oxygen atom pulls both bonding pairs towards
itself - in the sigma bond and the pi bond. That leaves the oxygen fairly
negative and the carbon fairly positive.

Note: You can read about addition reactions or additionelimination reactions of compounds containing carbonoxygen double bonds elswhere on this site.

http://www.chemguide.co.uk/basicorg/bonding/eneg.html (7 of 8)30/12/2004 11:02:05

electronegativity - polar bonds in organic compounds

Where would you like to go now?


To look at electronegativity in a wider context. . .
To the organic bonding menu. . .
To menu of basic organic chemistry. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/basicorg/bonding/eneg.html (8 of 8)30/12/2004 11:02:05

electrophilic addition - symmetrical alkenes and hydrogen halides

THE REACTION BETWEEN SYMMETRICAL


ALKENES AND THE HYDROGEN HALIDES

This page gives you the facts and a simple, uncluttered mechanism for
the electrophilic addition reactions between the hydrogen halides and
alkenes like ethene and cyclohexene. Hydrogen halides include
hydrogen chloride and hydrogen bromide. If you want the mechanisms
explained to you in detail, there is a link at the bottom of the page.

Electrophilic addition reactions involving hydrogen


bromide
The facts
Alkenes react with hydrogen bromide in the cold. The double bond
breaks and a hydrogen atom ends up attached to one of the carbons and
a bromine atom to the other.
In the case of ethene, bromoethane is formed.

Note: Be careful when you write the names of the addition


products that you change the ene ending in the original
alkene (showing the C=C) into an ane ending (showing that it
has been replaced by C-C).

http://www.chemguide.co.uk/mechanisms/eladd/symhbr.html (1 of 5)30/12/2004 11:02:10

electrophilic addition - symmetrical alkenes and hydrogen halides

With cyclohexene you get bromocyclohexane.

The structures of the cyclohexene and the bromocyclohexane are often


simplified:

Note: Each corner in one of these diagrams represents a


carbon atom. Each carbon atom has enough hydrogens
attached to make the total number of bonds up to 4.
In the case of the bromocyclohexane, it isn't necessary to
write the new hydrogen into the diagram, but it is helpful to
put it there to emphasise that addition has happened.

Be sure that you understand the relationship between these simplified


diagrams and the full structures.
The mechanisms
The reactions are examples of electrophilic addition.
With ethene and HBr:

http://www.chemguide.co.uk/mechanisms/eladd/symhbr.html (2 of 5)30/12/2004 11:02:10

electrophilic addition - symmetrical alkenes and hydrogen halides

and with cyclohexene:

Electrophilic addition reactions involving the other


hydrogen halides
The facts
Hydrogen chloride and the other hydrogen halides add on in exactly the
same way. For example, hydrogen chloride adds to ethene to make
chloroethane:

The only difference is in how fast the reactions happen with the different
hydrogen halides. The rate of reaction increases as you go from HF to
HCl to HBr to HI.
HF slowest reaction
HCl
HBr
HI

fastest reaction

http://www.chemguide.co.uk/mechanisms/eladd/symhbr.html (3 of 5)30/12/2004 11:02:10

electrophilic addition - symmetrical alkenes and hydrogen halides

The reason for this is that as the halogen atoms get bigger, the strength
of the hydrogen-halogen bond falls. Bond strengths (measured in
kilojoules per mole) are:
H-F 568
H-Cl 432
H-Br 366
H-I

298

Note: You may find slightly different values depending on


which data source you use. It doesn't matter - the differences
are minor and the pattern is always the same.

As you have seen in the HBr case, in the first step of the mechanism the
hydrogen-halogen bond gets broken. If the bond is weaker, it will break
more readily and so the reaction is more likely to happen.
The mechanisms
The reactions are still examples of electrophilic addition.
With ethene and HCl, for example:

This is exactly the same as the mechanism for the reaction between
ethene and HBr, except that we've replaced Br by Cl.
http://www.chemguide.co.uk/mechanisms/eladd/symhbr.html (4 of 5)30/12/2004 11:02:10

electrophilic addition - symmetrical alkenes and hydrogen halides

All the other mechanisms for symmetrical alkenes and the hydrogen
halides would be done in the same way.

Suggestion: Find out if your syllabus mentions a particular


hydrogen halide, and learn that mechanism. You can simply
swap the halogen atom if a different hydrogen halide comes
up in an exam.
Haven't got a syllabus? Follow this link to find out how to get
one.

Where would you like to go now?


Help! Talk me through this mechanism . . .
Look at the same reactions involving unsymmetrical
alkenes . . .
To menu of electrophilic addition reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/eladd/symhbr.html (5 of 5)30/12/2004 11:02:10

Explaining electrophilic addition involving hydrogen halides

EXPLAINING THE REACTION BETWEEN


SYMMETRICAL ALKENES AND THE HYDROGEN
HALIDES

This page guides you through the mechanism for the electrophilic
addition of hydrogen halides such as hydrogen bromide with symmetrical
alkenes like ethene or cyclohexene. Unsymmetrical alkenes are covered
separately, and you will find a link at the bottom of the page.

Electrophilic addition reactions involving hydrogen


bromide
Hydrogen bromide is chosen as a typical
hydrogen halide. Bromine is more
electronegative than hydrogen. That means
that the bonding pair of electrons is pulled
towards the bromine end of the bond, and so
the hydrogen bromide molecule is polar.

Note: If you aren't sure about electronegativity and bond


polarity follow this link before you read on.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/mechanisms/eladd/symhbrtt.html (1 of 8)30/12/2004 11:02:15

Explaining electrophilic addition involving hydrogen halides

The reaction of ethene with hydrogen bromide


The structure of ethene is shown in the
diagram on the right. The pi bond is an
orbital above and below the plane of the rest
of the molecule, and relatively exposed to
things around it. The two electrons in this
orbital are highly attractive to anything which
is positively charged.

Note: If you aren't sure about this, it would be a good idea to


read the introductory page on electrophilic addition before
you go on.
Use the BACK button on your browser to return to this page.

The slightly positive hydrogen atom in the hydrogen bromide acts as an


electrophile, and is strongly attracted to the electrons in the pi bond.

Electrophile: A substance with a strong attraction to a


negative region in another substance. Electrophiles are either
fully positive ions, or the slightly positive end of a polar
molecule.

The electrons from the pi bond move down towards the


slightly positive hydrogen atom.
In the process, the electrons in the H-Br bond are
repelled down until they are entirely on the bromine
atom, producing a bromide ion.

http://www.chemguide.co.uk/mechanisms/eladd/symhbrtt.html (2 of 8)30/12/2004 11:02:15

Explaining electrophilic addition involving hydrogen halides

Help! If you aren't sure about the use of curly arrows in


mechanisms, you must follow this link before you go on.
Use the BACK button on your browser to return to this page.

That leaves you with these two ions at this half-way stage of the reaction:

The ion with a positive charge on the carbon atom is called a


carbocation or carbonium ion (an older term).
Why is there a positive charge on the carbon atom? The pi bond was
originally made up of an electron from each of the carbon atoms. Both of
those electrons have been used to make a new bond to the hydrogen.
That leaves the right-hand carbon an electron short - hence positively
charged.
In the second stage of the mechanism, the lone pair of electrons on the
bromide ion is strongly attracted to the positive carbon and moves
towards it until a bond is formed.

http://www.chemguide.co.uk/mechanisms/eladd/symhbrtt.html (3 of 8)30/12/2004 11:02:15

Explaining electrophilic addition involving hydrogen halides

Note: For clarity only one of the lone pairs around the
bromide ion is shown. That's perfectly acceptable, because
the other three lone pairs aren't involved in the process - they
are pointing in the wrong directions.

The overall mechanism is therefore

The reaction of cyclohexene with hydrogen bromide


Cyclohexene is chosen an an example of a fairly commonly used
symmetrical alkene. The fact that it is a ring structure doesn't make any
difference to the mechanism.
The full structure of cyclohexene is

but it is often abbreviated to

In this diagram, there is a carbon atom at each corner, and enough


hydrogens attached to each carbon to bring the total number of bonds
per carbon atom up to 4.

http://www.chemguide.co.uk/mechanisms/eladd/symhbrtt.html (4 of 8)30/12/2004 11:02:15

Explaining electrophilic addition involving hydrogen halides

The double bond has an easily attacked pi bond exactly as in ethene,


and the electrons in that bond are attracted towards the slightly positive
hydrogen atom in the HBr.
Once again, the pi bond electrons swing to make a bond with the
hydrogen, and push the electrons in the H-Br bond fully onto the
bromine, making a bromide ion.

Care! Think carefully about which way the pi bond electrons


swing. In this case, think of them as being pivotted about the
top carbon atom. It is therefore that carbon atom which is
joined to the new hydrogen.

In the second stage, one of the lone pairs of electrons on the bromide ion
is attracted to the positively charged carbon atom and forms a bond with
it.

The overall mechanism is therefore:

Electrophilic addition reactions involving the other


hydrogen halides
The mechanisms
The other hydrogen halides behave in exactly the same way as hydrogen
http://www.chemguide.co.uk/mechanisms/eladd/symhbrtt.html (5 of 8)30/12/2004 11:02:15

Explaining electrophilic addition involving hydrogen halides

bromide. For example, compare the reaction between ethene and


hydrogen bromide with the one between ethene and hydrogen chloride.

There's no need to learn both mechanisms. As long as you know one of


them, all you have to do is swap one halogen atom for another. That's
equally true for hydrogen fluoride or hydrogen iodide.
The different rates of reaction
The rate of reaction increases as you go from HF to HCl to HBr to HI.
HF slowest reaction
HCl
HBr
HI

fastest reaction

The reason for this is that as the halogen atoms get bigger, the strength
of the hydrogen-halogen bond falls. Bond strengths (measured in
kilojoules per mole) are:

http://www.chemguide.co.uk/mechanisms/eladd/symhbrtt.html (6 of 8)30/12/2004 11:02:15

Explaining electrophilic addition involving hydrogen halides

H-F 568
H-Cl 432
H-Br 366
H-I

298

In the first step of these mechanisms, the hydrogen-halogen bond breaks


as the electron pair is forced down onto the halogen atom.
Breaking bonds needs energy, and if the bond is weaker, it will break
more easily - needing less energy. That means that the activation energy
for the reactions will fall as you go from hydrogen fluoride to hydrogen
iodide. The lower the activation energy, the faster the reaction.

Activation energy: The minimum energy needed before a


reaction will occur. In this case it is the energy needed to
break the various bonds and make the carbocation and the
halide ion.

Beware! A red herring!


People sometimes get confused because there is another tempting place
to look for the reason why the reaction rates are different between the
various hydrogen halides.
The halogens have different electronegativities - with fluorine being the
most electronegative and iodine the least. That means that the hydrogen
in HF will have the greatest positive charge and so will be attracted most
strongly to the pi bond.
It would be tempting to think that that would produce the fastest reaction

http://www.chemguide.co.uk/mechanisms/eladd/symhbrtt.html (7 of 8)30/12/2004 11:02:15

Explaining electrophilic addition involving hydrogen halides

- but not so! Although the HF may well be attracted most strongly,
attraction alone isn't enough. If anything is to happen, bonds have to be
broken - and here HF is at a disadvantage, because the bond is very
strong.
The lesson from all this
When you are trying to find reasons for differing rates of reactions,
always look first at differences in bond strengths. Electronegativity
differences may be interesting, but rarely give you the answer you want!

Where would you like to go now?


Look at the same reactions involving unsymmetrical
alkenes . . .
To menu of electrophilic addition reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/eladd/symhbrtt.html (8 of 8)30/12/2004 11:02:15

The use of curly arrows to show electron movements in reaction mechanisms

USING CURLY ARROWS IN REACTION


MECHANISMS

This page explains the use of curly arrows to show the movement both of
electron pairs and of single electrons during organic reaction
mechanisms.
You can jump straight to the movement of single electrons further down
this page if that is all you are interested in for the moment (for example, if
you are currently working on free radical reactions).

Using curly arrows to show the movement of electron pairs


Curly arrows (and that's exactly what they are called!) are used in
mechanisms to show the various electron pairs moving around. You
mustn't use them for any other purpose.

The arrow tail is where the electron pair starts from. That's always
fairly obvious, but you must show the electron pair either as a
bond or, if it is a lone pair, as a pair of dots. Remember that a lone
pair is a pair of electrons at the bonding level which isn't currently
being used to join on to anything else.
The arrow head is where you want the electron pair to end up.

For example, in the reaction between ethene and


hydrogen bromide, one of the two bonds between the
two carbon atoms breaks. That bond is simply a pair of
electrons.
Those electrons move to form a new bond with the
hydrogen from the HBr. At the same time the pair of electrons in the
hydrogen-bromine bond moves down on to the bromine atom.

http://www.chemguide.co.uk/basicorg/conventions/curlies.html (1 of 4)30/12/2004 11:02:20

The use of curly arrows to show electron movements in reaction mechanisms

There's no need to draw the pairs of electrons in the


bonds as two dots. Drawing the bond as a line is
enough, but you could put two dots in as well if you
wanted to.
Notice that the arrow head points between the C and
H because that's where the electron pair ends up. Notice also that the
electron movement between the H and Br is shown as a curly arrow even
though the electron pair moves straight down. You have to show electron
pair movements as curly arrows - not as straight ones.
The second stage of this reaction nicely illustrates how you use a curly
arrow if a lone pair of electrons is involved.
The first stage leaves you with a positive charge on
the right hand carbon atom and a negative bromide
ion. You can think of the electrons shown on the
bromide ion as being the ones which originally made
up the hydrogen-bromine bond.

Note: There are another three lone pairs around the outside
of the bromide ion - making four in all. These aren't normally
shown because they don't actually do anything new and
interesting!
However, it is essential that you show the lone pair you are
interested in as a pair of dots. If you don't, you risk losing
marks in an exam.

http://www.chemguide.co.uk/basicorg/conventions/curlies.html (2 of 4)30/12/2004 11:02:20

The use of curly arrows to show electron movements in reaction mechanisms

The lone pair on the bromide ion moves to form a new bond between the
bromine and the right hand carbon atom. That movement is again shown
by a curly arrow. Notice again, that the curly arrow points between the
carbon and the bromine because that's where the electron pair ends up.
That leaves you with the product of this reaction, bromoethane:

Note: You can read a full description of this mechanism


together with other similar reactions of ethene and the other
alkenes by following this link.

Using curly arrows to show the movement of single


electrons
The most common use of "curly arrows" is to show the movement of
pairs of electrons. You can also use similar arrows to show the
movement of single electrons - except that the heads of these arrows
only have a single line rather than two lines.
shows the movement of an electron pair

shows the movement of a single electron


The first stage of the polymerisation of ethene, for example, could be
shown as:

http://www.chemguide.co.uk/basicorg/conventions/curlies.html (3 of 4)30/12/2004 11:02:20

The use of curly arrows to show electron movements in reaction mechanisms

You should draw the dots showing the interesting electrons. The half
arrows show where they go. This is very much a "belt-and-braces" job,
and the arrows don't add much.
Whether you choose to use these half arrows to show the movement of a
single electron should be governed by what your syllabus says. If your
syllabus encourages the use of these arrows, then it makes sense to use
them. If not - if the syllabus says that they "may" be used, or just ignores
them altogether - then they are as well avoided.
There is some danger of confusing them with the arrows showing
electron pair movements, which you will use all the time. If, by mistake,
you use an ordinary full arrow to show the movement of a single electron
you run the risk of losing marks.

Help! You must have a copy of your syllabus! If you haven't


got a copy, find out how to get a syllabus by following this link.

Where would you like to go now?


To the organic conventions menu. . .
To menu of basic organic chemistry. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/basicorg/conventions/curlies.html (4 of 4)30/12/2004 11:02:20

electrophilic addition - unsymmetrical alkenes and hydrogen halides

THE REACTION BETWEEN UNSYMMETRICAL


ALKENES AND THE HYDROGEN HALIDES

This page gives you the facts and a simple, uncluttered mechanism for
the electrophilic addition reactions between the hydrogen halides and
alkenes like propene. Hydrogen halides include hydrogen chloride and
hydrogen bromide. If you want the mechanisms explained to you in
detail, there is a link at the bottom of the page.
An unsymmetrical alkene is one like propene in which the groups or
atoms attached to either end of the carbon-carbon double bond are
different.
For example, in propene there are a hydrogen and a methyl group at one
end, but two hydrogen atoms at the other end of the double bond. But-1ene is another unsymmetrical alkene.

Electrophilic addition reactions involving hydrogen


bromide
The facts
As with all alkenes, unsymmetrical alkenes like propene react with
hydrogen bromide in the cold. The double bond breaks and a hydrogen
atom ends up attached to one of the carbons and a bromine atom to the
other.
In the case of propene, 2-bromopropane is formed.

http://www.chemguide.co.uk/mechanisms/eladd/unsymhbr.html (1 of 5)30/12/2004 11:02:26

electrophilic addition - unsymmetrical alkenes and hydrogen halides

This would normally be written in a more condensed form as

The product is 2-bromopropane.

Note: There is another possible reaction between


unsymmetrical alkenes and hydrogen bromide (but not the
other hydrogen halides) unless the hydrogen bromide and
alkene are absolutely pure.
A different mechanism happens (a free radical chain reaction
- not on UK A' level syllabuses) which leads to the hydrogen
and bromine adding the opposite way round. For A' level
purposes, you don't need to worry about that. However, if you
are interested, you will find the free radical addition
mechanism by following this link.
Use the BACK button on your browser to return to this page
later.

This is in line with Markovnikov's Rule which says:


When a compound HX is added to an unsymmetrical alkene,
the hydrogen becomes attached to the carbon with the most
hydrogens attached to it already.
In this case, the hydrogen becomes attached to the CH2 group, because
the CH2 group has more hydrogens than the CH group.
Notice that only the hydrogens directly attached to the carbon atoms at
either end of the double bond count. The ones in the CH3 group are
http://www.chemguide.co.uk/mechanisms/eladd/unsymhbr.html (2 of 5)30/12/2004 11:02:26

electrophilic addition - unsymmetrical alkenes and hydrogen halides

totally irrelevant.

Warning! Markovnikov's Rule is a useful guide for you to


work out which way round to add something across a double
bond, but it isn't the reason why things add that way. As a
general principle, don't quote Markovnikov's Rule in an exam
unless you are specifically asked for it.

The mechanism
This is an example of electrophilic addition.

The addition is this way around because the intermediate carbocation


(previously called a carbonium ion) formed is secondary. This is more
stable (and so is easier to form) than the primary carbocation which
would be produced if the hydrogen became attached to the centre
carbon atom and the bromine to the end one.

Electrophilic addition reactions involving the other


hydrogen halides
The facts
Hydrogen fluoride, hydrogen chloride and hydrogen iodide all add on in
exactly the same way as hydrogen bromide.
The only differences lie in the rates of reaction:

http://www.chemguide.co.uk/mechanisms/eladd/unsymhbr.html (3 of 5)30/12/2004 11:02:26

electrophilic addition - unsymmetrical alkenes and hydrogen halides

HF slowest reaction
HCl
HBr
HI

fastest reaction

This is because the hydrogen-halogen bond gets weaker as the halogen


atom gets bigger. If the bond is weaker, it breaks more easily and so the
reaction is faster.
If the halogen is given the symbol X, the equation for the reaction with
propene is:

Notice that the product is still in line with Markovnikov's Rule.


The mechanism
These are still examples of electrophilic addition.
Again using X to stand for any halogen:

Again, the intermediate carbocation formed is secondary. This is more


stable than the primary carbocation ion which would be formed if the
hydrogen attached to the centre carbon atom and the X to the end one. If
http://www.chemguide.co.uk/mechanisms/eladd/unsymhbr.html (4 of 5)30/12/2004 11:02:26

electrophilic addition - unsymmetrical alkenes and hydrogen halides

it is more stable it will be easier to make.

Where would you like to go now?


Help! Talk me through this mechanism . . .
To menu of electrophilic addition reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/eladd/unsymhbr.html (5 of 5)30/12/2004 11:02:26

hydrogen bromide and alkenes - the peroxide effect

HYDROGEN BROMIDE AND ALKENES: THE


PEROXIDE EFFECT

This page gives you the facts and simple uncluttered mechanisms for the
free radical addition of hydrogen bromide to alkenes - often known as the
"peroxide effect". If you want the mechanisms explained to you in more
detail, there is a link at the bottom of the page.

Addition to symmetrical alkenes


A symmetrical alkene is one like ethene where the groups at both ends
of the carbon-carbon double bond are the same.
The facts
The reaction happens at room temperature in the presence of organic
peroxides or some oxygen from the air. Alkenes react very slowly with
oxygen to produce traces of organic peroxides - so the two possible
conditions are equivalent to each other.
The reaction is a simple addition of the hydrogen bromide. For example,
with ethene:

With a symmetrical alkene you get exactly the same product in the
absence of the organic peroxides or oxygen - but the mechanism is
different.

The mechanism
Hydrogen halides (hydrogen chloride, hydrogen bromide and the rest)
usually react with alkenes using an electrophilic addition mechanism.
However, in the presence of organic peroxides, hydrogen bromide adds
http://www.chemguide.co.uk/mechanisms/freerad/alkenehbr.html (1 of 6)30/12/2004 11:02:33

hydrogen bromide and alkenes - the peroxide effect

by a different mechanism.

Note: If you are interested, you will find the electrophilic


addition mechanism for the addition of hydrogen bromide and
other hydrogen halides to alkenes if you follow this link. You
may need to explore several pages in this section.
Use the BACK button (or the HISTORY file or GO menu) on
your browser to return to this page.

With the organic peroxides present you get a free radical chain
reaction.
Chain initiation
The chain is initiated by free radicals produced by an oxygen-oxygen
bond in the organic peroxide breaking.

These free radicals extract a hydrogen atom from a hydrogen bromide


molecule to produce bromine radicals.

Chain propagation
A bromine radical joins to the ethene using one of the electrons in the pi
bond. That creates a new radical with the single electron on the other
carbon atom.

That radical reacts with another HBr molecule to produce bromoethane


and another bromine radical to continue the process.

http://www.chemguide.co.uk/mechanisms/freerad/alkenehbr.html (2 of 6)30/12/2004 11:02:33

hydrogen bromide and alkenes - the peroxide effect

etc
Chain termination
Eventually two free radicals hit each other and produce a molecule of
some sort. The process stops here because no new free radicals are
formed.

Addition to unsymmetrical alkenes


An unsymmetrical alkene is one like propene where the groups at either
end of the carbon-carbon double bond are different.
The facts
The reaction happens under the same conditions as with a symmetrical
alkene, but there is a complication because the hydrogen and the
bromine can add in two different ways. Which way they add depends on
whether there are organic peroxides (or oxygen) present or not.

Normally, when a molecule HX adds to a carbon-carbon double bond,


the hydrogen becomes attached to the carbon with the more hydrogens
on already. This is known as Markovnikov's Rule.
Because the HBr adds on the "wrong way around " in the presence of
http://www.chemguide.co.uk/mechanisms/freerad/alkenehbr.html (3 of 6)30/12/2004 11:02:33

hydrogen bromide and alkenes - the peroxide effect

organic peroxides, this is often known as the peroxide effect or antiMarkovnikov addition.
In the absence of peroxides, hydrogen bromide adds to propene via an
electrophilic addition mechanism. That gives the product predicted by
Markovnikov's Rule.

The free radical mechanism


Chain initiation
This is exactly the same as in the ethene case above.

Chain propagation
When the bromine radical joins to the propene, it attaches so that a
secondary radical is formed. This is more stable (and so easier to form)
than the primary radical which would be formed if it attached to the other
carbon atom.

That radical reacts with another HBr molecule to produce 1bromopropane and another bromine radical to continue the process.

etc
Chain termination
Eventually two free radicals hit each other and produce a molecule of
some sort. The process stops here because no new free radicals are
formed.

http://www.chemguide.co.uk/mechanisms/freerad/alkenehbr.html (4 of 6)30/12/2004 11:02:33

hydrogen bromide and alkenes - the peroxide effect

Why don't the other hydrogen halides behave in the same way?
The reason that hydrogen bromide adds in an anti-Markovnikov fashion
in the presence of organic peroxides is simply a question of reaction
rates. The free radical mechanism is much faster than the alternative
electrophilic addition mechanism.
Both mechanisms happen, but most of the product is the one from the
free radical mechanism because that is working faster.
With the other hydrogen halides, the opposite is true.
Hydrogen fluoride
The hydrogen-fluorine bond is so strong that fluorine radicals aren't
formed in the initiation step.
Hydrogen chloride
With hydrogen chloride, the second half of the propagation stage is very
slow. If you do a bond enthalpy sum, you will find that the following
reaction is endothermic.

This is due to the relatively high hydrogen-chlorine bond strength.


Hydrogen iodide
In this case, the first step of the propagation stage turns out to be
endothermic and this slows the reaction down. Not enough energy is
released when the weak carbon-iodine bond is formed.

In the case of hydrogen bromide, both steps of the propagation stage are
http://www.chemguide.co.uk/mechanisms/freerad/alkenehbr.html (5 of 6)30/12/2004 11:02:33

hydrogen bromide and alkenes - the peroxide effect

exothermic.

Where would you like to go now?


Help! Talk me through these mechanisms . . .
To menu of free radical reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/mechanisms/freerad/alkenehbr.html (6 of 6)30/12/2004 11:02:33

explaining the peroxide effect in the addition of hydrogen bromide to alkenes

EXPLAINING THE "PEROXIDE EFFECT" IN THE


REACTION BETWEEN HYDROGEN BROMIDE
AND ALKENES

This page guides you through the mechanism for the free radical addition
of hydrogen bromide to alkenes - often known as the "peroxide effect".

Note: If you just want the facts and mechanism with a


minimum of discussion you will find them by following this link.

Free radical addition to a carbon-carbon double bond


If you have read the introductory page (see above), you will know that
hydrogen bromide adds to the carbon-carbon double bond in alkenes via
a free radical mechanism in the presence of organic peroxides or oxygen
from the air.
Oxygen reacts slowly with alkenes to produce small amounts of organic
peroxides, so we don't need to look at that as a separate case.
We'll start by looking at the general case without worrying about what is
attached at either end of the carbon-carbon double bond.

The function of the organic peroxides - chain initiation


Organic peroxides are compounds containing an oxygen-oxygen single
bond, and are commonly given a general formula R-O-O-R.
The "R" groups can be quite complicated and aren't necessarily just
simple alkyl groups.
http://www.chemguide.co.uk/mechanisms/freerad/alkenehbrtt.html (1 of 8)30/12/2004 11:02:41

explaining the peroxide effect in the addition of hydrogen bromide to alkenes

The oxygen-oxygen bond is quite weak, and breaks easily so that each
oxygen gets a single electron. Free radicals are formed.

If these free radicals collide with a hydrogen bromide molecule, a


hydrogen atom is transferred, breaking the hydrogen-bromine bond to
produce bromine radicals.

Chain propagation
In any alkene (like ethene, for example), the two pairs of electrons which
make up the double bond aren't the same. One pair is held securely on
the line between the two carbon nuclei in a bond called a sigma bond.
The other pair is more loosely held in an orbital above and below the
plane of the molecule known as a pi bond.

Note: It would be helpful - but not essential - if you read


about the structure of ethene before you went on. If the
diagram above is unfamiliar to you, then you certainly ought
to read this background material.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/mechanisms/freerad/alkenehbrtt.html (2 of 8)30/12/2004 11:02:41

explaining the peroxide effect in the addition of hydrogen bromide to alkenes

Imagine what happens if a free radical approaches the pi bond in an


alkene. Once again, we'll draw it as if it is ethene - but it would apply to
any case.
The bromine radical uses one of the electrons in the pi bond to help to
form a new bond between itself and the left hand carbon atom. The other
electron returns to the right hand carbon.

Note: Don't worry that we've gone back to a simpler


diagram. It is perfectly adequate for this discussion.

The sigma bond between the carbon atoms isn't affected by any of this.
Now this new free radical reacts with a hydrogen bromide molecule. It
takes a hydrogen atom from it, leaving another bromine radical.

The bromine radical can then react with another carbon-carbon double
bond, which eventually produces a new bromine radical - and so on, and
so on . . . There is a chain reaction.

http://www.chemguide.co.uk/mechanisms/freerad/alkenehbrtt.html (3 of 8)30/12/2004 11:02:41

explaining the peroxide effect in the addition of hydrogen bromide to alkenes

Chain termination
The chain will be broken when any two radicals happen to hit each other
and form a new bond using both of the single electrons. Removing a free
radical from the system without producing a new one immediately stops
that particular chain.

What happens if the alkene is unsymmetrical?


Unsymmetrical alkenes?
An unsymmetrical alkene is one like propene, CH3CH=CH2. At one end
of the double bond there is a CH3 group and a hydrogen atom. At the
other end there are two hydrogen atoms.
A question of orientation
The problem with these unsymmetrical alkenes is that you could get two
different products depending on which end of the bond the hydrogen and
the bromine add.
In fact, under free radical conditions, most of the product is 1bromopropane.

This happens because when the bromine radical attacks the pi bond, it
joins to the carbon atom at the CH2 end of the double bond rather than
the CH end.

http://www.chemguide.co.uk/mechanisms/freerad/alkenehbrtt.html (4 of 8)30/12/2004 11:02:41

explaining the peroxide effect in the addition of hydrogen bromide to alkenes

Why does the bromine add this way?


You might think that there would be an equal chance of it attaching to
either end, but where it attaches is controlled by the stability of the free
radical formed. The more stable radical will be formed more quickly.
Think of this in terms of activation energy.

The activation energy will be lower for the reaction where the bromine
attaches to the end carbon because the radical produced is more
energetically stable.

http://www.chemguide.co.uk/mechanisms/freerad/alkenehbrtt.html (5 of 8)30/12/2004 11:02:41

explaining the peroxide effect in the addition of hydrogen bromide to alkenes

The stability of various sorts of radicals


What matters is the number of carbon atoms attached to the carbon with
the single electron.
Looking at the simplest possibilities:

Tertiary radicals are more stable than secondary ones, and secondary
radicals are more stable than primary.
In the case of a bromine radical attacking the double bond in propene, it
forms a secondary radical rather than a primary one because it is more
stable.

http://www.chemguide.co.uk/mechanisms/freerad/alkenehbrtt.html (6 of 8)30/12/2004 11:02:41

explaining the peroxide effect in the addition of hydrogen bromide to alkenes

Note: The order of stability of the various tertiary, secondary


and primary free radicals exactly reflects the order of stability
of carbocations (carbonium ions).
If you are interested in following this link, use the BACK
button on your browser to return to this page.

The rest of the reaction


Once the bromine has attached to the carbon to form the secondary
radical, there is nothing different about the rest of the reaction. The new
radical takes a hydrogen from a hydrogen bromide molecule.
This produces the 1-bromopropane and a bromine radical.

The bromine radical now goes into another cycle exactly as before to
continue the chain reaction.

Where would you like to go now?


To menu of free radical reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/mechanisms/freerad/alkenehbrtt.html (7 of 8)30/12/2004 11:02:41

explaining the peroxide effect in the addition of hydrogen bromide to alkenes

http://www.chemguide.co.uk/mechanisms/freerad/alkenehbrtt.html (8 of 8)30/12/2004 11:02:41

carbocations (or carbonium ions)

CARBOCATIONS (or CARBONIUM IONS)


All carbocations (previously known as carbonium ions) carry a positive
charge on a carbon atom. The name tells you that - a cation is a positive
ion, and the "carbo" bit refers to a carbon atom. However there are
important differences in the structures of various types of carbocations.

The different kinds of carbocations


Primary carbocations
In a primary (1) carbocation, the carbon which carries the positive
charge is only attached to one other alkyl group.

Help! An alkyl group is a group such as methyl, CH3, or


ethyl, CH3CH2. These are groups containing chains of
carbon atoms which may be branched. Alkyl groups are
given the general symbol R.

Some examples of primary carbocations include:

Notice that it doesn't matter how complicated the attached alkyl group is.
All you are doing is counting the number of bonds from the positive
carbon to other carbon atoms. In all the above cases there is only one
such link.

http://www.chemguide.co.uk/mechanisms/eladd/carbonium.html (1 of 6)30/12/2004 11:02:46

carbocations (or carbonium ions)

Using the symbol R for an alkyl group, a primary


carbocation would be written as in the box.

Secondary carbocations
In a secondary (2) carbocation, the carbon with the positive charge is
attached to two other alkyl groups, which may be the same or different.
Examples:

A secondary carbocation has the general formula


shown in the box. R and R' represent alkyl groups
which may be the same or different.

Tertiary carbocations
In a tertiary (3) carbocation, the positive carbon atom is attached to
three alkyl groups, which may be any combination of same or different.

A tertiary carbocation has the general formula


shown in the box. R, R' and R" are alkyl groups
and may be the same or different.

The stability of the various carbocations


http://www.chemguide.co.uk/mechanisms/eladd/carbonium.html (2 of 6)30/12/2004 11:02:46

carbocations (or carbonium ions)

The "electron pushing effect" of alkyl groups


You are probably familiar with the idea that bromine is more
electronegative than hydrogen, so that in a H-Br bond the electrons are
held closer to the bromine than the hydrogen. A bromine atom attached
to a carbon atom would have precisely the same effect - the electrons
being pulled towards the bromine end of the bond. The bromine has a
negative inductive effect.

Help! If you aren't familiar with all of this, follow this link to
read about electronegativity and bond polarity before you go
any further.
Use the BACK button on your browser to return to this page.

Alkyl groups do precisely the opposite and, rather than draw electrons
towards themselves, tend to "push" electrons away.

Note: The term "electron pushing" is only to help remember


what happens. The alkyl group doesn't literally "push" the
electrons away - the other end of the bond attracts them
more strongly.

http://www.chemguide.co.uk/mechanisms/eladd/carbonium.html (3 of 6)30/12/2004 11:02:46

carbocations (or carbonium ions)

This means that the alkyl group becomes slightly positive ( +) and the
carbon they are attached to becomes slightly negative ( -). The alkyl
group has a positive inductive effect.
This is sometimes shown as, for example:

The arrow shows the electrons being "pushed" away from the CH3
group. The plus sign on the left-hand end of it shows that the CH3 group
is becoming positive. The symbols + and - simply reinforce that idea.
The importance of spreading charge around in making ions stable
The general rule-of-thumb is that if a charge is very localised (all
concentrated on one atom) the ion is much less stable than if the charge
is spread out over several atoms.
Applying that to carbocations of various sorts . . .

You will see that the electron pushing effect of the CH3 group is placing
more and more negative charge on the positive carbon as you go from
primary to secondary to tertiary carbocations. The effect of this, of
course, is to cut down that positive charge.
At the same time, the region around the various CH3 groups is becoming
somewhat positive. The net effect, then, is that the positive charge is
being spread out over more and more atoms as you go from primary to
secondary to tertiary ions.
The more you can spread the charge around, the more stable the ion
becomes.

http://www.chemguide.co.uk/mechanisms/eladd/carbonium.html (4 of 6)30/12/2004 11:02:46

carbocations (or carbonium ions)

Order of stability of carbocations


primary < secondary < tertiary

Note: The symbol "<" means "is less than". So what this is
saying is that primary ions are less stable than secondary
ones which in turn are less stable than tertiary ones.

The stability of the carbocations in terms of energetics


When we talk about secondary
carbocations being more stable than
primary ones, what exactly do we mean?
We are actually talking about energetic
stability - secondary carbocations are
lower down an energy "ladder" than
primary ones.
This means that it is going to take more
energy to make a primary carbocation
than a secondary one.
If there is a choice between making a
secondary ion or a primary one, it will be much easier to make the
secondary one.
Similarly, if there is a choice between making a tertiary ion or a
secondary one, it will be easier to make the tertiary one.
This has important implications in the reactions of unsymmetrical
alkenes. If you are interested in these, follow the link below to the
electrophilic addition reactions menu.

Where would you like to go now?

http://www.chemguide.co.uk/mechanisms/eladd/carbonium.html (5 of 6)30/12/2004 11:02:46

carbocations (or carbonium ions)

To menu of electrophilic addition reactions. . .


To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/eladd/carbonium.html (6 of 6)30/12/2004 11:02:46

Understanding Chemistry - Organic Mechanisms Menu

Understanding Chemistry

ORGANIC MECHANISMS MENU


Free radical reactions . . .
Free radical substitution reactions in alkanes and alkyl groups.
Free radical addition during the polymerisation of ethene and the
reaction between HBr and alkenes in the presence of organic
peroxides.
Electrophilic addition reactions . . .
Addition reactions of alkenes such as ethene and propene.
Electrophilic substitution reactions . . .
Substitution reactions of benzene and other simple arenes.
Nucleophilic substitution reactions . . .
Substitution reactions of halogenoalkanes such as bromoethane.
Elimination reactions . . .
The formation of alkenes from halogenoalkanes such as 2bromopropane, and the dehydration of alcohols.
Nucleophilic addition reactions . . .
Addition reactions of carbonyl compounds such as ethanal and
propanone.
Nucleophilic addition / elimination reactions . . .

http://www.chemguide.co.uk/mechmenu.html (1 of 2)30/12/2004 11:02:48

Understanding Chemistry - Organic Mechanisms Menu

The reactions of acyl chlorides (acid chlorides) with water,


alcohols, ammonia and amines.

Mechanisms described elsewhere on the site


These are discussed in a section on acid catalysis in organic chemistry.
The hydration of ethene to make ethanol . . .
The esterification reaction . . .
The acid catalysed hydrolysis of esters . . .

Go to Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/mechmenu.html (2 of 2)30/12/2004 11:02:48

Free Radical Mechanisms Menu

Understanding Chemistry

FREE RADICAL MECHANISMS MENU

Free radical substitution


What is free radical substitution? . . .
A brief explanation of the terms free radical and substitution.
methane and chlorine . . .
Free radical substitution of hydrogen atoms in methane by chlorine
atoms. Includes multiple substitution if you need it.
methane and bromine . . .
Free radical substitution of hydrogen atoms in methane by
bromine atoms. Includes multiple substitution if you need it.
methylbenzene and chlorine . . .
Free radical substitution of hydrogen atoms in the methyl group in
methylbenzene by chlorine atoms in the presence of UV light
(sunlight). Includes multiple substitution if you need it.

Free radical addition


polymerisation of ethene . . .
The mechanism for the polymerisation of ethene under free radical
http://www.chemguide.co.uk/mechanisms/frmenu.html (1 of 3)30/12/2004 11:02:49

Free Radical Mechanisms Menu

conditions.
alkenes with hydrogen bromide . . .
The mechanism for the addition of hydrogen bromide to alkenes
under free radical conditions. This is often known as the "peroxide
effect".

Go to menu of other types of mechanism. . .


Go to Main Menu . . .

You might also be interested in:


properties and reactions of alkanes . . .
A survey of all the physical and chemical properties of alkanes
required by UK A level syllabuses.
properties and reactions of alkenes . . .
For more about the reactions and physical properties of alkenes like
ethene and propene.
more reactions of methylbenzene (toluene) . . .
A link to the arenes (aromatic hydrocarbons) section of the site.

Jim Clark 2000 (modified 2004)


http://www.chemguide.co.uk/mechanisms/frmenu.html (2 of 3)30/12/2004 11:02:49

Free Radical Mechanisms Menu

http://www.chemguide.co.uk/mechanisms/frmenu.html (3 of 3)30/12/2004 11:02:49

What is free radical substitution?

FREE RADICAL SUBSTITUTION

Substitution reactions
These are reactions in which one atom in a molecule is replaced by
another atom or group of atoms. Free radical substitution for A' level
purposes involves breaking a carbon-hydrogen bond in alkanes such as
methane

CH4

ethane

CH3CH3

propane

CH3CH2CH3

A new bond is then formed to something else.


It also happens in alkyl groups like methyl, ethyl (and so on) wherever
these appear in more complicated molecules.
methyl

CH3

ethyl

CH3CH2

For example, ethanoic acid is CH3COOH and contains a methyl group.


The carbon-hydrogen bonds in the methyl group behave just like those in
methane, and can be broken and replaced by something else in the
same way.
A simple example of substitution is the reaction between methane and
chlorine in the presence of UV light (or sunlight).
CH4 + Cl2

CH3Cl + HCl

Notice that one of the hydrogen atoms in the methane has been replaced
by a chlorine atom. That's substitution.

Free radical reactions

http://www.chemguide.co.uk/mechanisms/freerad/whatis.html (1 of 3)30/12/2004 11:02:52

What is free radical substitution?

Free radicals are atoms or groups of atoms which have a single unpaired
electron. A free radical substitution reaction is one involving these
radicals.
Free radicals are formed if a bond splits evenly - each atom getting one
of the two electrons. The name given to this is homolytic fission.

Note: If a bond were to split unevenly (one atom getting both


electrons, and the other none), ions would be formed. The
atom that got both electrons would become negatively
charged, while the other one would become positive. This is
called heterolytic fission.
Warning! It is important that you get these terms the right
way round. "Fission" is obvious - it just means "splitting".
"Homo" and "hetero" are used in the sense of "same" (homo)
or "different" (hetero). This is just like their use in
"homosexual" or "heterosexual".
So, homolytic fission is splitting a bond to produce two
particles which are the same in the sense that they both have
a single unpaired electron (both are free radicals). Heterolytic
fission produces two particles which are different because
one is a positive ion and the other a negative ion.

To show that a species (either an atom or a group of atoms) is a free


radical, the symbol is written with a dot attached to show the unpaired
electron. For example:
a chlorine radical
a methyl radical

http://www.chemguide.co.uk/mechanisms/freerad/whatis.html (2 of 3)30/12/2004 11:02:52

Cl
CH3

What is free radical substitution?

Note: If you wanted to be fussy, the dot showing the electron


really ought to be written next to the carbon atom in the
methyl radical, because that's the atom with the unpaired
electron - in other words as CH3. This isn't normally done
unless the radical gets more complicated. Examples of this
will crop up when you look at the mechanisms.

Where would you like to go now?


To menu of free radical reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/freerad/whatis.html (3 of 3)30/12/2004 11:02:52

Electrophilic Substitution Mechanisms Menu

Understanding Chemistry

ELECTROPHILIC SUBSTITUTION MECHANISMS


MENU
What is electrophilic substitution? . . .
Some background on benzene (including links to more detailed
discussions) and a general mechanism which covers several of
benzene's reactions.
The nitration of benzene . . .
The mechanism for the formation of nitrobenzene from benzene.
The Friedel-Crafts acylation of benzene . . .
The mechanism for the substitution of an acyl group such as
CH3CO into benzene.
The Friedel-Crafts alkylation of benzene . . .
The mechanism for the substitution of an alkyl group such as CH3
into benzene.
An industrial alkylation of benzene . . .
The mechanism for the substitution of an alkyl group such as
CH3CH2 into benzene, by a reaction involving an alkene such as
ethene.
The halogenation of benzene . . .
The mechanism for the substitution of atoms like chlorine and
bromine into benzene rings.
http://www.chemguide.co.uk/mechanisms/elsubmenu.html (1 of 2)30/12/2004 11:02:53

Electrophilic Substitution Mechanisms Menu

The sulphonation of benzene . . .


The mechanism for reaction between benzene and concentrated
sulphuric acid to produce benzenesulphonic acid
Some substitution reactions of methylbenzene . . .
Illustrates how to cope with the problem of substituting things into
rings which already have something else attached.

Go to menu of other types of mechanism. . .


Go to Main Menu . . .

You might also be interested in:


properties and reactions of arenes (aromatic hydrocarbons). . .
Covers all the physical and chemical properties of arenes like
benzene and methylbenzene required by UK A level syllabuses.

Jim Clark 2000 (modified 2004)

http://www.chemguide.co.uk/mechanisms/elsubmenu.html (2 of 2)30/12/2004 11:02:53

What is electrophilic substitution?

ELECTROPHILIC SUBSTITUTION

Background
Electrophilic substitution happens in many of the reactions of compounds
containing benzene rings - the arenes. For simplicity, we'll only look for
now at benzene itself.

Note: Before you start it would be a good idea if you had a


clear idea about the structure of benzene. Check your
syllabus now to find out what you need to know, and then
read the page on the modern orbital view of benzene in the
organic bonding section of this site. Don't forget to look in the
section(s) in your syllabus on bonding as well as organic
chemistry.
Haven't got a syllabus? Follow this link to find out how to get
one.

This is what you need to understand for the purposes of the electrophilic
substitution mechanisms:

Benzene, C6H6, is a planar molecule containing a ring of six


carbon atoms each with a hydrogen atom attached.
There are delocalised electrons above and below the plane of the
ring.

http://www.chemguide.co.uk/mechanisms/elsub/whatis.html (1 of 6)30/12/2004 11:02:58

What is electrophilic substitution?

The presence of the delocalised electrons makes benzene


particularly stable.
Benzene resists addition reactions because that would involve
breaking the delocalisation and losing that stability.
Benzene is represented by this symbol, where the
circle represents the delocalised electrons, and
each corner of the hexagon has a carbon atom
with a hydrogen attached.

Electrophilic substitution reactions involving positive ions


Benzene and electrophiles
Because of the delocalised electrons exposed above and below the
plane of the rest of the molecule, benzene is obviously going to be highly
attractive to electrophiles - species which seek after electron rich areas in
other molecules.

Species: A useful word which can mean any particle you


want it to mean - an atom, a molecule, an ion or a free radical.

The electrophile will either be a positive ion, or the slightly positive end of
a polar molecule.

Help! If you aren't sure what a polar molecule is, read about
electronegativity and polar bonds before you go on.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/mechanisms/elsub/whatis.html (2 of 6)30/12/2004 11:02:58

What is electrophilic substitution?

The delocalised electrons above and below the plane of the benzene
molecule are open to attack in the same way as those above and below
the plane of an ethene molecule. However, the end result will be different.

If benzene underwent addition reactions in the same way as ethene, it


would need to use some of the delocalised electrons to form bonds with
the new atoms or groups. This would break the delocalisation - and this
costs energy.

Note: You can read about electrophilic addition to ethene if


you are interested.
Use the BACK button on your browser to return to this page.

Instead, it can maintain the delocalisation if it replaces a hydrogen atom


by something else - a substitution reaction. The hydrogen atoms aren't
involved in any way with the delocalised electrons.
In most of benzene's reactions, the electrophile is a positive ion, and
these reactions all follow a general pattern.
The general mechanism
The first stage
Suppose the electrophile is a positive ion X+.
Two of the electrons in the delocalised system are attracted towards the
http://www.chemguide.co.uk/mechanisms/elsub/whatis.html (3 of 6)30/12/2004 11:02:58

What is electrophilic substitution?

X+ and form a bond with it. This has the effect of breaking the
delocalisation, although not completely.

Note: If you aren't sure about the use of curly arrows in


mechanisms, you must follow this link before you go on.
Use the BACK button on your browser to return to this page.

The ion formed in this step isn't the final product. It immediately goes on
to react with something else. It is just an intermediate.
There is still delocalisation in the intermediate formed, but
it only covers part of the ion. When you write one of these
mechanisms, draw the partial delocalisation to take in all
the carbon atoms apart from the one that the X has
become attached to.
The intermediate ion carries a positive charge because
you are joining together a neutral molecule and a positive
ion. This positive charge is spread over the delocalised part of the ring.
Simply draw the "+" in the middle of the ring.
The hydrogen at the top isn't new - it's the hydrogen that was already
attached to that carbon. We need to show that it is there for the next
stage.
The second stage

http://www.chemguide.co.uk/mechanisms/elsub/whatis.html (4 of 6)30/12/2004 11:02:58

What is electrophilic substitution?

Here we've introduced a new ion, Y-. Where did this come from? You
have to remember that it is impossible to get a positive ion on its own in a
chemical system - so Y- is simply the negative ion that was originally
associated with X+. Don't worry about this at the moment - it's much
easier to see when you've got a real example in front of you.
A lone pair of electrons on Y- forms a bond with
the hydrogen atom at the top of the ring. That
means that the pair of electrons joining the
hydrogen onto the ring aren't needed any more.
These then move down to plug the gap in the
delocalised electrons, so restoring the delocalised
ring of electrons which originally gave the benzene its special stability.
The energetics of the reaction
The complete delocalisation is temporarily broken as X replaces H on the
ring, and this costs energy. However, that energy is recovered when the
delocalisation is re-established. This initial input of energy is simply the
activation energy for the reaction. In this case, it is going to be high
(something around 150 kJ mol-1), and this means that benzene's
reactions tend to be slow.

Electrophilic substitution reactions not involving positive


ions
Halogenation and sulphonation
In these reactions, the electrophiles are polar molecules rather than fully
positive ions.
Because these mechanisms are different from what's gone before (and

http://www.chemguide.co.uk/mechanisms/elsub/whatis.html (5 of 6)30/12/2004 11:02:58

What is electrophilic substitution?

from each other), there isn't any point in dealing with them in a general
way. You will find them explained in full if you follow the link to the
electrophilic substitution menu below.

Where would you like to go now?


To menu of electrophilic substitution reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/elsub/whatis.html (6 of 6)30/12/2004 11:02:58

bonding in benzene - sp2 hybridisation and delocalisation

BONDING IN BENZENE

Important! This article builds on knowledge about the


bonding in methane, and the bonding in ethene.
You will find the current page much easier to understand if
you read these other ones first.
You may also find it useful to read the article on orbitals if you
aren't sure about simple orbital theory.
You can also read about the evidence which leads to the
structure described in this article. That page includes the
Kekul structure for benzene and the reasons that it isn't very
satisfactory.

An orbital model for the benzene structure


Building the orbital model
Benzene is built from hydrogen atoms (1s1) and carbon atoms
(1s22s22px12py1).
Each carbon atom has to join to three other atoms (one hydrogen and
two carbons) and doesn't have enough unpaired electrons to form the
required number of bonds, so it needs to promote one of the 2s2 pair into
the empty 2pz orbital. This is exactly the same as happens whenever
carbon forms bonds - whatever else it ends up joined to.

http://www.chemguide.co.uk/basicorg/bonding/benzene2.html (1 of 5)30/12/2004 11:03:01

bonding in benzene - sp2 hybridisation and delocalisation

Important! If you have any doubts about this then you


should follow the links at the top of the page.

Because each carbon is only joining to three other atoms, when the
carbon atoms hybridise their outer orbitals before forming bonds, they
only need to hybridise three of the orbitals rather than all four. They use
the 2s electron and two of the 2p electrons, but leave the other 2p
electron unchanged.
The new orbitals formed are called sp2 hybrids, because they are made
by an s orbital and two p orbitals reorganising themselves. The three sp2
hybrid orbitals arrange themselves as far apart as possible - which is at
120 to each other in a plane. The remaining p orbital is at right angles to
them.
Each carbon atom now looks like the
diagram on the right. This is all exactly
the same as happens in ethene.
The difference in benzene is that each
carbon atom is joined to two other similar
carbon atoms instead of just one. Each
carbon atom uses the sp2 hybrids to
form sigma bonds with two other
carbons and one hydrogen atom. The next diagram shows the sigma
bonds formed, but for the moment leaves the p orbitals alone.

Remember: A sigma bond is formed by the end-to-end


overlap between atomic orbitals.

http://www.chemguide.co.uk/basicorg/bonding/benzene2.html (2 of 5)30/12/2004 11:03:01

bonding in benzene - sp2 hybridisation and delocalisation

Only a part of the ring is shown because the diagram gets extremely
cluttered if you try to draw any more.
Notice that the p electron on each carbon atom is overlapping with those
on both sides of it. This extensive sideways overlap produces a system
of pi bonds which are spread out over the whole carbon ring. Because
the electrons are no longer held between just two carbon atoms, but are
spread over the whole ring, the electrons are said to be delocalised. The
six delocalised electrons go into three molecular orbitals - two in each.

Remember: A molecular orbital is the region of space which


contains a bonding pair of electrons.
Warning! Be very careful how you phrase this in exams.
You must never talk about the p orbitals on the carbons
overlapping sideways to produce a delocalised pi bond. This
upsets examiners because a pi bond can only hold 2
electrons - whereas in benzene there are 6 delocalised
electrons. Talk instead about a "pi system" - or just about the
delocalised electrons.

http://www.chemguide.co.uk/basicorg/bonding/benzene2.html (3 of 5)30/12/2004 11:03:01

bonding in benzene - sp2 hybridisation and delocalisation

In common with the great majority of descriptions of the bonding in


benzene, we are only going to show one of these delocalised molecular
orbitals for simplicity.
In the diagram, the sigma bonds
have been shown as simple lines to
make the diagram less confusing.
The two rings above and below the
plane of the molecule represent
one molecular orbital. The two
delocalised electrons can be found
anywhere within those rings. The
other four delocalised electrons live in two similar (but not identical)
molecular orbitals.
Relating the orbital model to the properties of benzene

Note: To get the best out of this section you ought to read
the article on the Kekul structure for benzene.

The shape of benzene


This is easily explained. Benzene is a regular hexagon because all the
bonds are identical. The delocalisation of the electrons means that there
aren't alternating double and single bonds.
The energetic stability of benzene
This is accounted for by the delocalisation. As a general principle, the
more you can spread electrons around - in other words, the more they
are delocalised - the more stable the molecule becomes. The extra
stability of benzene is often referred to as "delocalisation energy".
The reluctance of benzene to undergo addition reactions
With the delocalised electrons in place, benzene is about 150 kJ mol-1
more stable than it would otherwise be. If you added other atoms to a
benzene ring you would have to use some of the delocalised electrons to
http://www.chemguide.co.uk/basicorg/bonding/benzene2.html (4 of 5)30/12/2004 11:03:01

bonding in benzene - sp2 hybridisation and delocalisation

join the new atoms to the ring. That would disrupt the delocalisation and
the system would become less stable.
Since about 150 kJ per mole of benzene would have to be supplied to
break up the delocalisation, this isn't going to be an easy thing to do.
The symbol for benzene
Although you will still come across the Kekul structure
for benzene, for most purposes we use the structure on
the right.
The hexagon shows the ring of six carbon atoms, each
of which has one hydrogen attached. (You have to know
that - counting bonds to find out how many hydrogens to add doesn't
work in this particular case.)
The circle represents the delocalised electrons. It is essential that you
include the circle. If you miss it out, you are drawing cyclohexane and not
benzene.

Where would you like to go now?


To read about the Kekul structure for benzene. . .
To the organic bonding menu. . .
To menu of basic organic chemistry. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/basicorg/bonding/benzene2.html (5 of 5)30/12/2004 11:03:01

electrophilic substitution - the nitration of benzene

THE NITRATION OF BENZENE

This page gives you the facts and a simple, uncluttered mechanism for
the electrophilic substitution reaction between benzene and a mixture of
concentrated nitric acid and concentrated sulphuric acid. If you want the
nitration mechanism explained to you in detail, there is a link at the
bottom of the page.

The electrophilic substitution reaction between benzene


and nitric acid
The facts
Benzene is treated with a mixture of concentrated nitric acid and
concentrated sulphuric acid at a temperature not exceeding 50C. As
temperature increases there is a greater chance of getting more than one
nitro group, -NO2, substituted onto the ring.
Nitrobenzene is formed.

or:

The concentrated sulphuric acid is acting as a catalyst.

The formation of the electrophile


The electrophile is the "nitronium ion" or the "nitryl cation", NO2+. This is
http://www.chemguide.co.uk/mechanisms/elsub/nitration.html (1 of 3)30/12/2004 11:03:04

electrophilic substitution - the nitration of benzene

formed by reaction between the nitric acid and the sulphuric acid.

The electrophilic substitution mechanism


Stage one

Stage two

Where would you like to go now?


Help! Talk me through this mechanism . . .
To menu of electrophilic substitution reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

http://www.chemguide.co.uk/mechanisms/elsub/nitration.html (2 of 3)30/12/2004 11:03:04

electrophilic substitution - the nitration of benzene

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/elsub/nitration.html (3 of 3)30/12/2004 11:03:04

Explaining the nitration of benzene - electrophilic substitution

EXPLAINING THE NITRATION OF BENZENE

This page guides you through the mechanism for the nitration of benzene
involving an electrophilic substitution reaction between benzene and
nitric acid.

Important! It would help if you first read the page What is


electrophilic substitution? before you went on.

The electrophilic substitution reaction between benzene


and nitric acid
The formation of the electrophile
If you are going to substitute an -NO2 group into the ring, then the
electrophile must be NO2+. This is called the "nitronium ion" or the "nitryl
cation", and is formed by reaction between the nitric acid and sulphuric
acid.

Note: If you don't understand why the electrophile has got to


be NO2+, then you really should look at What is electrophilic
substitution? before you go on.
If you are going to substitute X onto the ring, then the
electrophile must be X+. If you are going to insert an -NO2
group onto the ring, then the electrophile must be NO2+.

http://www.chemguide.co.uk/mechanisms/elsub/nitrationtt.html (1 of 4)30/12/2004 11:03:07

Explaining the nitration of benzene - electrophilic substitution

The equation

Important! Check past exam papers to see whether you will


need to quote this equation in an exam. It's probable that you
will, so learn it! There are ways of building it up, but they
involve much more effort than learning it in the first place.

The hydrogensulphate ion, HSO4-, will also be involved in the


mechanism. The hydroxonium ion, H3O+, isn't involved.

Note: The hydroxonium ion (also called the hydronium ion or


oxonium ion) is simply a hydrogen ion attached to a water
molecule - what is often written more simply as H+(aq).

The electrophilic substitution mechanism


Stage one
As the NO2+ ion approaches the delocalised electrons in the benzene,
those electrons are strongly attracted towards the positive charge.
Two electrons from the delocalised system are used to form a new bond
with the NO2+ ion. Because those two electrons aren't a part of the
delocalised system any longer, the delocalisation is partly broken, and in
the process the ring gains a positive charge.

http://www.chemguide.co.uk/mechanisms/elsub/nitrationtt.html (2 of 4)30/12/2004 11:03:07

Explaining the nitration of benzene - electrophilic substitution

The hydrogen shown on the ring is the one which was already attached
to that top carbon atom - it's nothing new or subtle! We need to show it
there because it has to be removed in the second stage.
Stage two
The second stage involves a hydrogensulphate ion,
HSO4-, which was produced at the same time as the
NO2+ ion (refer back to the equation showing the
formation of the electrophile if you've forgotten).

Note: Only one of the lone pairs in the hydrogensulphate ion


is shown. There are lots more, but those aren't involved in the
reaction.

This removes a hydrogen from the ring to form sulphuric acid - the
catalyst has therefore been regenerated. The electrons which originally
joined the hydrogen to the ring are now used to re-establish the
delocalised system.
Stage two - a sloppy way of writing the same thing!
You will often find the second stage of this reaction simplified in many (or
even most!) books. The second stage is shown as:

http://www.chemguide.co.uk/mechanisms/elsub/nitrationtt.html (3 of 4)30/12/2004 11:03:07

Explaining the nitration of benzene - electrophilic substitution

The hydrogen is shown as "falling off" the ring as a hydrogen ion. This is
sloppy and unsatisfactory on two counts:

Hydrogen ions never exist on their own in a chemical reaction. A


hydrogen ion is a raw proton - the most intensely positive thing you
can imagine. It will always be attached to something else.
By not showing the hydrogensulphate ion, you can't show that the
sulphuric acid catalyst has been regenerated. That's simply
unsatisfying!

Showing exactly how the hydrogen is removed from the ring isn't difficult
- do it properly!

Where would you like to go now?


To menu of electrophilic substitution reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/elsub/nitrationtt.html (4 of 4)30/12/2004 11:03:07

electrophilic substitution - the acylation of benzene

THE FRIEDEL-CRAFTS ACYLATION OF


BENZENE

This page gives you the facts and a simple, uncluttered mechanism for
the electrophilic substitution reaction between benzene and ethanoyl
chloride in the presence of an aluminium chloride catalyst. If you want
the Friedel-Crafts acylation mechanism explained to you in detail, there
is a link at the bottom of the page.

The electrophilic substitution reaction between benzene


and ethanoyl chloride
What is acylation?
An acyl group is an alkyl group attached to a carbonoxygen double bond. If "R" represents any alkyl group,
then an acyl group has the formula RCO-. Acylation
means substituting an acyl group into something - in this
case, into a benzene ring.
The most commonly used acyl group is CH3CO-. This is called the
ethanoyl group. In the example which follows we are substituting a
CH3CO- group into the ring, but you could equally well use any other
alkyl group instead of the CH3.
The facts
The most reactive substance containing an acyl group is an acyl chloride
(also known as an acid chloride). These have the general formula RCOCl.
Benzene is treated with a mixture of ethanoyl chloride, CH3COCl, and
aluminium chloride as the catalyst. A ketone called phenylethanone is
formed.

http://www.chemguide.co.uk/mechanisms/elsub/fcacyl.html (1 of 3)30/12/2004 11:03:11

electrophilic substitution - the acylation of benzene

Note: Ketones: A family of compounds containing a carbonoxygen double bond with a hydrocarbon group either side of
it. In this case there is a methyl group on one side and a
benzene ring on the other.
Don't worry too much about the name "phenylethanone" - all
that matters is that you can draw the structure.

or better:

The aluminium chloride isn't written into these equations because it is


acting as a catalyst. If you wanted to include it, you could write AlCl3 over
the top of the arrow.

The formation of the electrophile


The electrophile is CH3CO+. It is formed by
reaction between the ethanoyl chloride and the
aluminium chloride catalyst.

The electrophilic substitution mechanism


Stage one
http://www.chemguide.co.uk/mechanisms/elsub/fcacyl.html (2 of 3)30/12/2004 11:03:11

electrophilic substitution - the acylation of benzene

Stage two

The hydrogen is removed by the AlCl4- ion which was formed at the
same time as the CH3CO+ electrophile. The aluminium chloride catalyst
is re-generated in this second stage.

Where would you like to go now?


Help! Talk me through this mechanism . . .
To menu of electrophilic substitution reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/elsub/fcacyl.html (3 of 3)30/12/2004 11:03:11

Explaining the Friedel-Crafts acylation of benzene - electrophilic substitution

EXPLAINING THE FRIEDEL-CRAFTS


ACYLATION OF BENZENE

This page guides you through the mechanism for the Friedel-Crafts
acyation of benzene involving an electrophilic substitution reaction
between benzene and ethanoyl chloride in the presence of an aluminium
chloride catalyst.

Important! It would help if you first read the page What is


electrophilic substitution? before you went on.

The electrophilic substitution reaction between benzene


and ethanoyl chloride
The formation of the electrophile
If you are going to replace a hydrogen atom in a
benzene ring by CH3CO, then the electrophile
must be the ion CH3CO+. The positive charge must be on the carbon
atom, because that's what gets attached to the ring.

Note: If you don't understand why the electrophile has got to


be CH3CO+, then you really should look at What is
electrophilic substitution? before you go on.
If you are going to substitute X onto the ring, then the
electrophile must be X+. If you are going to insert an CH3CO
group onto the ring, then the electrophile must be CH3CO+.

http://www.chemguide.co.uk/mechanisms/elsub/fcacyltt.html (1 of 4)30/12/2004 11:03:14

Explaining the Friedel-Crafts acylation of benzene - electrophilic substitution

Aluminium chloride, AlCl3, is an electron deficient molecule. It is


covalently bonded, but because the aluminium is only forming 3 bonds,
and has no lone pairs, there are only 6 electrons around the aluminium
atom rather than 8. It takes a chlorine (as a chloride ion) from the
ethanoyl chloride, and forms a co-ordinate (dative covalent) bond with it.

Help! A co-ordinate bond is a covalent bond in which both


electrons originally came from the same atom. In this case,
both the electrons in the bond come from the chlorine being
removed from the ethanoyl chloride.

The equation simplified

The electrophilic substitution mechanism


Stage one
As the CH3CO+ ion approaches the delocalised electrons in the
benzene, those electrons are strongly attracted towards the positive
charge.
Two electrons from the delocalised system are used to form a new bond
with the CH3CO+ ion. Because those two electrons aren't a part of the
delocalised system any longer, the delocalisation is partly broken, and in
the process the ring gains a positive charge.

http://www.chemguide.co.uk/mechanisms/elsub/fcacyltt.html (2 of 4)30/12/2004 11:03:14

Explaining the Friedel-Crafts acylation of benzene - electrophilic substitution

Help! Don't be confused that we have reversed the way we


have written the CH3CO+ ion to show it as +COCH3.
This is just so that it is easier to draw the mechanism tidily in
the same way that the general mechanism was drawn on the
page What is electrophilic substitution?

The hydrogen shown on the ring is the one which was already attached
to that top carbon atom. We need to show it there because it has to be
removed in the second stage.
Stage two
The second stage involves the AlCl4-, which was produced at the same
time as the CH3CO+ ion.

One of the aluminium-chlorine bonds breaks and both electrons from it


are used to join to the hydrogen. This removes the hydrogen from the
ring to form HCl, and re-generates the aluminium chloride catalyst in the
process. The electrons which originally joined the hydrogen to the ring
are now used to re-establish the delocalised system.

http://www.chemguide.co.uk/mechanisms/elsub/fcacyltt.html (3 of 4)30/12/2004 11:03:14

Explaining the Friedel-Crafts acylation of benzene - electrophilic substitution

Where would you like to go now?


To menu of electrophilic substitution reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/elsub/fcacyltt.html (4 of 4)30/12/2004 11:03:14

electrophilic substitution - the alkylation of benzene

THE FRIEDEL-CRAFTS ALKYLATION OF


BENZENE

This page gives you the facts and a simple, uncluttered mechanism for
the electrophilic substitution reaction between benzene and
chloromethane in the presence of an aluminium chloride catalyst. Any
other chloroalkane would work similarly. If you want the Friedel-Crafts
alkylation mechanism explained to you in detail, there is a link at the
bottom of the page.

The electrophilic substitution reaction between benzene


and chloromethane
What is alkylation?
Alkylation means substituting an alkyl group into something - in this case
into a benzene ring. A hydrogen on the ring is replaced by a group like
methyl or ethyl and so on.
The facts
Benzene is treated with a chloroalkane (for example, chloromethane or
chloroethane) in the presence of aluminium chloride as a catalyst. On
this page, we will look at substituting a methyl group, but any other alkyl
group could be used in the same way.
Substituting a methyl group gives methylbenzene - once known as
toluene.

or better:

http://www.chemguide.co.uk/mechanisms/elsub/fcalkyl.html (1 of 3)30/12/2004 11:03:19

electrophilic substitution - the alkylation of benzene

The aluminium chloride isn't written into these equations because it is


acting as a catalyst. If you wanted to include it, you could write AlCl3 over
the top of the arrow.

Note: The methylbenzene formed is more reactive than the


original benzene, and so the reaction doesn't stop there. You
get further methyl groups substituted around the ring.
You won't have to worry about this for A' level.

The formation of the electrophile


The electrophile is CH3+. It is formed by reaction between the
chloromethane and the aluminium chloride catalyst.

The electrophilic substitution mechanism


Stage one

Stage two

http://www.chemguide.co.uk/mechanisms/elsub/fcalkyl.html (2 of 3)30/12/2004 11:03:19

electrophilic substitution - the alkylation of benzene

The hydrogen is removed by the AlCl4- ion which was formed at the
same time as the CH3+ electrophile. The aluminium chloride catalyst is regenerated in this second stage.

Where would you like to go now?


Help! Talk me through this mechanism . . .
To menu of electrophilic substitution reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/elsub/fcalkyl.html (3 of 3)30/12/2004 11:03:19

Explaining the Friedel-Crafts alkylation of benzene - electrophilic substitution

EXPLAINING THE FRIEDEL-CRAFTS


ALKYLATION OF BENZENE

This page guides you through the mechanism for the Friedel-Crafts
alkyation of benzene involving an electrophilic substitution reaction
between benzene and a chloroalkane like chloromethane in the presence
of an aluminium chloride catalyst.

Important! It would help if you first read the page What is


electrophilic substitution? before you went on.

The electrophilic substitution reaction between benzene


and chloromethane
The formation of the electrophile
If you are going to replace a hydrogen atom in a benzene ring by CH3,
then the electrophile must be the ion CH3+.

Note: If you don't understand why the electrophile has got to


be CH3+, then you really should look at What is electrophilic
substitution? before you go on.
If you are going to substitute X onto the ring, then the
electrophile must be X+. If you are going to insert an CH3
group onto the ring, then the electrophile must be CH3+.

http://www.chemguide.co.uk/mechanisms/elsub/fcalkyltt.html (1 of 4)30/12/2004 11:03:21

Explaining the Friedel-Crafts alkylation of benzene - electrophilic substitution

Aluminium chloride, AlCl3, is an electron deficient molecule. It is


covalently bonded, but because the aluminium is only forming 3 bonds,
and has no lone pairs, there are only 6 electrons around the aluminium
atom rather than 8. It takes a chlorine (as a chloride ion) from the
chloromethane, and forms a co-ordinate (dative covalent) bond with it.

Help! A co-ordinate bond is a covalent bond in which both


electrons originally came from the same atom. In this case,
both the electrons in the bond come from the chlorine being
removed from the chloromethane.

The equation simplified

The electrophilic substitution mechanism


Stage one
As the CH3+ ion approaches the delocalised electrons in the benzene,
those electrons are strongly attracted towards the positive charge.
Two electrons from the delocalised system are used to form a new bond
with the CH3+ ion. Because those two electrons aren't a part of the
delocalised system any longer, the delocalisation is partly broken, and in
the process the ring gains a positive charge.

http://www.chemguide.co.uk/mechanisms/elsub/fcalkyltt.html (2 of 4)30/12/2004 11:03:21

Explaining the Friedel-Crafts alkylation of benzene - electrophilic substitution

The hydrogen shown on the ring is the one which was already attached
to that top carbon atom. We need to show it there because it has to be
removed in the second stage.
Stage two
The second stage involves the AlCl4-, which was produced at the same
time as the CH3+ ion.

One of the aluminium-chlorine bonds breaks and both electrons from it


are used to join to the hydrogen. This removes the hydrogen from the
ring to form HCl, and re-generates the aluminium chloride catalyst in the
process. The electrons which originally joined the hydrogen to the ring
are now used to re-establish the delocalised system.

Where would you like to go now?


To menu of electrophilic substitution reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

http://www.chemguide.co.uk/mechanisms/elsub/fcalkyltt.html (3 of 4)30/12/2004 11:03:21

Explaining the Friedel-Crafts alkylation of benzene - electrophilic substitution

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/elsub/fcalkyltt.html (4 of 4)30/12/2004 11:03:21

electrophilic substitution - an industrial alkylation of benzene

AN INDUSTRIAL ALKYLATION OF BENZENE

This page gives you the facts and simple, uncluttered mechanisms for
the electrophilic substitution reaction between benzene and alkenes in
the presence of a mixture of aluminium chloride and hydrogen chloride
as the catalyst. If you want these mechanisms explained to you in detail,
there is a link at the bottom of the page.

Note: The process described below is one of several ways


of making alkylbenzenes like ethylbenzene from benzene
and alkenes, using a variety of different catalysts and
conditions.
The process to produce ethylbenzene is currently asked for
by the UK A level Exam Board AQA, and I am including it
solely to satisfy their requirements. If you don't actually need
to read this, don't!

The electrophilic substitution reaction between benzene


and ethene
The facts
Industrially, alkyl groups can be substituted into a benzene ring using a
variant on Friedel-Crafts alkylation. One possibility is that instead of
using a chloroalkane with an aluminium chloride catalyst, they use an
alkene and a mixture of aluminium chloride and hydrogen chloride as the
catalyst.
This is a cheaper method because it saves having to make the
chloroalkane first.

http://www.chemguide.co.uk/mechanisms/elsub/indalkyl.html (1 of 5)30/12/2004 11:03:27

electrophilic substitution - an industrial alkylation of benzene

Note: If you haven't already done so, you might like to look
at the Friedel-Crafts alkylation reaction before you go on.

To put an ethyl group on the ring (to make ethylbenzene), benzene is


treated with a mixture of ethene, HCl and aluminium chloride.

or better:

The aluminium chloride and HCl aren't written into these equations
because they are acting as catalysts. If you wanted to include them, you
could write AlCl3 and HCl over the top of the arrow.
The formation of the electrophile
The electrophile is CH3CH2+. It is formed by reaction between the ethene
and the HCl - exactly as if you were beginning to add the HCl to the
ethene.

The chloride ion is immediately picked up by the aluminium chloride to


form an AlCl4- ion. That prevents the chloride ion from reacting with the
CH3CH2+ ion to form chloroethane.

http://www.chemguide.co.uk/mechanisms/elsub/indalkyl.html (2 of 5)30/12/2004 11:03:27

electrophilic substitution - an industrial alkylation of benzene

Note: It wouldn't matter if it did react, because chloroethane


will react with benzene using a simple Friedel-Crafts
alkylation reaction to give the product you want anyway.

The electrophilic substitution mechanism


Stage one

Stage two

The hydrogen is removed by the AlCl4- ion which was formed at the
same time as the CH3CH2+ electrophile. The aluminium chloride and
hydrogen chloride catalysts are re-generated in this second stage.

The electrophilic substitution reaction between benzene


and propene
The facts
The problem with more complicated alkenes like propene is that you
have to be careful about the structure of the product. In each case, you
can only really be sure of that structure if you work through the
mechanism first.
http://www.chemguide.co.uk/mechanisms/elsub/indalkyl.html (3 of 5)30/12/2004 11:03:27

electrophilic substitution - an industrial alkylation of benzene

For example, the propyl group becomes attached to the ring via its
middle carbon atom - and not its end one.

You still need a mixture of aluminium chloride and hydrogen chloride as


catalysts.
The formation of the electrophile
When the propene reacts with the HCl, the hydrogen becomes attached
to the end carbon atom. A secondary carbocation (carbonium ion) is
formed because it is more stable than the primary one which would have
been formed if the addition was the other way round.

Because the positive charge is on the centre carbon atom, that is the one
which will become attached to the ring.
The electrophilic substitution mechanism
Stage one

Stage two
http://www.chemguide.co.uk/mechanisms/elsub/indalkyl.html (4 of 5)30/12/2004 11:03:27

electrophilic substitution - an industrial alkylation of benzene

Again, the hydrogen is removed by the AlCl4- ion. The aluminium


chloride and hydrogen chloride catalysts are re-generated in this second
stage.

Where would you like to go now?


Help! Talk me through these mechanisms . . .
To menu of electrophilic substitution reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000 (slightly modified 2004)

http://www.chemguide.co.uk/mechanisms/elsub/indalkyl.html (5 of 5)30/12/2004 11:03:27

Explaining the industrial alkylation of benzene - electrophilic substitution

EXPLAINING THE INDUSTRIAL ALKYLATION OF


BENZENE

This page guides you through the mechanisms for the electrophilic
substitution reaction between benzene and alkenes in the presence of a
mixture of aluminium chloride and hydrogen chloride as catalysts.

Important! There is some quite complicated chemistry


involved on this page. Be sure that you actually need to know
about it!
Although these mechanisms weren't actually required by the
syllabus, in the past one Exam Board (NEAB - now part of
AQA) frequently asked guided questions about them. Check
recent exam papers to see if that's the case with your current
Exam Board. If you haven't already got any, find out how to
get copies of recent past papers by accessing your Exam
Board's web site (links on the syllabuses page).

The electrophilic substitution reaction between benzene


and ethene
The formation of the electrophile
If you are going to replace a hydrogen atom in a benzene ring by
CH3CH2, then the electrophile must be the ion CH3CH2+.

http://www.chemguide.co.uk/mechanisms/elsub/indalkyltt.html (1 of 7)30/12/2004 11:03:31

Explaining the industrial alkylation of benzene - electrophilic substitution

Note: If you don't understand why the electrophile has got to


be CH3CH2+, then you should look at What is electrophilic
substitution? before you go on.
If you are going to substitute X onto the ring, then the
electrophile must be X+. If you are going to insert an CH3CH2
group onto the ring, then the electrophile must be CH3CH2+.

The electrophile is formed by reaction between the ethene and the HCl exactly as if you were beginning to add the HCl to the ethene.

Note: If you haven't done any alkene chemistry recently, it


would be worth your while looking at the reaction of ethene
with hydrogen halides, if you aren't sure about it.

The chloride ion is immediately picked up by the aluminium chloride to


form an AlCl4- ion. That prevents the chloride ion from reacting with the
CH3CH2+ ion to form chloroethane.
Aluminium chloride is an electron deficient molecule, with only 3 pairs of
electrons around the aluminium. By forming a fourth bond with the
chloride ion, it becomes more energetically stable.

http://www.chemguide.co.uk/mechanisms/elsub/indalkyltt.html (2 of 7)30/12/2004 11:03:31

Explaining the industrial alkylation of benzene - electrophilic substitution

Note: We are only showing one of the lone pairs around the
chloride ion. The other three aren't involved in this reaction.

The electrophilic substitution mechanism

Note: From now on the mechanism is exactly the same as


the Friedel-Crafts alkylation of benzene. You might like to
compare the two.

Stage one
As the CH3CH2+ ion approaches the delocalised electrons in the
benzene, those electrons are strongly attracted towards the positive
charge.
Two electrons from the delocalised system are used to form a new bond
with the positive carbon atom. Because those two electrons aren't a part
of the delocalised system any longer, the delocalisation is partly broken,
and in the process the ring gains a positive charge.

The hydrogen shown on the ring is the one which was already attached
to that top carbon atom. We need to show it there because it has to be
removed in the second stage.
Stage two
The second stage involves the AlCl4-.

http://www.chemguide.co.uk/mechanisms/elsub/indalkyltt.html (3 of 7)30/12/2004 11:03:31

Explaining the industrial alkylation of benzene - electrophilic substitution

One of the aluminium-chlorine bonds breaks and both electrons from it


are used to join to the hydrogen. Removing the hydrogen from the ring reforms the HCl and the aluminium chloride catalysts in the process. The
electrons which originally joined the hydrogen to the ring are now used to
re-establish the delocalised system.

The electrophilic substitution reaction between benzene


and propene
Important! Don't even think about reading this section until
you are sure that you understand everything on this page so
far! The next bit adds another layer of difficulty.

The problem lies in how you attach the carbon chain to the ring. You
might remember that the propene joins on via the middle carbon atom.

The formation of the electrophile


When the propene reacts with the HCl, there are two different ways that
the initial stage of the addition reaction could happen.
The pi bond could swing so that the hydrogen atom becomes attached to
the left-hand carbon. This produces a primary carbocation (carbonium
ion).

http://www.chemguide.co.uk/mechanisms/elsub/indalkyltt.html (4 of 7)30/12/2004 11:03:31

Explaining the industrial alkylation of benzene - electrophilic substitution

Alternatively, the pi bond electrons could move the other way and form a
secondary carbocation.

Note: If you don't understand about primary and secondary


carbocations (carbonium ions) follow this link before you go
any further.
You might also find it useful to read about the addition of
hydrogen halides to unsymmetrical alkenes as well.

The secondary carbocation (with the positive charge on the centre


carbon atom) is the more stable, and so that forms more readily.
Because the electrophile has the positive charge on the centre carbon,
that will be where it attaches to the benzene ring.
The electrophilic substitution mechanism
The mechanism is now exactly the same as the one involving ethene that
we've already looked at in detail.
Stage one
Two electrons from the delocalised system are used to form a new bond
with the positive carbon atom. Because those two electrons aren't a part
of the delocalised system any longer, the delocalisation is partly broken,
http://www.chemguide.co.uk/mechanisms/elsub/indalkyltt.html (5 of 7)30/12/2004 11:03:31

Explaining the industrial alkylation of benzene - electrophilic substitution

and in the process the ring gains a positive charge.

Stage two
In the second stage, one of the aluminium-chlorine bonds breaks and
both electrons from it are used to join to the hydrogen. Removing the
hydrogen from the ring re-forms the HCl and the aluminium chloride
catalysts in the process. The electrons which originally joined the
hydrogen to the ring are now used to re-establish the delocalised system.

Where would you like to go now?


To menu of electrophilic substitution reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/elsub/indalkyltt.html (6 of 7)30/12/2004 11:03:31

Explaining the industrial alkylation of benzene - electrophilic substitution

http://www.chemguide.co.uk/mechanisms/elsub/indalkyltt.html (7 of 7)30/12/2004 11:03:31

Explaining electrophilic addition involving hydrogen halides and unsymmetrical alkenes

EXPLAINING THE REACTION BETWEEN


UNSYMMETRICAL ALKENES AND THE
HYDROGEN HALIDES

This page guides you through the mechanism for the electrophilic
addition of hydrogen halides such as hydrogen bromide to unsymmetrical
alkenes like propene.

Important! To make sense of this page, you will need to


understand about the structure and stability of carbocations
(previously called carbonium ions) and be confident about
electrophilic addition to simple alkenes like ethene.
If you aren't sure about either of these things, follow these
links first.
You would also find it easier if you first read about the
electrophilic addition reactions between hydrogen halides
and symmetrical alkenes like ethene, and addition to
unsymmetrical alkenes in general. If you have just come from
those, ignore these links!

Electrophilic addition reactions involving hydrogen


bromide
If you want the mechanism for one of the other hydrogen halides, simply
replace Br by whatever else you are interested in - F or Cl or I. There is
no difference whatsoever in the mechanisms.
You might, however, need to be aware that there is an alternative
mechanism involving hydrogen bromide and alkenes if the reaction
mixture is impure in the presence of organic peroxides or oxygen from
the air.

http://www.chemguide.co.uk/mechanisms/eladd/unsymhbrtt.html (1 of 6)30/12/2004 11:03:36

Explaining electrophilic addition involving hydrogen halides and unsymmetrical alkenes

Note: The different mechanism (a free radical chain reaction


- not on UK A' level syllabuses) leads to the hydrogen and
bromine adding the opposite way round. For A' level
purposes, you don't need to worry about that. However, if you
are interested, you will find the free radical addition
mechanism by following this link.
Use the BACK button on your browser to return to this page
later.

Hydrogen bromide as an electrophile


Hydrogen bromide is chosen as a typical
hydrogen halide. Bromine is more
electronegative than hydrogen. That means
that the bonding pair of electrons is pulled towards the bromine end of
the bond, and so the hydrogen bromide molecule is polar.

Note: If you aren't sure about electronegativity and bond


polarity follow this link before you read on.
Use the BACK button on your browser to return to this page.

The slightly positive hydrogen atom will be attracted to negative regions


in other molecules, and is therefore an electrophile.

http://www.chemguide.co.uk/mechanisms/eladd/unsymhbrtt.html (2 of 6)30/12/2004 11:03:36

Explaining electrophilic addition involving hydrogen halides and unsymmetrical alkenes

Electrophile: A substance with a strong attraction to a


negative region in another substance. Electrophiles are either
fully positive ions, or the slightly positive end of a polar
molecule.

The reaction of propene with hydrogen bromide


The double bond in all alkenes is made up of two different parts. One
pair of electrons lies on the line between the two nuclei where you would
expect them to be. This is called a sigma bond.
The other pair lies in an orbital above and below the plane of the rest of
the molecule, and is called a pi bond. The pi bond is weaker than a
sigma bond and is very vulnerable to attack.

Note: If this isn't fairly obvious to you, you really ought to


read the page introducing electrophilic addition before you go
on.
Use the BACK button on your browser to return to this page.

As the HBr approaches the pi bond, the electrons in that bond are
attracted towards the slightly positive hydrogen atom. That repels the
electrons in the hydrogen-bromine bond down towards the bromine.

http://www.chemguide.co.uk/mechanisms/eladd/unsymhbrtt.html (3 of 6)30/12/2004 11:03:36

Explaining electrophilic addition involving hydrogen halides and unsymmetrical alkenes

The electron movements continue until a new bond is made between one
of the carbon atoms and the hydrogen. The bromine now has both
electrons from the H-Br bond, and so is negatively charged as a bromide
ion.
The problem is that there are two possible ways that the pi bond
electrons could move.
They could form a bond between the hydrogen and the left-hand carbon:

or they could form a bond with the right-hand one:

It's the second of these changes that happens more readily. In that case,
a secondary carbocation is formed - and that's more energetically stable
than the primary one formed in the first possibility.
http://www.chemguide.co.uk/mechanisms/eladd/unsymhbrtt.html (4 of 6)30/12/2004 11:03:36

Explaining electrophilic addition involving hydrogen halides and unsymmetrical alkenes

Because the secondary ion is more energetically stable, it will form more
easily and so the reaction needs less activation energy.

Important! If you don't understand about the structure and


stability of carbocations (carbonium ions) follow this link.
Activation energy: The minimum energy needed before a
reaction will occur. In this case it is the energy needed to
break the various bonds and make the carbocation and the
bromide ion.

Once the ions have been formed, the lone pair on the bromide ion is
strongly attracted towards the positive carbon atom. It moves towards it
and forms a bond.

Note: There are actually 4 lone pairs around the bromide


ion, but we are only interested in the one shown.

That leaves you with the over-all mechanism:

http://www.chemguide.co.uk/mechanisms/eladd/unsymhbrtt.html (5 of 6)30/12/2004 11:03:36

Explaining electrophilic addition involving hydrogen halides and unsymmetrical alkenes

Important! If you've had problems with this page you might


find it useful to read about addition to unsymmetrical alkenes
in general, and then come back here again.

Where would you like to go now?


To menu of electrophilic addition reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/eladd/unsymhbrtt.html (6 of 6)30/12/2004 11:03:36

electrophilic addition to unsymmetrical alkenes

ELECTROPHILIC ADDITION TO
UNSYMMETRICAL ALKENES

Important! To make sense of this page, you will need to


understand about the structure and stability of carbocations
(previously called carbonium ions) and be confident about
electrophilic addition to simple alkenes like ethene.
If you aren't sure about either of these things, follow these
links first. Trying to build on shaky foundations risks confusing
you and undermining your confidence.

The Problem
The addition of H-X to an unsymmetrical alkene like propene
An unsymmetrical alkene is one like propene or but-1-ene in which the
groups or atoms attached to either end of the carbon-carbon double bond
are different.
For example, in propene there are a hydrogen and a methyl group at one
end, but two hydrogen atoms at the other end of the double bond.

With these unsymmetrical alkenes, it is possible to get two different


products during some addition reactions.
During the addition of a molecule HX to propene, you could in principle
get either this reaction:

http://www.chemguide.co.uk/mechanisms/eladd/unsymprob.html (1 of 7)30/12/2004 11:03:42

electrophilic addition to unsymmetrical alkenes

or this one:

It depends on which way around you add the HX across the double bond.
In fact, in most cases, it's mainly the second reaction which happens.
The hydrogen atom becomes attached to the right-hand carbon atom as
we've drawn it.
Markovnikov's Rule

Note: Don't worry too much about the spelling of


Markovnikov - there are nearly as many versions as there are
text books. It's an English attempt at a Russian name and
Markovnikov wouldn't actually have recognised any of them!

When a compound HX is added to an unsymmetrical alkene, the


hydrogen becomes attached to the carbon with the most hydrogens
attached to it already.
Remember that the HX has to attach itself to the carbon atoms which
were originally part of the double bond. So in this case, adding HX to
CH3CH=CH2, the hydrogen is attached to the CH2 group, because the
CH2 group has more hydrogens than the CH group.
Notice that only the hydrogens directly attached count. The ones in the
CH3 group are totally irrelevant.

http://www.chemguide.co.uk/mechanisms/eladd/unsymprob.html (2 of 7)30/12/2004 11:03:42

electrophilic addition to unsymmetrical alkenes

Warning! Markovnikov's Rule is a useful guide for you to


work out which way round to add something across a double
bond, but it isn't the reason why things add that way. Propene
has never even heard of Markovnikov! When the question
arises in an exam, you will need a much more fundamental
explanation which is coming up next. As a general principle,
don't quote Markovnikov's Rule in an exam unless you are
specifically asked for it.

The Mechanism
HX as an electrophile
In each of the cases we are interested in, X is
more electronegative than hydrogen. That
means that the bonding pair of electrons will be
dragged towards the X end of the bond, and so
the hydrogen becomes slightly positively
charged.
The slightly positive hydrogen is the electrophile, and is attracted to the pi
bond in the propene.
What happens next determines which way around the HX adds across
the double bond.
The two possible mechanisms
In both possibilities, the pi bond breaks and the electron pair swings
down to form a bond with the hydrogen atom. At the same time, the
electrons in the H-X bond are repelled right down on to the X to give an
X- ion.
The difference lies in which way the electron pair in the pi bond swings.
First possibility
The electron pair moves to form a bond between the hydrogen and the
left-hand carbon.
http://www.chemguide.co.uk/mechanisms/eladd/unsymprob.html (3 of 7)30/12/2004 11:03:42

electrophilic addition to unsymmetrical alkenes

When the second stage of the mechanism happens, and the lone pair on
X- forms a bond with the positive carbon atom, the product of this
mechanism is not the one which Markovnikov's Rule predicts.
Second possibility
The electron pair moves to form a bond between the hydrogen and the
right-hand carbon.

This time, the overall mechanism leads to the correct product.


Why does one of these work better than the other?
That's the truth of the situation! One of these mechanisms works better
than the other one. The second mechanism works much faster than the
first, and so most of the product that you get is CH3-CHX-CH3.
There will be small amounts of CH3-CH2-CH2X despite what
Markovnikov says!
The difference between the two mechanisms lies in the intermediates the things formed at the half-way stage.
In the mechanism that works best, you get a secondary carbocation
formed as one of the intermediate ions.

http://www.chemguide.co.uk/mechanisms/eladd/unsymprob.html (4 of 7)30/12/2004 11:03:42

electrophilic addition to unsymmetrical alkenes

In the slow mechanism which produces hardly any product, you get a
primary carbocation formed instead.

It is much easier to form the secondary carbocation because it is more


energetically stable. The activation energy for the reaction will be less,
and so most of the reaction happens via that mechanism.

Activation energy: The minimum energy needed before a


reaction will occur. In this case it is the energy needed to
break the various bonds and make the carbocation and the Xion.

How to attack this sort of question in an exam


Suppose you were asked for the mechanism for the addition of HX to
some alkene you hadn't come across before.
First, you need to decide whether the alkene is symmetrical or not. If it is
symmetrical, there's no problem - it wouldn't matter which way around
you added the HX. If it is unsymmetrical, you need to decide which way
round the HX is going to add.
For example, supposed you were asked for the mechanism for the
addition of HX to but-1-ene, CH3-CH2-CH=CH2.
First, use Markovnikov's Rule to decide which carbon to attach the
hydrogen to. In this case, the hydrogen would get attached to the CH2
end of the double bond, because that carbon has more hydrogens than
the CH end.

http://www.chemguide.co.uk/mechanisms/eladd/unsymprob.html (5 of 7)30/12/2004 11:03:42

electrophilic addition to unsymmetrical alkenes

Warning! Markovnikov's Rule is only to help you decide.


Don't give the examiners any hint that you are using it unless they specifically ask.

Now write the mechanism, taking care to draw the curly arrow showing
the movement of the pi bond so that the hydrogen gets attached to that
particular CH2 carbon.

If you are asked why the HX adds this way round, look at the carbocation
formed as an intermediate and decide whether it is secondary or tertiary.
Here it is a secondary ion. Then think about what sort of ion would be
formed if the HX added the other way around. In this case that would be
a primary ion.

Then say something like:


"The secondary carbocation formed in this reaction is more energetically
stable than the primary one which would be formed if the addition was
the other way round, and so less activation energy is needed."

http://www.chemguide.co.uk/mechanisms/eladd/unsymprob.html (6 of 7)30/12/2004 11:03:42

electrophilic addition to unsymmetrical alkenes

Important! If there were bits of this that you found you


couldn't understand, you are probably trying to do too much
too quickly.
Follow the links to carbocations (carbonium ions) and
electrophilic addition to simple alkenes, get the basics sorted
out properly, and then try again.

Where would you like to go now?


To menu of electrophilic addition reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/eladd/unsymprob.html (7 of 7)30/12/2004 11:03:42

electrophilic substitution - the halogenation of benzene

THE HALOGENATION OF BENZENE

This page gives you the facts and a simple, uncluttered mechanism for
the electrophilic substitution reaction between benzene and chlorine or
bromine in the presence of a catalyst such as aluminium chloride or iron.
If you want this mechanism explained to you in detail, there is a link at
the bottom of the page.

The electrophilic substitution reaction between benzene


and chlorine or bromine
The facts
Benzene reacts with chlorine or bromine in an electrophilic substitution
reaction, but only in the presence of a catalyst. The catalyst is either
aluminium chloride (or aluminium bromide if you are reacting benzene
with bromine) or iron.
Strictly speaking iron isn't a catalyst, because it gets permanently
changed during the reaction. It reacts with some of the chlorine or
bromine to form iron(III) chloride, FeCl3, or iron(III) bromide, FeBr3.

These compounds act as the catalyst and behave exactly like aluminium
chloride in these reactions.
The reaction with chlorine
The reaction between benzene and chlorine in the presence of either
aluminium chloride or iron gives chlorobenzene.

http://www.chemguide.co.uk/mechanisms/elsub/halogenation.html (1 of 4)30/12/2004 11:03:48

electrophilic substitution - the halogenation of benzene

or:

The reaction with bromine


The reaction between benzene and bromine in the presence of either
aluminium bromide or iron gives bromobenzene. Iron is usually used
because it is cheaper and more readily available.

or:

The formation of the electrophile


We are going to explore the reaction using chlorine and aluminium
chloride. If you want one of the other combinations, all you have to do is
to replace each Cl by Br, or each Al by Fe.
As a chlorine molecule approaches the benzene ring, the delocalised
electrons in the ring repel electrons in the chlorine-chlorine bond.

It is the slightly positive end of the chlorine molecule which acts as the
electrophile. The presence of the aluminium chloride helps this
http://www.chemguide.co.uk/mechanisms/elsub/halogenation.html (2 of 4)30/12/2004 11:03:48

electrophilic substitution - the halogenation of benzene

polarisation.
The electrophilic substitution mechanism
Stage one

Stage two

The hydrogen is removed by the AlCl4- ion which was formed in the first
stage. The aluminium chloride catalyst is re-generated in this second
stage.

Where would you like to go now?


Help! Talk me through this mechanism . . .
To menu of electrophilic substitution reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

http://www.chemguide.co.uk/mechanisms/elsub/halogenation.html (3 of 4)30/12/2004 11:03:48

electrophilic substitution - the halogenation of benzene

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/elsub/halogenation.html (4 of 4)30/12/2004 11:03:48

Explaining the halogenation of benzene - electrophilic substitution

EXPLAINING THE HALOGENATION OF BENZENE

This page guides you through the mechanism for the electrophilic
substitution reaction between benzene and chlorine in the presence of an
aluminium chloride or an iron catalyst.
The reaction involving bromine is exactly the same, except that iron
would be the preferred catalyst. Aluminium bromide could be used as an
alternative.
In what follows, if you want one of the other combinations, all you have to
do is to replace each Cl by Br, or each Al by Fe.

Note: The reason that iron functions in the same way as the
aluminium compounds is explained in the "facts" section of
the introductory page on halogenation of benzene. If you've
forgotten, you might like to look back at that before you go on.

The electrophilic substitution reaction between benzene


and chlorine
The formation of the electrophile
Many of the electrophilic substitution reactions of benzene involve an
attack on the benzene by a positive ion. In the chlorine case, forming a Cl
+ ion needs too much energy.
As the chlorine molecule approaches a benzene ring, the delocalised
electrons in the ring repel the electrons in the chlorine-chlorine bond.
That induces a dipole in the chlorine.

http://www.chemguide.co.uk/mechanisms/elsub/halogenationtt.html (1 of 4)30/12/2004 11:03:52

Explaining the halogenation of benzene - electrophilic substitution

Note: If you aren't happy about the structure of benzene,


you could follow this link.
The formation of the induced dipole is much the same as
happens in the addition of bromine to ethene. If you aren't
sure about induced dipoles, you might like to have a look at
the beginning of that page.

Also nearby is the aluminium chloride, and this encourages the


polarisation of the chlorine. The aluminium chloride is an electron
deficient molecule, with the aluminium only having 3 pairs of electrons in
its bonding level. The aluminium is strongly attracted to the slightly
negative end of the chlorine molecule, and pulls electrons even more
towards that end.
The electrophilic substitution mechanism
Stage one
Two electrons from the delocalised system are used to form a new bond
with the slightly positive chlorine atom. Because those two electrons
aren't a part of the delocalised system any longer, the delocalisation is
partly broken, and in the process the ring gains a positive charge.

http://www.chemguide.co.uk/mechanisms/elsub/halogenationtt.html (2 of 4)30/12/2004 11:03:52

Explaining the halogenation of benzene - electrophilic substitution

The hydrogen shown on the ring is the one which was already attached
to that top carbon atom. We need to show it there because it has to be
removed in the second stage.
Notice that the chlorine-chlorine bond breaks, transferring a chloride ion
to the AlCl3 to make an AlCl4- ion.
Stage two
The second stage involves that AlCl4-.

One of the aluminium-chlorine bonds breaks and both electrons from it


are used to join to the hydrogen. Removing the hydrogen from the ring
forms the HCl which is also produced in the reaction, and the aluminium
chloride catalyst is re-generated. The electrons which originally joined
the hydrogen to the ring are now used to re-establish the delocalised
system.

Where would you like to go now?


To menu of electrophilic substitution reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .
http://www.chemguide.co.uk/mechanisms/elsub/halogenationtt.html (3 of 4)30/12/2004 11:03:52

Explaining the halogenation of benzene - electrophilic substitution

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/elsub/halogenationtt.html (4 of 4)30/12/2004 11:03:52

Explaining electrophilic addition involving bromine and symmetrical alkenes

EXPLAINING THE REACTION BETWEEN


SYMMETRICAL ALKENES AND BROMINE

This page guides you through the mechanism for the electrophilic
addition of bromine to symmetrical alkenes like ethene or cyclohexene.
Unsymmetrical alkenes are covered separately, and you will find a link at
the bottom of the page.

The electrophilic addition of bromine to ethene


The structure of ethene
The structure of ethene is shown in the
diagram on the right. The pi bond is an
orbital above and below the plane of the rest
of the molecule, and relatively exposed to
things around it.

Note: If you aren't sure about this, then you should read the
page What is electrophilic addition? before you go on.
Use the BACK button on your browser to return to this page.

Bromine as an electrophile
Since two identical bromine atoms are joined together in the bromine
molecule there is no reason why one atom should pull the bonding pair of
electrons towards itself - they must be equally electronegative and so
there won't be any separation of charge, + or -. How, then, can
bromine be an electrophile?

http://www.chemguide.co.uk/mechanisms/eladd/symbr2tt.html (1 of 8)30/12/2004 11:03:59

Explaining electrophilic addition involving bromine and symmetrical alkenes

Note: If you aren't sure about electronegativity and bond


polarity follow this link before you read on.
Equally, if you aren't sure about terms like electrophile, then it
really would be a good idea to read the page What is
electrophilic addition? before you go on.
Use the BACK button on your browser to return to this page.

In fact, bromine is a very polarisable molecule - in other words, the


electrons in the bond are very easily pushed to one end or the other. As
the bromine molecule approaches the ethene, the electrons in the pi
bond tend to repel the electrons in the bromine-bromine bond, leaving
the nearer bromine slightly positive and the further one slightly negative.

The bromine molecule therefore acquires an induced dipole which is


automatically lined up the right way round for a successful attack on the
ethene.

http://www.chemguide.co.uk/mechanisms/eladd/symbr2tt.html (2 of 8)30/12/2004 11:03:59

Explaining electrophilic addition involving bromine and symmetrical alkenes

Help! What is an "induced dipole"? A dipole is simply a


separation of charge between + at one end and - at the
other. "Induced" means that it has been created by some
external influence (in this case the approach of the pi bond),
and didn't already exist.
Where it does already exist - as, for example, in HBr - it is
called a permanent dipole.

The simplified version of the mechanism

Note: Use this version unless your examiners insist on the


more accurate one.
If you've come into this web site from a search engine directly
to this page, read the notes on the introductory page to this
reaction before you go any further.

The electrons from the pi bond move down towards the slightly positive
bromine atom.

In the process, the electrons in the Br-Br bond are repelled down until
they are entirely on the bottom bromine atom, producing a bromide ion.

http://www.chemguide.co.uk/mechanisms/eladd/symbr2tt.html (3 of 8)30/12/2004 11:03:59

Explaining electrophilic addition involving bromine and symmetrical alkenes

Help! If you aren't sure about the use of curly arrows in


mechanisms, you must follow this link before you go on.
Use the BACK button on your browser to return to this page.

The ion with a positive charge on the carbon atom is called a


carbocation or carbonium ion.
Why is there a positive charge on the carbon atom? The pi bond was
originally made up of an electron from each of the carbon atoms. Both of
those electrons have been used to make a new bond to the bromine.
That leaves the right-hand carbon an electron short - hence positively
charged.
In the second stage of the mechanism, the lone pair of electrons on the
bromide ion is strongly attracted to the positive carbon and moves
towards it until a bond is formed.

The overall mechanism is therefore

The more accurate version of the mechanism

http://www.chemguide.co.uk/mechanisms/eladd/symbr2tt.html (4 of 8)30/12/2004 11:03:59

Explaining electrophilic addition involving bromine and symmetrical alkenes

Note: Don't learn this unless you have to. There is a real risk
of getting confused. If your examiners are happy to accept
the simple version, there's no point in making life difficult for
yourself.

The reaction starts off just the same as in the simplified version, with the
pi bond electrons moving down towards the slightly positive bromine
atom.
But this time, the top bromine atom becomes attached to both carbon
atoms, with the positive charge being found on the bromine rather than
on one of the carbons. A bromonium ion is formed.

The bromonium ion is then attacked from the back by a bromide ion
formed in a nearby reaction. It can't be attacked by its original bromide
ion because the bromonium ion is completely cluttered up with a positive
bromine on that side.

It doesn't matter which of the carbon atoms the bromide ion attacks - the
end result would be just the same.

http://www.chemguide.co.uk/mechanisms/eladd/symbr2tt.html (5 of 8)30/12/2004 11:03:59

Explaining electrophilic addition involving bromine and symmetrical alkenes

Note: You can't really draw this mechanism tidily in one line
because the bromide ion has to be in a different place at the
beginning of the second stage than it was at the end of the
first stage.

The electrophilic addition of bromine to cyclohexene


The simplified version of the mechanism

Note: Use this version unless your examiners insist on the


more accurate one.

The electrons from the pi bond move towards the slightly positive
bromine atom.

In the process, the electrons in the bromine-bromine bond are repelled


until they are entirely on the right-hand bromine atom, producing a
bromide ion.
Exactly as with ethene, a carbocation is formed. The bottom carbon atom
lost one of its electrons when the pi bond swung towards the bromine.
In the second stage of the mechanism, the lone pair of electrons on the
bromide ion is strongly attracted to the positive carbon and moves
towards it until a bond is formed.

http://www.chemguide.co.uk/mechanisms/eladd/symbr2tt.html (6 of 8)30/12/2004 11:03:59

Explaining electrophilic addition involving bromine and symmetrical alkenes

The overall mechanism is therefore

The alternative version of the mechanism

Note: Don't learn this unless your examiners insist on it.


Keep life simple!

The reaction starts off just the same as in the simplified version, with the
pi bond electrons moving towards the slightly positive bromine atom.
But this time, the left-hand bromine atom becomes attached to both
carbon atoms, with the positive charge being found on the bromine rather
than on one of the carbons. A bromonium ion is formed.

The bromonium ion is then attacked from the back by a bromide ion
formed in a nearby reaction. It can't be attacked by its original bromide
ion because approach from that side is hindered by the positive bromine
atom.

http://www.chemguide.co.uk/mechanisms/eladd/symbr2tt.html (7 of 8)30/12/2004 11:03:59

Explaining electrophilic addition involving bromine and symmetrical alkenes

It doesn't matter which of the carbon atoms on either end of the original
double bond the bromide ion attacks - the end result would be just the
same.

Note: Once again, you can't really draw this mechanism


tidily in one line because of the need to move the bromide ion.

Where would you like to go now?


Look at the same reactions involving unsymmetrical
alkenes . . .
To menu of electrophilic addition reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/eladd/symbr2tt.html (8 of 8)30/12/2004 11:03:59

electrophilic addition - symmetrical alkenes and bromine

THE REACTION BETWEEN SYMMETRICAL


ALKENES AND BROMINE

This page gives you the facts and a simple, uncluttered mechanism for
the electrophilic addition reactions between bromine (and the other
halogens) and alkenes like ethene and cyclohexene. If you want the
mechanisms explained to you in detail, there is a link at the bottom of the
page.

The electrophilic addition of bromine to ethene


The facts
Alkenes react in the cold with pure liquid bromine, or with a solution of
bromine in an organic solvent like tetrachloromethane. The double bond
breaks, and a bromine atom becomes attached to each carbon. The
bromine loses its original red-brown colour to give a colourless liquid. In
the case of the reaction with ethene, 1,2-dibromoethane is formed.

This decolourisation of bromine is often used as a test for a carboncarbon double bond. If an aqueous solution of bromine is used ("bromine
water"), you get a mixture of products. The presence of the water
complicates the mechanism beyond what is required by current A'level
syllabuses.
The other halogens, apart from fluorine, behave similarly. (Fluorine
reacts explosively with all hydrocarbons - including alkenes - to give
carbon and hydrogen fluoride.)
If you are interested in the reaction with, say, chlorine, all you have to do
is to replace Br by Cl in all the equations on this page.

http://www.chemguide.co.uk/mechanisms/eladd/symbr2.html (1 of 6)30/12/2004 11:04:04

electrophilic addition - symmetrical alkenes and bromine

The mechanism for the reaction between ethene and bromine


The reaction is an example of electrophilic addition.

Warning! There are two versions of the ethene / bromine


mechanism in common use, and you must know which your
examiners will accept.
One version is simplified to bring it into line with the other
alkene electrophilic addition mechanisms. You will probably
find that your examiners will accept this one, but you must
find out to be sure.
You almost certainly won't be able to tell this from your
syllabus. You need to refer to recent mark schemes, or to
any support material that your Exam Board provides. If you
still aren't sure, contact your Exam Board direct. You can find
out how to do this by using the link to your Board's web site
on the syllabuses page.
The person you need to contact will probably have the title
Subject Officer for Chemistry or something similar. Ask
whether they want the mechanism for the reaction between
bromine and alkenes which proceeds via a carbocation or via
a bromonium ion intermediate.

Bromine as an electrophile
The bromine is a very "polarisable" molecule and the approaching pi
bond in the ethene induces a dipole in the bromine molecule. If you draw
this mechanism in an exam, write the words "induced dipole" next to the
bromine molecule - to show that you understand what's going on.
The simplified version of the mechanism

http://www.chemguide.co.uk/mechanisms/eladd/symbr2.html (2 of 6)30/12/2004 11:04:04

electrophilic addition - symmetrical alkenes and bromine

Note: Use this version unless your examiners insist on the


more accurate one.

The more accurate version of the mechanism

Note: Don't learn this unless you have to. There is a real risk
of getting confused. If your examiners are happy to accept
the simple version, there's no point in making life difficult for
yourself.

In the first stage of the reaction, one of the bromine atoms becomes
attached to both carbon atoms, with the positive charge being found on
the bromine atom. A bromonium ion is formed.

The bromonium ion is then attacked from the back by a bromide ion
formed in a nearby reaction.

http://www.chemguide.co.uk/mechanisms/eladd/symbr2.html (3 of 6)30/12/2004 11:04:04

electrophilic addition - symmetrical alkenes and bromine

The electrophilic addition of bromine to cyclohexene


The facts
Cyclohexene reacts with bromine in the same way and under the same
conditions as any other alkene. 1,2-dibromocyclohexane is formed.

The mechanism for the reaction between cyclohexene and bromine


The reaction is an example of electrophilic addition.

Warning! Again, there are two versions of this mechanism in


common use, and you must know which your examiners will
accept.

Bromine as an electrophile
Again, the bromine is polarised by the approaching pi bond in the
cyclohexene. Don't forget to write the words "induced dipole" next to the
bromine molecule.
The simplified version of the mechanism

http://www.chemguide.co.uk/mechanisms/eladd/symbr2.html (4 of 6)30/12/2004 11:04:04

electrophilic addition - symmetrical alkenes and bromine

Note: Use this version unless your examiners insist on the


more accurate one.

The alternative version of the mechanism

Note: Don't learn this unless you have to. If your examiners
are happy to accept the simple version, there's no point in
making life difficult for yourself.

In the first stage of the reaction, one of the bromine atoms becomes
attached to both carbon atoms, with the positive charge being found on
the bromine atom. A bromonium ion is formed.

The bromonium ion is then attacked from the back by a bromide ion
formed in a nearby reaction.

http://www.chemguide.co.uk/mechanisms/eladd/symbr2.html (5 of 6)30/12/2004 11:04:04

electrophilic addition - symmetrical alkenes and bromine

Where would you like to go now?


Help! Talk me through these mechanisms . . .
Look at the same reactions involving unsymmetrical
alkenes . . .
To menu of electrophilic addition reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/eladd/symbr2.html (6 of 6)30/12/2004 11:04:04

electrophilic addition - unsymmetrical alkenes and bromine

THE REACTION BETWEEN UNSYMMETRICAL


ALKENES AND BROMINE

Important! This page assumes that you have already read


the page on the addition of bromine to symmetrical alkenes. If
you haven't, you must read it before you go on. It contains
important advice that you will need to make best use of this
page.

This page gives you the facts and a simple, uncluttered mechanism for
the electrophilic addition reactions between bromine and alkenes like
propene. If you want the mechanism explained to you in detail, there is a
link at the bottom of the page.
An unsymmetrical alkene is one like propene in which the groups or
atoms attached to either end of the carbon-carbon double bond are
different.
For example, in propene there are a hydrogen and a methyl group at one
end, but two hydrogen atoms at the other end of the double bond. But-1ene is another unsymmetrical alkene.

The electrophilic addition of bromine to propene


The facts
In common with all other alkenes, propene reacts in the cold with pure
liquid bromine, or with a solution of bromine in an organic solvent like
tetrachloromethane. The double bond breaks, and a bromine atom
becomes attached to each carbon. The bromine loses its original redhttp://www.chemguide.co.uk/mechanisms/eladd/unsymbr2.html (1 of 4)30/12/2004 11:04:08

electrophilic addition - unsymmetrical alkenes and bromine

brown colour to give a colourless liquid. In the case of the reaction with
propene, 1,2-dibromopropane is formed.

The other halogens, apart from fluorine, behave similarly. (Fluorine


reacts explosively with all hydrocarbons - including alkenes - to give
carbon and hydrogen fluoride.)
If you are interested in the reaction with, say, chlorine, all you have to do
is to replace Br by Cl in all the equations on this page.
The mechanism for the reaction between propene and bromine
The reaction is an example of electrophilic addition.

Warning! Just as with symmetrical alkenes, there are two


versions of the propene / bromine mechanism in common
use, and you must know which your examiners will accept.
How you can find out which one your examiners expect is
explained on the page on the addition of bromine to
symmetrical alkenes.

Bromine as an electrophile
The bromine is a very "polarisable" molecule and the approaching pi
bond in the propene induces a dipole in the bromine molecule. If you
draw this mechanism in an exam, write the words "induced dipole" next
to the bromine molecule.
The simplified version of the mechanism

http://www.chemguide.co.uk/mechanisms/eladd/unsymbr2.html (2 of 4)30/12/2004 11:04:08

electrophilic addition - unsymmetrical alkenes and bromine

Note: Use this version unless your examiners insist on the


more accurate one.

The more accurate version of the mechanism

Note: Don't learn this unless your examiners insist on it.


There's no point in making life difficult for yourself.

In the first stage of the reaction, one of the bromine atoms becomes
attached to both carbon atoms, with the positive charge being found on
the bromine atom. A bromonium ion is formed.

The bromonium ion is then attacked from the back by a bromide ion
formed in a nearby reaction.

http://www.chemguide.co.uk/mechanisms/eladd/unsymbr2.html (3 of 4)30/12/2004 11:04:08

electrophilic addition - unsymmetrical alkenes and bromine

Where would you like to go now?


Help! Talk me through these mechanisms . . .
To menu of electrophilic addition reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/eladd/unsymbr2.html (4 of 4)30/12/2004 11:04:08

Explaining electrophilic addition involving bromine and unsymmetrical alkenes

EXPLAINING THE REACTION BETWEEN


UNSYMMETRICAL ALKENES AND BROMINE

This page guides you through the mechanism for the electrophilic
addition of bromine to unsymmetrical alkenes like propene.

Important! You will find it easier to make sense of this page


if you first read about the electrophilic addition reactions
between bromine and symmetrical alkenes like ethene, and
addition to unsymmetrical alkenes in general. You may want
to follow other links from those pages as well before you
come back here again.

The electrophilic addition of bromine to propene


The attraction between the propene and the bromine
The double bond in all alkenes is made up of two different parts. One pair
of electrons lies on the line between the two nuclei where you would
expect them to be. This is called a sigma bond.
The other pair lies in an orbital above and below the plane of the rest of
the molecule, and is called a pi bond. The pi bond is weaker than a
sigma bond and is very vulnerable to attack.

Note: If this isn't fairly obvious to you, you really should


follow the links at the top of the page before you go on - and
perhaps explore other simpler reactions from the electrophilic
addition menu as well.

http://www.chemguide.co.uk/mechanisms/eladd/unsymbr2tt.html (1 of 6)30/12/2004 11:04:10

Explaining electrophilic addition involving bromine and unsymmetrical alkenes

As the bromine molecule approaches the pi bond, the electrons in that


bond repel the electrons in the bromine-bromine bond down towards the
bottom bromine. That produces an induced dipole in the bromine
molecule.

Help! What is an "induced dipole"? A dipole is simply a


separation of charge between + at one end and - at the
other. "Induced" means that it has been created by some
external influence (in this case the approach of the pi bond),
and didn't already exist.
Where it does already exist - as, for example, in HBr - it is
called a permanent dipole.

The simplified version of the mechanism

http://www.chemguide.co.uk/mechanisms/eladd/unsymbr2tt.html (2 of 6)30/12/2004 11:04:10

Explaining electrophilic addition involving bromine and unsymmetrical alkenes

Note: Use this version unless your examiners insist on the


more accurate one.
If you've come into this web site from a search engine directly
to this page, read the notes on the addition of bromine to
ethene before you go any further.
Use the BACK button on your browser to return to this page.

The electrons from the pi bond move down towards the slightly positive
bromine atom.

In the process, the electrons in the Br-Br bond are repelled down until
they are entirely on the bottom bromine atom, producing a bromide ion.
Notice the way that the pi bond electrons have moved. By swinging so
that the bromine is attached to the right-hand carbon, a secondary
carbocation has been formed. That is more stable than the primary one
which would have been formed if the pi electrons had swung the other
way.

Note: If this doesn't make sense to you, read about


carbocations (previously called carbonium ions) and addition
to unsymmetrical alkenes in general.

http://www.chemguide.co.uk/mechanisms/eladd/unsymbr2tt.html (3 of 6)30/12/2004 11:04:10

Explaining electrophilic addition involving bromine and unsymmetrical alkenes

In the second stage of the mechanism, the lone pair of electrons on the
bromide ion is strongly attracted to the positive carbon and moves
towards it until a bond is formed.

The overall mechanism is therefore

The more accurate version of the mechanism

Note: Don't learn this unless you have to. There is a real risk
of getting confused. If your examiners are happy to accept
the simple version, there's no point in making life difficult for
yourself.

http://www.chemguide.co.uk/mechanisms/eladd/unsymbr2tt.html (4 of 6)30/12/2004 11:04:10

Explaining electrophilic addition involving bromine and unsymmetrical alkenes

The reaction starts off just the same as in the simplified version, with the
pi bond electrons moving down towards the slightly positive bromine
atom.
But this time, the top bromine atom becomes attached to both carbon
atoms, with the positive charge being found on the bromine rather than
on one of the carbons. A bromonium ion is formed.

The bromonium ion is then attacked from the back by a bromide ion
formed in a nearby reaction. It can't be attacked by its original bromide
ion because the bromonium ion is completely cluttered up with a positive
bromine on that side.

It doesn't matter which of the carbon atoms which were originally part of
the double bond the bromide ion attacks - the end result would be just
the same.

Note: You can't really draw this mechanism tidily in one line
because the bromide ion has to be in a different place at the
beginning of the second stage than it was at the end of the
first stage.

http://www.chemguide.co.uk/mechanisms/eladd/unsymbr2tt.html (5 of 6)30/12/2004 11:04:10

Explaining electrophilic addition involving bromine and unsymmetrical alkenes

Where would you like to go now?


To menu of electrophilic addition reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/eladd/unsymbr2tt.html (6 of 6)30/12/2004 11:04:10

electrophilic substitution - the sulphonation of benzene

THE SULPHONATION OF BENZENE

This page gives you the facts and a simple, uncluttered mechanism for
the electrophilic substitution reaction between benzene and sulphuric
acid (or sulphur trioxide). If you want this mechanism explained to you in
detail, there is a link at the bottom of the page.

The electrophilic substitution reaction between benzene


and sulphuric acid
The facts
There are two equivalent ways of sulphonating benzene:

Heat benzene under reflux with concentrated sulphuric acid for


several hours.
Warm benzene under reflux at 40C with fuming sulphuric acid for
20 to 30 minutes.

Or:

The product is benzenesulphonic acid.


The electrophile is actually sulphur trioxide, SO3, and you may find the
equation for the sulphonation reaction written:

http://www.chemguide.co.uk/mechanisms/elsub/sulphonation.html (1 of 3)30/12/2004 11:04:21

electrophilic substitution - the sulphonation of benzene

Note: Which version of this equation you use will depend on


what question you are being asked. If the question refers to
the reaction with sulphuric acid, then you must use that one.
If the question refers to SO3 as the electrophile, then you
could use this one.

The formation of the electrophile


The sulphur trioxide electrophile arises in one of two ways depending on
which sort of acid you are using.
Concentrated sulphuric acid contains traces of SO3 due to slight
dissociation of the acid.

Fuming sulphuric acid, H2S2O7, can be thought of as a solution of SO3 in


sulphuric acid - and so is a much richer source of the SO3.
Sulphur trioxide is an electrophile because it is a highly polar molecule
with a fair amount of positive charge on the sulphur atom. It is this which
is attracted to the ring electrons.
The electrophilic substitution mechanism
Stage one

http://www.chemguide.co.uk/mechanisms/elsub/sulphonation.html (2 of 3)30/12/2004 11:04:21

electrophilic substitution - the sulphonation of benzene

Stage two

The second stage of the reaction involves a transfer of the hydrogen


from the ring to the negative oxygen.

Where would you like to go now?


Help! Talk me through this mechanism . . .
To menu of electrophilic substitution reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000 (updated 2004)

http://www.chemguide.co.uk/mechanisms/elsub/sulphonation.html (3 of 3)30/12/2004 11:04:21

Explaining the sulphonation of benzene - electrophilic substitution

EXPLAINING THE SULPHONATION OF BENZENE

This page guides you through the mechanism for the electrophilic
substitution reaction between benzene and sulphuric acid (or sulphur
trioxide)

The electrophilic substitution reaction between benzene


and sulphuric acid
The formation of the electrophile
The electrophile is sulphur trioxide, and this arises in one of two ways
depending on which sort of acid you are using.
Concentrated sulphuric acid contains traces of SO3 due to slight
dissociation of the acid.

Dissociation: This is a reversible splitting up of a


compound. In this case, the sulphuric acid splits into water
and SO3, and at the same time these combine back together
again to make sulphuric acid. The overall effect is that
concentrated sulphuric acid contains small amounts of SO3.

Fuming sulphuric acid, H2S2O7, can be thought of as a solution of SO3


in sulphuric acid - and so is a much richer source of the SO3.

http://www.chemguide.co.uk/mechanisms/elsub/sulphonationtt.html (1 of 4)30/12/2004 11:04:24

Explaining the sulphonation of benzene - electrophilic substitution

Note: You could think of the formula as being essentially


H2SO4.SO3.

Although sulphur trioxide isn't ionic, it is highly polar.


The three oxygens are more electronegative than
the sulphur and so draw electrons towards
themselves. That leaves the sulphur atom fairly
positively charged. It is this + sulphur atom which
attacks the benzene ring.

Note: If you aren't sure about electronegativity and polar


bonds you might like to follow this link.
Use the BACK button on your browser to return to this page.

The electrophilic substitution mechanism


Stage one
Two electrons from the delocalised system are used to form a new bond
with the slightly positive sulphur atom. Because those two electrons
aren't a part of the delocalised system any longer, the delocalisation is
partly broken, and in the process the ring gains a positive charge.
To make room for the new bond between the ring and the sulphur, two of
the electrons joining the sulphur to one of the oxygens are forced right
out on to the oxygen atom, giving it a negative charge.

http://www.chemguide.co.uk/mechanisms/elsub/sulphonationtt.html (2 of 4)30/12/2004 11:04:24

Explaining the sulphonation of benzene - electrophilic substitution

The hydrogen shown on the ring is the one which was already attached
to that top carbon atom. We need to show it there because it has to be
removed in the second stage.

Note: If you aren't sure about the use of curly arrows follow
this link before you go on.
Use the BACK button on your browser to return to this page.

Stage two
This second step is different from all the other benzene electrophilic
substitution reactions you might have already looked at on this site.
This time there isn't a separate negative ion to remove the hydrogen
atom from the ring. Instead it is removed by a lone pair on the negative
oxygen atom.

The lone pair forms a bond with the hydrogen atom, releasing the
electrons in the hydrogen-to-ring bond so that they can re-establish the
delocalisation.
http://www.chemguide.co.uk/mechanisms/elsub/sulphonationtt.html (3 of 4)30/12/2004 11:04:24

Explaining the sulphonation of benzene - electrophilic substitution

This particular mechanism needs to be drawn with rather more thought


than the other electrophilic substitution mechanisms. In particular, you
have to make sure that you put the negative charge on the right oxygen
in the intermediate ion. It must be the one which is closest to the
hydrogen you intend to take off the ring, otherwise there is no way of
drawing sensible curly arrows in the second stage!

Where would you like to go now?


To menu of electrophilic substitution reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/elsub/sulphonationtt.html (4 of 4)30/12/2004 11:04:24

electrophilic substitution in methylbenzene and nitrobenzene

ELECTROPHILIC SUBSTITUTION INTO


ALREADY SUBSTITUTED BENZENE RINGS

This page discusses the problems which arise if you try to write the
mechanism for an electrophilic substitution reaction into a benzene ring
which already has something else attached to it.
There are two problems you might come across:

Whereabouts in the ring does the substitution happen? Does this


make a difference to how you draw the mechanisms?
Can the group already attached to the ring get involved in any
way?

Important! This page assumes that you can already write


the mechanisms for substitution into a simple benzene ring. If
you can't, go back to the electrophilic substitution menu and
read about the reactions you are interested in before you
tackle this page.

Electrophilic substitution in methylbenzene


The nitration of methylbenzene
If you substitute a nitro group, -NO2, into the benzene ring in
methylbenzene, you could possibly get any of the following products:

http://www.chemguide.co.uk/mechanisms/elsub/toluene.html (1 of 7)30/12/2004 11:04:30

electrophilic substitution in methylbenzene and nitrobenzene

The carbon with the methyl group attached is thought of as the number 1
carbon, and the ring is then numbered around from 1 to 6. You number in
a direction (in this case, clockwise) which produces the smaller number
in the name - hence 2-nitromethylbenzene rather than 6nitromethylbenzene.
In the case of methylbenzene, whatever you attach to the ring, you
always get a mixture consisting mainly of the 2- and 4- isomers. The
methyl group is said to be 2,4-directing, in the sense that it seems to
"push" incoming groups into those positions.

Isomers: Molecules which have the same molecular formula


(i.e. contain exactly the same number and type of atoms), but
with a different spatial arrangement of those atoms.

Some other groups which might already be on the ring (for example, the NO2 group in nitrobenzene) "push" incoming groups into the 3- position.
We'll have a quick look at this later on this page.
What to expect in exams
At A' level, you will not be expected to explain why different groups have
different directing effects.
It is possible, however, that you may be expected to remember the
directing effect of one or two groups. It is very difficult to tell this from the
syllabuses, most of which don't specifically exclude the possibility - but
that could simply be an oversight. Your best option is to check past exam
papers, or any support material published by your Exam Board. If in
doubt, contact them and find out exactly what they expect.
http://www.chemguide.co.uk/mechanisms/elsub/toluene.html (2 of 7)30/12/2004 11:04:30

electrophilic substitution in methylbenzene and nitrobenzene

Note: You will find links to help you to contact your Exam
Board on the syllabuses page.

In an exam, then, you may have to remember the directing effect of a


particular group, or more likely you will be told it. Your problem may then
be to write the mechanism.
How to write the mechanism for the nitration of methylbenzene

Important! This assumes that you are already familiar with


the nitration of benzene. If you aren't, read about it before
you go any further.

Reacting methylbenzene with a mixture of concentrated nitric and


sulphuric acids gives both 2-nitromethylbenzene and 4nitromethylbenzene. The mechanism is exactly the same as the nitration
of benzene. You just have to be careful about the way that you draw the
structure of the intermediate ion.
Making 2-nitromethylbenzene (the first step)

This just shows the first step of the electrophilic substitution reaction.
Notice that the partial delocalisation in the intermediate ion covers all the
carbon atoms in the ring except for the one that the -NO2 group gets
attached to.
That is the only point of interest in this example - everything else is just
the same as with the nitration of benzene. The hydrogen atom is then

http://www.chemguide.co.uk/mechanisms/elsub/toluene.html (3 of 7)30/12/2004 11:04:30

electrophilic substitution in methylbenzene and nitrobenzene

removed by an HSO4- ion - exactly as in the benzene case.


Making 4-nitromethylbenzene (the first step)

Once again, the only point of interest is in the way the partial
delocalisation in the intermediate ion is drawn - again, it covers all the
carbon atoms in the ring apart from the one with the -NO2 group attached.
The electrophilic substitution reaction between methylbenzene and
chlorine
This is a good example of a case where what is already attached to the
ring can also get involved in the reaction.
It is possible to get two quite different substitution reactions between
methylbenzene and chlorine depending on the conditions used. The
chlorine can substitute into the ring or into the methyl group.
Here we are only interested in substitution into the ring. This happens in
the presence of aluminium chloride or iron, and in the absence of UV
light.

Note: If you are also interested in substitution into the methyl


group (in the presence of UV light - and with no catalyst
present), you will find this explained in the page on the free
radical reaction between methylbenzene and chlorine.

http://www.chemguide.co.uk/mechanisms/elsub/toluene.html (4 of 7)30/12/2004 11:04:30

electrophilic substitution in methylbenzene and nitrobenzene

Substituting into the ring gives a mixture of 2-chloromethylbenzene and 4chloromethylbenzene.

The mechanisms are exactly the same as the substitution of chlorine into
benzene - although you would have to be careful about the way you draw
the intermediate ion.

Important! This assumes that you are already familiar with


the chlorination of benzene. If you aren't, read about it before
you go any further.

For example, the complete mechanism for substitution into the 4position is:
Stage one

Stage two

http://www.chemguide.co.uk/mechanisms/elsub/toluene.html (5 of 7)30/12/2004 11:04:30

electrophilic substitution in methylbenzene and nitrobenzene

Electrophilic substitution in nitrobenzene


Substitution into the 3- position (the first step)
Methyl groups direct new groups into the 2- and 4- positions, but a nitro
group, -NO2, already on the ring directs incoming groups into the 3position.
For example, if the temperature is raised above 50C, the nitation of
benzene doesn't just produce nitrobenzene - it also produces some 1,3dinitrobenzene. A second nitro group is substituted into the ring in the 3position.

The mechanism is exactly the same as the nitration of benzene or of


methylbenzene - you just have to be careful in drawing the intermediate
ion. Draw the partial delocalisation to include all the carbons except for
the one the new -NO2 group gets attached to.

http://www.chemguide.co.uk/mechanisms/elsub/toluene.html (6 of 7)30/12/2004 11:04:30

electrophilic substitution in methylbenzene and nitrobenzene

In the second stage, the hydrogen atom is then removed by an HSO4ion - exactly as in the benzene case. This isn't shown because there's
nothing new.

Where would you like to go now?


To menu of electrophilic substitution reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/elsub/toluene.html (7 of 7)30/12/2004 11:04:30

free radical substitution in the methylbenzene and chlorine reaction

THE REACTION BETWEEN METHYLBENZENE


AND CHLORINE

A Free Radical Substitution Reaction


This page gives you the facts and a simple, uncluttered
mechanism for the free radical substitution reaction
between methylbenzene (previously known as toluene)
and chlorine. If you want the mechanism explained to you
in detail, there is a link at the bottom of the page.
Methylbenzene has a methyl group attached to a benzene ring. The
hexagon with the circle inside is the standard symbol for this ring. There
is a carbon atom at each corner of the hexagon, and a hydrogen atom on
each carbon apart from the one with the methyl group attached.

Note: There is no need to worry about the bonding in the


benzene ring at this point. If you are interested, you can
follow the link - but it isn't important for now.

The facts
The reaction we are going to explore happens between methylbenzene
and chlorine in the presence of ultraviolet light - typically sunlight. This is
a good example of a photochemical reaction - a reaction brought about
by light.

Note: These reactions are sometimes described as


examples of photocatalysis - reactions catalysed by light. It
is better to use the term "photochemical" and keep the keep
the word "catalysis" for reactions speeded up by actual
substances rather than light.

http://www.chemguide.co.uk/mechanisms/freerad/tolueneandcl2.html (1 of 4)30/12/2004 11:04:34

free radical substitution in the methylbenzene and chlorine reaction

The organic product is (chloromethyl)benzene. The brackets in the name


emphasise that the chlorine is part of the attached methyl group, and isn't
on the ring.
One of the hydrogen atoms in the methyl group has been replaced by a
chlorine atom, so this is a substitution reaction. However, the reaction
doesn't stop there, and all three hydrogens in the methyl group can in
turn be replaced by chlorine atoms. Multiple substitution is dealt with on a
separate page, and you will find a link to that at the bottom of this page.

Important! There is another reaction which happens


between methylbenzene and chlorine in the absence of light
and in the presence of a number of possible catalysts. In that
one, substitution happens in the benzene ring instead of in
the methyl group. You will find this reaction discussed under
electrophilic substitution reactions.

The mechanism
The mechanism involves a chain reaction. During a chain reaction, for
every reactive species you start off with, a new one is generated at the
end - and this keeps the process going.

Species: a useful word which is used in chemistry to mean


any sort of particle you want it to mean. It covers molecules,
ions, atoms, or (in this case) free radicals.

The over-all process is known as free radical substitution, or as a free


radical chain reaction.

http://www.chemguide.co.uk/mechanisms/freerad/tolueneandcl2.html (2 of 4)30/12/2004 11:04:34

free radical substitution in the methylbenzene and chlorine reaction

Note: If you aren't sure about the words free radical or


substitution, read the page What is free radical substitution?
Use the BACK button on your browser to return quickly to this
page.

Chain initiation
The chain is initiated (started) by UV light breaking a chlorine molecule
into free radicals.
Cl2

2Cl

Chain propagation reactions


These are the reactions which keep the chain going.

Chain termination reactions


These are reactions which remove free radicals from the system without
replacing them by new ones. If any two free radicals collide, they will join
together without producing any new radicals.
The simplest example of this is a collision between two chlorine radicals.
2Cl

Cl2

Where would you like to go now?

http://www.chemguide.co.uk/mechanisms/freerad/tolueneandcl2.html (3 of 4)30/12/2004 11:04:34

free radical substitution in the methylbenzene and chlorine reaction

Help! Talk me through this mechanism . . .


Look at multiple substitution in this reaction . . .
To menu of free radical reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/freerad/tolueneandcl2.html (4 of 4)30/12/2004 11:04:34

Explaining the methylbenzene - chlorine free radical substitution mechanism

EXPLAINING THE REACTION BETWEEN


METHYLBENZENE AND CHLORINE

A Free Radical Substitution Reaction


This page guides you through the mechanism for the substitution of one
of the hydrogen atoms in methylbenzene by one chlorine atom. Multiple
substitution is covered separately, and you will find a link at the bottom of
the page.
We are going to talk through this mechanism in a very detailed way so
that you get a feel for what is going on. You couldn't possibly do the
same thing in an exam. At the bottom of the page, you will find the
condensed down version corresponding to the sort of answer you would
produce in an exam.
The role of the UV light
The ultraviolet light is simply a source of energy, and is being used to
break bonds. In fact, the energies in UV are exactly right to break the
bonds in chlorine molecules to produce chlorine atoms.

Note: Only the outer electrons of the chlorine are shown.


Notice also that it is quite acceptable to use a simple view of
atomic structure. There is no point in using a complicated
model of the atom if a simple one will do the job.

http://www.chemguide.co.uk/mechanisms/freerad/tolcl2tt.html (1 of 7)30/12/2004 11:04:38

Explaining the methylbenzene - chlorine free radical substitution mechanism

Because we want to stress the fact that the chlorine atoms have single
unpaired electrons, then we call them chlorine free radicals - or more
usually just chlorine radicals.
To show that a species (either an atom or a group of atoms) is a free
radical, the symbol is written with a dot attached to show the unpaired
electron. The splitting of the chlorine molecule would be shown as:
Cl2

2Cl

Free radicals are formed if a bond splits evenly - each atom getting one
of the two electrons. The name given to this is homolytic fission.
What happens to the chlorine radicals?
Reactions happen because things hit each other. In this case, you need
to think about what the chlorine radicals are likely to hit, and what could
happen as a result of that collision.
At the moment the mixture contains

lots of methylbenzene molecules


lots of chlorine molecules (only a few will have been fractured by
the UV light)
a few chlorine radicals

Let's start with the unproductive collisions.


The least likely collision is between two chlorine radicals. There aren't
very many of them in the mixture and so the chances of them hitting
each other are relatively small. If they do collide, they will combine to
form a chlorine molecule. That's worse than useless because it removes
the active free radicals from the system.
2Cl

Cl2

A chlorine radical could also hit a chlorine molecule. If this happens there
could possibly be an exchange of chlorine atoms, but nothing new would
http://www.chemguide.co.uk/mechanisms/freerad/tolcl2tt.html (2 of 7)30/12/2004 11:04:38

Explaining the methylbenzene - chlorine free radical substitution mechanism

be formed. It is just a wasted collision.


Cl + Cl-Cl
Cl

Cl-Cl +

Note: There is no difference between the chlorine atoms


shown in bold type or ordinary type. They are shown
differently so that the exchange is made clear.

The productive collision happens if a chlorine radical hits a


methylbenzene molecule.

The chlorine radical removes a hydrogen atom from the methyl group.
That hydrogen atom only needs to bring one electron with it to form a
new bond to the chlorine, and so one electron is left behind on the
carbon atom. A new free radical is formed - called a phenylmethyl radical.

Note: Don't worry about the name of this new radical. All that
matters is that you can draw its structure.

http://www.chemguide.co.uk/mechanisms/freerad/tolcl2tt.html (3 of 7)30/12/2004 11:04:38

Explaining the methylbenzene - chlorine free radical substitution mechanism

What happens to the phenylmethyl radicals?


It depends what they collide with. There are three interesting collisions
which need to be explored. Two of these involve a set-back to the
reaction, and only one is useful.
Luckily, the two unhelpful collisions don't happen very often, because
they involve collisions between two free radicals - and there won't be
many of these present in the mixture at any one time.

A phenylmethyl radical hits a chlorine radical. These will combine


to make what you want - (chloromethyl)benzene - but the reaction
removes the active free radicals from the system. That stops any
further reactions happening.

Even worse, two phenylmethyl radicals could hit each other. Not
only does this remove radicals from the system, but produces an
unwanted side reaction.

So what is the useful collision? If a phenylmethyl radical hits a chlorine


molecule (something that's quite likely to occur), the following change
can happen:

The phenylmethyl radical takes one of the chlorine atoms to form


(chloromethyl)benzene (which is what we want to make), but in the
process generates another chlorine radical. This new chlorine radical can
now go through the whole sequence again, and at the end will produce
yet another chlorine radical - and so on and so on.
http://www.chemguide.co.uk/mechanisms/freerad/tolcl2tt.html (4 of 7)30/12/2004 11:04:38

Explaining the methylbenzene - chlorine free radical substitution mechanism

The process is described as a free radical chain reaction. The chain


continues because for every chlorine radical that goes in at the
beginning, a new one is generated at the end.
Chain termination
Does this mean that one tiny burst of UV light, splitting one chlorine
molecule into two free radicals, is enough to convert a whole reactionsworth of methylbenzene and chlorine into (chloromethyl)benzene and
HCl?
Sadly, no! As we've seen, there are collisions which result in the removal
of free radicals without producing any new ones. These radicals can only
be replaced by starting the process all over again with a new burst of
light energy. In practice, then, the chains propagate many thousands of
times, but eventually any chain will be brought to an end by one of these
chain termination processes.

Simplifying all this for exam purposes:


The over-all process is known as free radical substitution, or as a free
radical chain reaction.
Chain initiation
The chain is initiated (started) by UV light breaking a chlorine molecule
into free radicals.
Cl2

2Cl

Chain propagation reactions


These are the reactions which keep the chain going.

http://www.chemguide.co.uk/mechanisms/freerad/tolcl2tt.html (5 of 7)30/12/2004 11:04:38

Explaining the methylbenzene - chlorine free radical substitution mechanism

Chain termination reactions


These are reactions which remove free radicals from the system without
replacing them by new ones. If any two free radicals collide, they will join
together without producing any new radicals.
2Cl

Cl2

Important! If you have found this mechanism difficult


because of the names and structures involved it would be
worth looking at the methane and chlorine reaction. The two
mechanisms are identical as far as the substitution is
concerned, but the methane / chlorine one looks easier!

Where would you like to go now?


Look at multiple substitution in this reaction . . .
To menu of free radical reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .
http://www.chemguide.co.uk/mechanisms/freerad/tolcl2tt.html (6 of 7)30/12/2004 11:04:38

Explaining the methylbenzene - chlorine free radical substitution mechanism

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/freerad/tolcl2tt.html (7 of 7)30/12/2004 11:04:38

free radical substitution in the methane and chlorine reaction

THE REACTION BETWEEN METHANE AND


CHLORINE

A Free Radical Substitution Reaction


This page gives you the facts and a simple, uncluttered mechanism for
the free radical substitution reaction between methane and chlorine. If
you want the mechanism explained to you in detail, there is a link at the
bottom of the page.
The facts
If a mixture of methane and chlorine is exposed to a flame, it explodes producing carbon and hydrogen chloride. This isn't a very useful reaction!
The reaction we are going to explore is a more gentle one between
methane and chlorine in the presence of ultraviolet light - typically
sunlight. This is a good example of a photochemical reaction - a
reaction brought about by light.

Note: These reactions are sometimes described as


examples of photocatalysis - reactions catalysed by light. It
is better to use the term "photochemical" and keep the keep
the word "catalysis" for reactions speeded up by actual
substances rather than light.

http://www.chemguide.co.uk/mechanisms/freerad/ch4andcl2.html (1 of 4)30/12/2004 11:04:40

free radical substitution in the methane and chlorine reaction

CH4 + Cl2

CH3Cl + HCl

The organic product is chloromethane.


One of the hydrogen atoms in the methane has been replaced by a
chlorine atom, so this is a substitution reaction. However, the reaction
doesn't stop there, and all the hydrogens in the methane can in turn be
replaced by chlorine atoms. Multiple substitution is dealt with on a
separate page, and you will find a link to that at the bottom of this page.
The mechanism
The mechanism involves a chain reaction. During a chain reaction, for
every reactive species you start off with, a new one is generated at the
end - and this keeps the process going.

Species: a useful word which is used in chemistry to mean


any sort of particle you want it to mean. It covers molecules,
ions, atoms, or (in this case) free radicals.

The over-all process is known as free radical substitution, or as a free


radical chain reaction.

Note: If you aren't sure about the words free radical or


substitution, read the page What is free radical substitution?
Use the BACK button on your browser to return quickly to this
page.

http://www.chemguide.co.uk/mechanisms/freerad/ch4andcl2.html (2 of 4)30/12/2004 11:04:40

free radical substitution in the methane and chlorine reaction

Chain initiation
The chain is initiated (started) by UV light breaking a chlorine molecule
into free radicals.
Cl2

2Cl

Chain propagation reactions


These are the reactions which keep the chain going.
CH4 + Cl
CH3

CH3

+ Cl2

+ HCl

CH3Cl + Cl

Chain termination reactions


These are reactions which remove free radicals from the system without
replacing them by new ones.
2Cl
CH3
Cl

Cl2
+

CH3 +
CH3

CH3Cl
CH3CH3

Where would you like to go now?


Help! Talk me through this mechanism . . .
Look at multiple substitution in this reaction . . .
Look at why side reactions happen in this reaction . . .
To menu of free radical reactions. . .
To menu of other types of mechanism. . .

http://www.chemguide.co.uk/mechanisms/freerad/ch4andcl2.html (3 of 4)30/12/2004 11:04:40

free radical substitution in the methane and chlorine reaction

To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/freerad/ch4andcl2.html (4 of 4)30/12/2004 11:04:40

Explaining the methane - chlorine free radical substitution mechanism

EXPLAINING THE REACTION BETWEEN


METHANE AND CHLORINE

A Free Radical Substitution Reaction


This page guides you through the mechanism for the substitution of one
of the hydrogen atoms in methane by one chlorine atom. Multiple
substitution is covered separately, and you will find a link at the bottom of
the page.
We are going to talk through this mechanism in a very detailed way so
that you get a feel for what is going on. You couldn't possibly do the
same thing in an exam. At the bottom of the page, you will find the
condensed down version corresponding to the sort of answer you would
produce in an exam.
The role of the UV light
The ultraviolet light is simply a source of energy, and is being used to
break bonds. In fact, the energies in UV are exactly right to break the
bonds in chlorine molecules to produce chlorine atoms.

Note: Only the outer electrons of the chlorine are shown.


Notice also that it is quite acceptable to use a simple view of
atomic structure. There is no point in using a complicated
model of the atom if a simple one will do the job.

http://www.chemguide.co.uk/mechanisms/freerad/ch4andcl2tt.html (1 of 6)30/12/2004 11:04:44

Explaining the methane - chlorine free radical substitution mechanism

Because we want to stress the fact that the chlorine atoms have single
unpaired electrons, then we call them chlorine free radicals - or more
usually just chlorine radicals.
To show that a species (either an atom or a group of atoms) is a free
radical, the symbol is written with a dot attached to show the unpaired
electron. The splitting of the chlorine molecule would be shown as:
Cl2

2Cl

Free radicals are formed if a bond splits evenly - each atom getting one
of the two electrons. The name given to this is homolytic fission.
What happens to the chlorine radicals?
There's nothing magic about reaction mechanisms. Reactions happen
because things hit each other. If the conditions are right something useful
might happen. In this case, you need to think about what the chlorine
radicals are likely to hit, and what could happen as a result of that
collision.
At the moment the mixture contains

lots of methane molecules


lots of chlorine molecules (only a few will have been fractured by
the UV light)
a few chlorine radicals

Let's start with the unproductive collisions.


The least likely collision is between two chlorine radicals. There aren't
very many of them in the mixture and so the chances of them hitting
each other are relatively small. If they do collide, they will combine to
form a chlorine molecule. That's worse than useless because it removes
the active free radicals from the system.
2Cl

Cl2

http://www.chemguide.co.uk/mechanisms/freerad/ch4andcl2tt.html (2 of 6)30/12/2004 11:04:44

Explaining the methane - chlorine free radical substitution mechanism

A chlorine radical could also hit a chlorine molecule. If this happens there
could possibly be an exchange of chlorine atoms, but nothing new would
be formed. It is just a wasted collision.
Cl + Cl-Cl
Cl

Cl-Cl +

Note: There is no difference between the chlorine atoms


shown in bold type or ordinary type. They are shown
differently so that the exchange is made clear.

The productive collision happens if a chlorine radical hits a methane


molecule.

The chlorine radical removes a hydrogen atom from the methane. That
hydrogen atom only needs to bring one electron with it to form a new
bond to the chlorine, and so one electron is left behind on the carbon
atom. A new free radical is formed - this time a methyl radical, CH3 .
CH4 + Cl

CH3

+ HCl

What happens to the methyl radicals?


It depends what they collide with. There are three interesting collisions
which need to be explored. Two of these involve a set-back to the
reaction, and only one is useful.
Luckily, the two unhelpful collisions don't happen very often, because
they involve collisions between two free radicals - and there won't be
many of these present in the mixture at any one time.

http://www.chemguide.co.uk/mechanisms/freerad/ch4andcl2tt.html (3 of 6)30/12/2004 11:04:44

Explaining the methane - chlorine free radical substitution mechanism

CH3
Cl

+
CH3Cl

CH3 +
CH3

CH3CH3

Even though the first reaction seems to produce what you want, the
problem with both of these reactions is that they use up the free radicals
in the system - we'll come back to that problem shortly. The second
reaction, of course, also introduces an impurity into the mixture.
So what is the useful collision? If a methyl radical hits a chlorine
molecule (something that's quite likely to occur), the following change
can happen:
CH3

+ Cl2

CH3Cl + Cl

The methyl radical takes one of the chlorine atoms to form


chloromethane (which is what we want to make), but in the process
generates another chlorine radical. This new chlorine radical can now go
through the whole sequence again, and at the end will produce yet
another chlorine radical - and so on and so on.
The process is described as a free radical chain reaction. The chain
continues because for every chlorine radical that goes in at the
beginning, a new one is generated at the end.
Chain termination
Does this mean that one tiny burst of UV light, splitting one chlorine
molecule into two free radicals, is enough to convert a whole reactionsworth of methane and chlorine into chloromethane and HCl?
Sadly, no! As we've seen, there are collisions which result in the removal
of free radicals without producing any new ones. These radicals can only
be replaced by starting the process all over again with a new burst of
light energy. In practice, then, the chains propagate many thousands of
times, but eventually any chain will be brought to an end by one of these
chain termination processes.

http://www.chemguide.co.uk/mechanisms/freerad/ch4andcl2tt.html (4 of 6)30/12/2004 11:04:44

Explaining the methane - chlorine free radical substitution mechanism

Simplifying all this for exam purposes:


The over-all process is known as free radical substitution, or as a free
radical chain reaction.
Chain initiation
The chain is initiated (started) by UV light breaking a chlorine molecule
into free radicals.
Cl2

2Cl

Chain propagation reactions


These are the reactions which keep the chain going.
CH4 + Cl
CH3

CH3

+ Cl2

+ HCl

CH3Cl + Cl

Chain termination reactions


These are reactions which remove free radicals from the system without
replacing them by new ones.
2Cl
CH3
Cl

Cl2
+

CH3 +
CH3

CH3Cl
CH3CH3

Where would you like to go now?


Look at multiple substitution in this reaction . . .
Look at why side reactions happen in this reaction . . .
To menu of free radical reactions. . .

http://www.chemguide.co.uk/mechanisms/freerad/ch4andcl2tt.html (5 of 6)30/12/2004 11:04:44

Explaining the methane - chlorine free radical substitution mechanism

To menu of other types of mechanism. . .


To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/freerad/ch4andcl2tt.html (6 of 6)30/12/2004 11:04:44

Multiple substitution in the methane and chlorine reaction

MULTIPLE SUBSTITUTION IN THE METHANE


AND CHLORINE REACTION

Warning! We are just about to muddy the water quite


considerably! Don't go on until you are sure that you
understand the mechanism for the production of
chloromethane - and are confident that you could write it in
an exam. If you aren't sure about it, go back to that reaction
and look at it again.
It would be worth checking your syllabus and past exam
papers to see if you need to know about these further
substitution reactions.

The facts
When a mixture of methane and chlorine is exposed to ultraviolet light typically sunlight - a substitution reaction occurs and the organic product
is chloromethane.
CH4 + Cl2

CH3Cl + HCl

However, the reaction doesn't stop there, and all the hydrogens in the
methane can in turn be replaced by chlorine atoms. That means that you
could get any of chloromethane, dichloromethane, trichloromethane or
tetrachloromethane.
CH4 + Cl2

CH3Cl + HCl

CH3Cl + Cl2

CH2Cl2 + HCl

CH2Cl2 + Cl2

CHCl3 + HCl

CHCl3 + Cl2

CCl4 + HCl

You might think that you could control which product you got by the
proportions of methane and chlorine you used, but it isn't as simple as
http://www.chemguide.co.uk/mechanisms/freerad/multisubcl.html (1 of 5)30/12/2004 11:04:46

Multiple substitution in the methane and chlorine reaction

that. If you use enough chlorine you will eventually get CCl4, but any
other proportions will always lead to a mixture of products.

The mechanisms
The formation of multiple substitution products like di-, tri- and
tetrachloromethane can be explained in just the same sort of way as the
formation of the original chloromethane. You just have to look at the
likely collisions as the reaction progresses.
Making dichloromethane
You will remember that the over-all equation for the first stage of the
reaction is
CH4 + Cl2

CH3Cl + HCl

As the reaction proceeds, the methane is getting used up and


chloromethane is taking its place. That means that the argument about
what a chlorine radical is likely to hit changes during the course of the
reaction. As time goes by there is an increasing chance of it hitting a
chloromethane molecule rather than a methane molecule.
When that happens, the chlorine radical can take a hydrogen from the
chloromethane just as well as it could from a methane. In this new case:
CH3Cl + Cl

CH2Cl + HCl

Notice: The dot representing the electron has been moved


against the carbon which is the atom with the unpaired
electron. It would be potentially confusing to leave it next to
the chlorine.

http://www.chemguide.co.uk/mechanisms/freerad/multisubcl.html (2 of 5)30/12/2004 11:04:46

Multiple substitution in the methane and chlorine reaction

The chloromethyl radical formed can then interact with a chlorine


molecule in a new propagation step . . .
CH2Cl + Cl2

CH2Cl2 + Cl

. . . and so dichloromethane is formed and a chlorine radical regenerated.


These propagation steps continue until the chain is terminated by any
two radicals colliding and combining together.
Making tri- and tetrachloromethane
Obviously, as time goes on, there is an increasing chance of the
dichloromethane being hit by a chlorine radical - producing these
propagation steps giving trichloromethane:
CH2Cl2 + Cl
CHCl2 + Cl2

CHCl2 + HCl
CHCl3 + Cl

Care! Don't just skip lightly over these equations. Look


carefully at each one so that you understand what is
happening, and can relate it to what has gone before. Talk
through the equations with yourself.
For example: "A chlorine radical hits the dichloromethane
molecule and steals a hydrogen. That leaves a new radical (I
don't know what it's called, but that doesn't really matter, as
long as I can work out its formula if I have to!), which then
bumps into a chlorine molecule - etc, etc."
Doing this helps you to focus properly on the equations. If
you just read them quickly, you'll have forgotten all about
them again in 15 seconds!

http://www.chemguide.co.uk/mechanisms/freerad/multisubcl.html (3 of 5)30/12/2004 11:04:46

Multiple substitution in the methane and chlorine reaction

As the amount of trichloromethane builds up, then you will get these
steps giving tetrachloromethane:
CHCl3 + Cl
CCl3 + Cl2

CCl3 + HCl
CCl4 + Cl

This is why you will always get a mixture of products whatever the
reaction proportions of methane and chlorine you use. The whole
process is simply governed by chance. Having produced some
chloromethane there is no way that you can prevent it from being hit by
chlorine radicals, and similarly for dichloromethane and trichloromethane.
Trying to produce mainly one product
If you wanted tetrachloromethane, you could of course get it by using a
large excess of chlorine, so that eventually all the hydrogens would be
replaced.
If you wanted mainly chloromethane, you could favour this by using a
huge excess of methane so that the chances were always greater of a
chlorine radical hitting a methane rather than anything else - but even so,
you would still get some mixture of products.
There is no obvious way of getting mainly dichloromethane or
trichloromethane.

Where would you like to go now?


Look at single substitution again . . .
Look at why side reactions happen in this reaction . . .
To menu of free radical reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

http://www.chemguide.co.uk/mechanisms/freerad/multisubcl.html (4 of 5)30/12/2004 11:04:46

Multiple substitution in the methane and chlorine reaction

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/freerad/multisubcl.html (5 of 5)30/12/2004 11:04:46

Side reactions in the methane and chlorine reaction

SIDE REACTIONS IN THE METHANE AND


CHLORINE REACTION

You may remember that one of the chain termination steps produces
ethane, CH3CH3.
CH3 +
CH3

CH3CH3

If chlorine radicals hit that, you are going to get chloroethane and
dichloroethane and so on - and in the course of those reactions you will
get ethyl radicals which could themselves become involved in chain
termination steps leading to propane (from methyl radical hitting ethyl
radical) or butane (from two ethyl radicals combining), which could then
start to undergo substitution - and on and on!
To be honest, all of these side products are going to be present in very
small amounts because the reaction producing ethane won't, by chance,
happen very often, but it nicely illustrates a typical organic chemistry
problem - when you do a reaction in the lab to produce an organic
chemical, a high proportion of your time is spent in purifying the product
from all the side reactions that have gone on!

Where would you like to go now?


To basic facts and mechanism for this reaction . . .
Look at multiple substitution in this reaction . . .
To menu of free radical reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

http://www.chemguide.co.uk/mechanisms/freerad/sidereactcl.html (1 of 2)30/12/2004 11:04:47

Side reactions in the methane and chlorine reaction

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/freerad/sidereactcl.html (2 of 2)30/12/2004 11:04:47

Multiple substitution in the methylbenzene and chlorine reaction

MULTIPLE SUBSTITUTION IN THE


METHYLBENZENE AND CHLORINE REACTION

Warning! Don't go on until you are sure that you understand


the mechanism for the production of (chloromethyl)benzene and are confident that you could write it in an exam. If you
aren't sure about it, go back to that reaction and look at it
again.
It would be worth checking your syllabus and past exam
papers to see if you need to know about these further
substitution reactions.

The facts
When a mixture of methylbenzene and chlorine is exposed to ultraviolet
light - typically sunlight - a substitution reaction occurs in the methyl
group and the organic product is (chloromethyl)benzene.

However, the reaction doesn't stop there, and all the hydrogens in the
methyl group can in turn be replaced by chlorine atoms. That means that
you could get any of (chloromethyl)benzene, (dichloromethyl)benzene, or
(trichloromethyl)benzene.

http://www.chemguide.co.uk/mechanisms/freerad/multisubtol.html (1 of 4)30/12/2004 11:04:53

Multiple substitution in the methylbenzene and chlorine reaction

Care! Look at these equations carefully so that you are sure


that you understand what's going on.
All that's happening is that the three hydrogens in the methyl
group are being replaced by chlorine atoms one at a time.

If you use enough chlorine you will eventually get (trichloromethyl)


benzene, but any other proportions will always lead to a mixture of
products.

The mechanisms
Making (dichloromethyl)benzene
You will remember that the over-all equation for the first stage of the
reaction is

As the reaction proceeds, the methylbenzene is getting used up and


(chloromethyl)benzene is taking its place. Remember that these
reactions happen because chlorine radicals bump into things. As time
http://www.chemguide.co.uk/mechanisms/freerad/multisubtol.html (2 of 4)30/12/2004 11:04:53

Multiple substitution in the methylbenzene and chlorine reaction

goes by there is an increasing chance of a chlorine radical hitting a


(chloromethyl)benzene molecule rather than a methylbenzene molecule.
When that happens, the chlorine radical can take a hydrogen from the
(chloromethyl)benzene just as well as it could from a methylbenzene
molecule. In this new case:

The new radical formed can then interact with a chlorine molecule in a
new propagation step . . .

. . . and so (dichloromethyl)benzene is formed and a chlorine radical


regenerated.
These propagation steps continue until the chain is terminated by any
two radicals colliding and combining together.
Making (trichloromethyl)benzene
Obviously, as time goes on, there is an increasing chance of the
(dichloromethyl)benzene being hit by a chlorine radical - producing these
propagation steps giving (trichloromethyl)benzene:

http://www.chemguide.co.uk/mechanisms/freerad/multisubtol.html (3 of 4)30/12/2004 11:04:53

Multiple substitution in the methylbenzene and chlorine reaction

Care! Nothing new is hapening here, but don't just glance


briefly at these equations and then move on. Talk them
through with yourself.
"A chlorine radical takes a hydrogen away from the first
molecule (I can't remember what it's called, but that doesn't
matter much, because I know how to draw it!) and forms a
new radical. That bumps into a chlorine molecule, and gives
the product and a new chlorine radical - which can go
through the process all over again. So it's a chain reaction."

You will always get a mixture of products whatever the reaction


proportions of methylbenzene and chlorine you use. The whole process
is simply governed by chance. Having produced some (chloromethyl)
benzene there is no way that you can prevent it from being hit by chlorine
radicals, and similarly for (dichloromethyl)benzene.

Where would you like to go now?


Look at single substitution again . . .
To menu of free radical reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/freerad/multisubtol.html (4 of 4)30/12/2004 11:04:53

Arenes (aromatic hydrocarbons) Menu

Understanding Chemistry

ARENES MENU

Arenes are aromatic hydrocarbons (compounds based on benzene rings)


such as benzene and methylbenzene.

Background . . .
An introduction to the arenes and their physical properties.
Manufacture . . .
The manufacture of arenes from petroleum by reforming.
Nitration . . .
The nitration of benzene and methylbenzene.
Halogenation . . .
The reactions of benzene and methylbenzene with chlorine and
bromine.
Friedel-Crafts reactions . . .
The alkylation and acylation of benzene and methylbenzene.
Other reactions of benzene and methylbenzene . . .
Combustion, hydrogenation, sulphonation, and the oxidation of

http://www.chemguide.co.uk/organicprops/arenesmenu.html (1 of 2)30/12/2004 11:04:54

Arenes (aromatic hydrocarbons) Menu

side chains on benzene rings.

Go to menu of other organic compounds . . .


Go to Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/arenesmenu.html (2 of 2)30/12/2004 11:04:54

an introduction to arenes (aromatic hydrocarbons)

INTRODUCING ARENES (AROMATIC


HYDROCARBONS)

This page looks at the structures and physical properties of the simplest
arenes (benzene and methylbenzene), together with a very brief
introduction to their reactivity. Much of this is covered in detail elsewhere
on the site - in sections on bonding and mechanisms, for example. You
will find links to all of these pages.

What are arenes?


Arenes are aromatic hydrocarbons. The term "aromatic" originally
referred to their pleasant smells, but now implies a particular sort of
delocalised bonding (see below).
The arenes are based on benzene rings. The simplest of them is
benzene itself, C6H6. The next simplest is methylbenzene (old name:
toluene) which has one of the hydrogen atoms attached to the ring
replaced by a methyl group - C6H5CH3.

The structure of benzene


The structure of benzene is covered in detail in two pages in the organic
bonding section of this site. It is important to understand these thoroughly
to make sense of benzene and methylbenzene chemistry. Unless you
have read these pages recently, you should spend some time on them
now before you go any further on this page.

http://www.chemguide.co.uk/organicprops/arenes/background.html (1 of 7)30/12/2004 11:04:58

an introduction to arenes (aromatic hydrocarbons)

Note: You will find two pages dealing with the bonding in
benzene. One deals with the Kelul structure, and the other
with the modern delocalised structure.
You should read both of these pages starting with the one
about the Kelul structure. You will find a link to the second
page at the bottom of the first one.
This is likely to take you some time because you will probably
have to follow up other links as well in order to fully
understand the second page. Don't try to short-cut this.
Use the BACK button (or more likely, the HISTORY file or
GO menu) on your browser to return to this page later.

What you need to understand about benzene is:

Benzene, C6H6, is a planar molecule containing a ring of six


carbon atoms each with a hydrogen atom attached.
The six carbon atoms form a perfectly regular hexagon. All the
carbon-carbon bonds have exactly the same lengths - somewhere
between single and double bonds.

There are delocalised electrons above and below the plane of the
ring.

This diagram shows one of the molecular orbitals containing two of


the delocalised electrons, which may be found anywhere within the
two "doughnuts". The other molecular orbitals are almost never
drawn.
http://www.chemguide.co.uk/organicprops/arenes/background.html (2 of 7)30/12/2004 11:04:58

an introduction to arenes (aromatic hydrocarbons)

The presence of the delocalised electrons makes benzene


particularly stable.
Benzene resists addition reactions because that would involve
breaking the delocalisation and losing that stability.
Benzene is represented by this symbol, where the
circle represents the delocalised electrons, and
each corner of the hexagon has a carbon atom
with a hydrogen attached.

The structure of methylbenzene (toluene)


Methylbenzene just has a methyl group attached to the benzene ring replacing one of the hydrogen atoms.

Attached groups are often drawn at the top of the ring, but you may
occasionally find them drawn in other places with the ring rotated.

Note: If you have to count up the hydrogens in a diagram of


this kind, don't forget that there isn't a hydrogen atom at any
corner of the hexagon where there is something else
attached. The molecular formula of methylbenzene, for
example, is C7H8. Check that you get that answer from this
diagram!

http://www.chemguide.co.uk/organicprops/arenes/background.html (3 of 7)30/12/2004 11:04:58

an introduction to arenes (aromatic hydrocarbons)

Physical properties
Boiling points
In benzene, the only attractions between neighbouring molecules are van
der Waals dispersion forces. There is no permanent dipole on the
molecule.
Benzene boils at 80C - rather higher than other hydrocarbons of similar
molecular size (pentane and hexane, for example). This is presumably
due to the ease with which temporary dipoles can be set up involving the
delocalised electrons.

Note: If you aren't happy about van der Waals dispersion


forces then you should follow this link before you go on.
Use the BACK button on your browser to return to this page.

Methylbenzene boils at 111C. It is a bigger molecule and so the van der


Waals dispersion forces will be bigger.
Methylbenzene also has a small permanent dipole, so there will be
dipole-dipole attractions as well as dispersion forces. The dipole is due to
the CH3 group's tendency to "push" electrons away from itself. This also
affects the reactivity of methylbenzene (see below).

Melting points
You might have expected that methylbenzene's melting point would be
higher than benzene's as well, but it isn't - it is much lower! Benzene
melts at 5.5C; methylbenzene at -95C.

http://www.chemguide.co.uk/organicprops/arenes/background.html (4 of 7)30/12/2004 11:04:58

an introduction to arenes (aromatic hydrocarbons)

Molecules must pack efficiently in the solid if they are to make best use
of their intermolecular forces. Benzene is a tidy, symmetrical molecule
and packs very efficiently. The methyl group sticking out in
methylbenzene tends to disrupt the closeness of the packing. If the
molecules aren't as closely packed, the intermolecular forces don't work
as well and so the melting point falls.

Solubility in water
The arenes are insoluble in water.
Benzene is quite large compared with a water molecule. In order for
benzene to dissolve it would have to break lots of existing hydrogen
bonds between water molecules. You also have to break the quite strong
van der Waals dispersion forces between benzene molecules. Both of
these cost energy.
The only new forces between the benzene and the water would be van
der Waals dispersion forces. These aren't as strong as hydrogen bonds
(or the original dispersion forces in the benzene), and so you wouldn't get
much energy released when they form.
It simply isn't energetically profitable for benzene to dissolve in water. It
would, of course, be even worse for larger arene molecules.

Reactivity
Benzene
It has already been pointed out above that benzene is resistant to
addition reactions. Adding something new to the ring would need you to
use some of the delocalised electrons to form bonds with whatever you
are adding. That results in a major loss of stability as the delocalisation is
broken.
Instead, benzene mainly undergoes substitution reactions - replacing one
or more of the hydrogen atoms by something new. That leaves the
delocalised electrons as they were.
http://www.chemguide.co.uk/organicprops/arenes/background.html (5 of 7)30/12/2004 11:04:58

an introduction to arenes (aromatic hydrocarbons)

Note: The delocalisation is, in fact, broken during a


substitution reaction, but reforms at the end. If you are
interested in the mechanisms for benzene's substitution
reactions you could follow this link.
Should you want to come back to this page later, it would
probably be best to use the HISTORY file or GO menu on
your browser.

Methylbenzene
You have to consider the reactivity of something like methylbenzene in
two distinct bits:

For example, if you explore other pages in this section, you will find that
alkyl groups attached to a benzene ring are oxidised by alkaline
potassium manganate(VII) solution. This doesn't happen in the absence
of the benzene ring.
The tendency of the CH3 group to "push" electrons away from itself also
has an effect on the ring, making methylbenzene react more quickly than
benzene itself. You will find this explored in other pages in this section as
well.

http://www.chemguide.co.uk/organicprops/arenes/background.html (6 of 7)30/12/2004 11:04:58

an introduction to arenes (aromatic hydrocarbons)

Where would you like to go now?


To the arenes menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/arenes/background.html (7 of 7)30/12/2004 11:04:58

intermolecular bonding - van der Waals forces

INTERMOLECULAR BONDING - VAN DER


WAALS FORCES

This page explains the origin of the two weaker forms of intermolecular
attractions - van der Waals dispersion forces and dipole-dipole
attractions. If you are also interested in hydrogen bonding there is a link
at the bottom of the page.

What are intermolecular attractions?


Intermolecular versus intramolecular bonds
Intermolecular attractions are attractions between one molecule and a
neighbouring molecule. The forces of attraction which hold an individual
molecule together (for example, the covalent bonds) are known as
intramolecular attractions. These two words are so confusingly similar
that it is safer to abandon one of them and never use it. The term
"intramolecular" won't be used again on this site.
All molecules experience intermolecular attractions, although in some
cases those attractions are very weak. Even in a gas like hydrogen, H2, if
you slow the molecules down by cooling the gas, the attractions are large
enough for the molecules to stick together eventually to form a liquid and
then a solid.
In hydrogen's case the attractions are so weak that the molecules have
to be cooled to 21 K (-252C) before the attractions are enough to
condense the hydrogen as a liquid. Helium's intermolecular attractions
are even weaker - the molecules won't stick together to form a liquid until
the temperature drops to 4 K (-269C).

van der Waals forces: dispersion forces


Dispersion forces (one of the two types of van der Waals force we are
http://www.chemguide.co.uk/atoms/bonding/vdw.html (1 of 9)30/12/2004 11:05:30

intermolecular bonding - van der Waals forces

dealing with on this page) are also known as "London forces" (named
after Fritz London who first suggested how they might arise).
The origin of van der Waals dispersion forces
Temporary fluctuating dipoles
Attractions are electrical in nature. In a symmetrical molecule like
hydrogen, however, there doesn't seem to be any electrical distortion to
produce positive or negative parts. But that's only true on average.

The lozenge-shaped diagram represents a small symmetrical molecule H2, perhaps, or Br2. The even shading shows that on average there is no
electrical distortion.
But the electrons are mobile, and at any one instant they might find
themselves towards one end of the molecule, making that end -. The
other end will be temporarily short of electrons and so becomes +.

Note: (read as "delta") means "slightly" - so + means


"slightly positive".

An instant later the electrons may well have moved up to the other end,
reversing the polarity of the molecule.

This constant "sloshing around" of the electrons in the molecule causes


rapidly fluctuating dipoles even in the most symmetrical molecule. It even
happens in monatomic molecules - molecules of noble gases, like
helium, which consist of a single atom.
http://www.chemguide.co.uk/atoms/bonding/vdw.html (2 of 9)30/12/2004 11:05:30

intermolecular bonding - van der Waals forces

If both the helium electrons happen to be on one side of the atom at the
same time, the nucleus is no longer properly covered by electrons for
that instant.

How temporary dipoles give rise to intermolecular attractions


Imagine a molecule which has a temporary polarity being approached by
one which happens to be entirely non-polar just at that moment. (A pretty
unlikely event, but it makes the diagrams much easier to draw! In reality,
one of the molecules is likely to have a greater polarity than the other at
that time - and so will be the dominant one.)

As the right hand molecule approaches, its electrons will tend to be


attracted by the slightly positive end of the left hand one.
This sets up an induced dipole in the approaching molecule, which is
orientated in such a way that the + end of one is attracted to the - end
of the other.

An instant later the electrons in the left hand molecule may well have
moved up the other end. In doing so, they will repel the electrons in the
right hand one.

http://www.chemguide.co.uk/atoms/bonding/vdw.html (3 of 9)30/12/2004 11:05:30

intermolecular bonding - van der Waals forces

The polarity of both molecules reverses, but you still have + attracting
-. As long as the molecules stay close to each other the polarities will
continue to fluctuate in synchronisation so that the attraction is always
maintained.

There is no reason why this has to be restricted to two molecules. As


long as the molecules are close together this synchronised movement of
the electrons can occur over huge numbers of molecules.

This diagram shows how a whole lattice of molecules could be held


together in a solid using van der Waals dispersion forces. An instant
later, of course, you would have to draw a quite different arrangement of
the distribution of the electrons as they shifted around - but always in
synchronisation.

The strength of dispersion forces


Dispersion forces between molecules are much weaker than the covalent
bonds within molecules. It isn't possible to give any exact value, because
the size of the attraction varies considerably with the size of the molecule
and its shape.
How molecular size affects the strength of the dispersion forces
The boiling points of the noble gases are
helium

http://www.chemguide.co.uk/atoms/bonding/vdw.html (4 of 9)30/12/2004 11:05:30

-269C

intermolecular bonding - van der Waals forces

neon
argon
krypton
xenon
radon

-246C
-186C
-152C
-108C
-62C

All of these elements have monatomic molecules.


The reason that the boiling points increase as you go down the group is
that the number of electrons increases, and so also does the radius of
the atom. The more electrons you have, and the more distance over
which they can move, the bigger the possible temporary dipoles and
therefore the bigger the dispersion forces.

Because of the greater temporary dipoles, xenon molecules are "stickier"


than neon molecules. Neon molecules will break away from each other at
much lower temperatures than xenon molecules - hence neon has the
lower boiling point.
This is the reason that (all other things being equal) bigger molecules
have higher boiling points than small ones. Bigger molecules have more
electrons and more distance over which temporary dipoles can develop and so the bigger molecules are "stickier".

How molecular shape affects the strength of the dispersion forces


The shapes of the molecules also matter. Long thin molecules can
develop bigger temporary dipoles due to electron movement than short
fat ones containing the same numbers of electrons.
Long thin molecules can also lie closer together - these attractions are at
their most effective if the molecules are really close.

http://www.chemguide.co.uk/atoms/bonding/vdw.html (5 of 9)30/12/2004 11:05:30

intermolecular bonding - van der Waals forces

For example, the hydrocarbon molecules butane and 2-methylpropane


both have a molecular formula C4H10, but the atoms are arranged
differently. In butane the carbon atoms are arranged in a single chain,
but 2-methylpropane is a shorter chain with a branch.

Butane has a higher boiling point because the dispersion forces are
greater. The molecules are longer (and so set up bigger temporary
dipoles) and can lie closer together than the shorter, fatter 2methylpropane molecules.

van der Waals forces: dipole-dipole interactions

Warning! There's a bit of a problem here with modern


A'level syllabuses. The majority of the syllabuses talk as if
dipole-dipole interactions were quite distinct from van der
Waals forces. Such a syllabus will talk about van der Waals
forces (meaning dispersion forces) and, separately, dipoledipole interactions.
All intermolecular attractions are known collectively as van
der Waals forces. The various different types were first
explained by different people at different times. Dispersion
forces, for example, were described by London in 1930;
dipole-dipole interactions by Keesom in 1912.
This oddity in the syllabuses doesn't matter in the least as far
as understanding is concerned - but you obviously must
know what your particular examiners mean by the terms they
use in the questions. Check your syllabus.
If you don't have a copy of your syllabus follow this link to find
out how to get one.

http://www.chemguide.co.uk/atoms/bonding/vdw.html (6 of 9)30/12/2004 11:05:30

intermolecular bonding - van der Waals forces

A molecule like HCl has a permanent dipole because chlorine is more


electronegative than hydrogen. These permanent, in-built dipoles will
cause the molecules to attract each other rather more than they
otherwise would if they had to rely only on dispersion forces.

Note: If you aren't happy about electronegativity and polar


molecules, follow this link before you go on.

It's important to realise that all molecules experience dispersion forces.


Dipole-dipole interactions are not an alternative to dispersion forces they occur in addition to them. Molecules which have permanent dipoles
will therefore have boiling points rather higher than molecules which only
have temporary fluctuating dipoles.
Surprisingly dipole-dipole attractions are fairly minor compared with
dispersion forces, and their effect can only really be seen if you compare
two molecules with the same number of electrons and the same size. For
example, the boiling points of ethane, CH3CH3, and fluoromethane,
CH3F, are

Why choose these two molecules to compare? Both have identical


numbers of electrons, and if you made models you would find that the
sizes were similar - as you can see in the diagrams. That means that the
dispersion forces in both molecules should be much the same.

http://www.chemguide.co.uk/atoms/bonding/vdw.html (7 of 9)30/12/2004 11:05:30

intermolecular bonding - van der Waals forces

The higher boiling point of fluoromethane is due to the large permanent


dipole on the molecule because of the high electronegativity of fluorine.
However, even given the large permanent polarity of the molecule, the
boiling point has only been increased by some 10.

Here is another example showing the


dominance of the dispersion forces.
Trichloromethane, CHCl3, is a highly polar
molecule because of the electronegativity of
the three chlorines. There will be quite strong
dipole-dipole attractions between one molecule
and its neighbours.

On the other hand, tetrachloromethane, CCl4,


is non-polar. The outside of the molecule is
uniformly - in all directions. CCl4 has to rely
only on dispersion forces.

So which has the highest boiling point? CCl4 does, because it is a bigger
molecule with more electrons. The increase in the dispersion forces more
than compensates for the loss of dipole-dipole interactions.
The boiling points are:
CHCl3

61.2C

CCl4

76.8C

Where would you like to go now?


To look at hydrogen bonding . . .

http://www.chemguide.co.uk/atoms/bonding/vdw.html (8 of 9)30/12/2004 11:05:30

intermolecular bonding - van der Waals forces

To the bonding menu . . .


To the atomic structure and bonding menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/atoms/bonding/vdw.html (9 of 9)30/12/2004 11:05:30

electronegativity

ELECTRONEGATIVITY

This page explains what electronegativity is, and how and why it varies
around the Periodic Table. It looks at the way that electronegativity
differences affect bond type and explains what is meant by polar bonds
and polar molecules.
If you are interested in electronegativity in an organic chemistry context,
you will find a link at the bottom of this page.

What is electronegativity
Definition
Electronegativity is a measure of the tendency of an atom to attract a
bonding pair of electrons.
The Pauling scale is the most commonly used. Fluorine (the most
electronegative element) is assigned a value of 4.0, and values range
down to caesium and francium which are the least electronegative at 0.7.

What happens if two atoms of equal electronegativity bond


together?
Consider a bond between two atoms, A and B. Each atom may be
forming other bonds as well as the one shown - but these are irrelevant
to the argument.

If the atoms are equally electronegative, both have the same tendency to
attract the bonding pair of electrons, and so it will be found on average
half way between the two atoms. To get a bond like this, A and B would
usually have to be the same atom. You will find this sort of bond in, for
http://www.chemguide.co.uk/atoms/bonding/electroneg.html (1 of 10)30/12/2004 11:05:36

electronegativity

example, H2 or Cl2 molecules.

Note: It's important to realise that this is an average picture.


The electrons are actually in a molecular orbital, and are
moving around all the time within that orbital.

This sort of bond could be thought of as being a "pure" covalent bond where the electrons are shared evenly between the two atoms.

What happens if B is slightly more electronegative than A?


B will attract the electron pair rather more than A does.

That means that the B end of the bond has more than its fair share of
electron density and so becomes slightly negative. At the same time, the
A end (rather short of electrons) becomes slightly positive. In the
diagram, " " (read as "delta") means "slightly" - so + means "slightly
positive".
Defining polar bonds
This is described as a polar bond. A polar bond is a covalent bond in
which there is a separation of charge between one end and the other - in
other words in which one end is slightly positive and the other slightly
negative. Examples include most covalent bonds. The hydrogen-chlorine
bond in HCl or the hydrogen-oxygen bonds in water are typical.

What happens if B is a lot more electronegative than A?


In this case, the electron pair is dragged right over to B's end of the bond.
To all intents and purposes, A has lost control of its electron, and B has
http://www.chemguide.co.uk/atoms/bonding/electroneg.html (2 of 10)30/12/2004 11:05:36

electronegativity

complete control over both electrons. Ions have been formed.

A "spectrum" of bonds
The implication of all this is that there is no clear-cut division between
covalent and ionic bonds. In a pure covalent bond, the electrons are held
on average exactly half way between the atoms. In a polar bond, the
electrons have been dragged slightly towards one end.
How far does this dragging have to go before the bond counts as ionic?
There is no real answer to that. You normally think of sodium chloride as
being a typically ionic solid, but even here the sodium hasn't completely
lost control of its electron. Because of the properties of sodium chloride,
however, we tend to count it as if it were purely ionic.

Note: Don't worry too much about the exact cut-off point
between polar covalent bonds and ionic bonds. At A'level,
examples will tend to avoid the grey areas - they will be
obviously covalent or obviously ionic. You will, however, be
expected to realise that those grey areas exist.

Lithium iodide, on the other hand, would be described as being "ionic


with some covalent character". In this case, the pair of electrons hasn't
moved entirely over to the iodine end of the bond. Lithium iodide, for
example, dissolves in organic solvents like ethanol - not something which
ionic substances normally do.

Summary

No electronegativity difference between two atoms leads to a pure


non-polar covalent bond.

http://www.chemguide.co.uk/atoms/bonding/electroneg.html (3 of 10)30/12/2004 11:05:36

electronegativity

A small electronegativity difference leads to a polar covalent bond.

A large electronegativity difference leads to an ionic bond.

Polar bonds and polar molecules


In a simple molecule like HCl, if the bond is polar, so also is the whole
molecule. What about more complicated molecules?
In CCl4, each bond is polar.

Note: Ordinary lines represent bonds in the plane of the


screen or paper. Dotted lines represent bonds going away
from you into the screen or paper. Wedged lines represent
bonds coming out of the screen or paper towards you.

The molecule as a whole, however, isn't polar - in the sense that it


doesn't have an end (or a side) which is slightly negative and one which
is slightly positive. The whole of the outside of the molecule is somewhat
negative, but there is no overall separation of charge from top to bottom,
or from left to right.
By contrast, CHCl3 is polar.

http://www.chemguide.co.uk/atoms/bonding/electroneg.html (4 of 10)30/12/2004 11:05:36

electronegativity

The hydrogen at the top of the molecule is less electronegative than


carbon and so is slightly positive. This means that the molecule now has
a slightly positive "top" and a slightly negative "bottom", and so is overall
a polar molecule.
A polar molecule will need to be "lop-sided" in some way.

Patterns of electronegativity in the Periodic Table


The most electronegative element is fluorine. If you remember that fact,
everything becomes easy, because electronegativity must always
increase towards fluorine in the Periodic Table.

Note: This simplification ignores the noble gases.


Historically this is because they were believed not to form
bonds - and if they don't form bonds, they can't have an
electronegativity value. Even now that we know that some of
them do form bonds, data sources still don't quote
electronegativity values for them.

http://www.chemguide.co.uk/atoms/bonding/electroneg.html (5 of 10)30/12/2004 11:05:36

electronegativity

Trends in electronegativity across a period


As you go across a period the electronegativity increases. The chart
shows electronegativities from sodium to chlorine - you have to ignore
argon. It doesn't have an electronegativity, because it doesn't form bonds.

Trends in electronegativity down a group


As you go down a group, electronegativity decreases. (If it increases up
to fluorine, it must decrease as you go down.) The chart shows the
patterns of electronegativity in Groups 1 and 7.

http://www.chemguide.co.uk/atoms/bonding/electroneg.html (6 of 10)30/12/2004 11:05:36

electronegativity

Explaining the patterns in electronegativity


The attraction that a bonding pair of electrons feels for a particular
nucleus depends on:

the number of protons in the nucleus;

the distance from the nucleus;

the amount of screening by inner electrons.

Note: If you aren't happy about the concept of screening or


shielding, it would pay you to read the page on ionisation
energies before you go on. The factors influencing ionisation
energies are just the same as those influencing
electronegativities.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/atoms/bonding/electroneg.html (7 of 10)30/12/2004 11:05:36

electronegativity

Why does electronegativity increase across a period?


Consider sodium at the beginning of period 3 and chlorine at the end
(ignoring the noble gas, argon). Think of sodium chloride as if it were
covalently bonded.

Both sodium and chlorine have their bonding electrons in the 3-level. The
electron pair is screened from both nuclei by the 1s, 2s and 2p electrons,
but the chlorine nucleus has 6 more protons in it. It is no wonder the
electron pair gets dragged so far towards the chlorine that ions are
formed.
Electronegativity increases across a period because the number of
charges on the nucleus increases. That attracts the bonding pair of
electrons more strongly.
Why does electronegativity fall as you go down a group?
Think of hydrogen fluoride and hydrogen chloride.

The bonding pair is shielded from the fluorine's nucleus only by the 1s2
electrons. In the chlorine case it is shielded by all the 1s22s22p6
electrons.
In each case there is a net pull from the centre of the fluorine or chlorine
of +7. But fluorine has the bonding pair in the 2-level rather than the 3level as it is in chlorine. If it is closer to the nucleus, the attraction is
greater.
http://www.chemguide.co.uk/atoms/bonding/electroneg.html (8 of 10)30/12/2004 11:05:36

electronegativity

As you go down a group, electronegativity decreases because the


bonding pair of electrons is increasingly distant from the attraction of the
nucleus.

The polarising ability of positive ions


What do we mean by "polarising ability"?
In the discussion so far, we've thought of ions as arising from highly
distorted covalent bonds. You can also think of it the other way round.
Solid aluminium chloride is covalent. Imagine instead that it was ionic. It
would contain Al3+ and Cl- ions.
The aluminium ion is very small and is packed with three positive
charges - the "charge density" is therefore very high. That will have a
considerable effect on any nearby electrons.

In the case of aluminium chloride, the electron pairs are dragged back
towards the aluminium to such an extent that the bonds becomes
covalent.

Factors affecting polarising ability


Positive ions can have the effect of polarising (electrically distorting)
nearby negative ions. The polarising ability depends on the charge
density in the positive ion.

http://www.chemguide.co.uk/atoms/bonding/electroneg.html (9 of 10)30/12/2004 11:05:36

electronegativity

Polarising ability increases as the positive ion gets smaller and the
number of charges gets larger.
As a negative ion gets bigger, it becomes easier to polarise. For
example, in an iodide ion, I-, the outer electrons are in the 5-level relatively distant from the nucleus.
A positive ion would be more effective in attracting a pair of electrons
from an iodide ion than the corresponding electrons in, say, a fluoride ion
where they are much closer to the nucleus.
Aluminium iodide is covalent because the electron pair is easily dragged
away from the iodide ion. On the other hand, aluminium fluoride is ionic
because the aluminium ion can't polarise the small fluoride ion sufficiently
to form a covalent bond.

Where would you like to go now?


To look at electronegativity in an organic chemistry
context . . .
To the bonding menu . . .
To the atomic structure and bonding menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/atoms/bonding/electroneg.html (10 of 10)30/12/2004 11:05:36

intermolecular bonding - hydrogen bonds

INTERMOLECULAR BONDING - HYDROGEN


BONDS

This page explains the origin of hydrogen bonding - a relatively strong


form of intermolecular attraction. If you are also interested in the weaker
intermolecular forces (van der Waals dispersion forces and dipole-dipole
interactions), there is a link at the bottom of the page.

The evidence for hydrogen bonding


Many elements form compounds with hydrogen - referred to as
"hydrides". If you plot the boiling points of the hydrides of the Group 4
elements, you find that the boiling points increase as you go down the
group.

The increase in boiling point happens because the molecules are getting
larger with more electrons, and so van der Waals dispersion forces
become greater.

Note: If you aren't sure about van der Waals dispersion


forces, it would pay you to follow this link before you go on.

http://www.chemguide.co.uk/atoms/bonding/hbond.html (1 of 9)30/12/2004 11:05:42

intermolecular bonding - hydrogen bonds

If you repeat this exercise with the hydrides of elements in Groups 5, 6


and 7, something odd happens.

Although for the most part the trend is exactly the same as in group 4 (for
exactly the same reasons), the boiling point of the hydride of the first
element in each group is abnormally high.
In the cases of NH3, H2O and HF there must be some additional
intermolecular forces of attraction, requiring significantly more heat
energy to break. These relatively powerful intermolecular forces are
described as hydrogen bonds.

The origin of hydrogen bonding


The molecules which have this extra bonding are:

http://www.chemguide.co.uk/atoms/bonding/hbond.html (2 of 9)30/12/2004 11:05:42

intermolecular bonding - hydrogen bonds

Note: The solid line represents a bond in the plane of the


screen or paper. Dotted bonds are going back into the screen
or paper away from you, and wedge-shaped ones are coming
out towards you.

Notice that in each of these molecules:

The hydrogen is attached directly to one of the most


electronegative elements, causing the hydrogen to acquire a
significant amount of positive charge.
Each of the elements to which the hydrogen is attached is not only
significantly negative, but also has at least one "active" lone pair.
Lone pairs at the 2-level have the electrons contained in a
relatively small volume of space which therefore has a high density
of negative charge. Lone pairs at higher levels are more diffuse
and not so attractive to positive things.

Note: If you aren't happy about electronegativity, you should


follow this link before you go on.

Consider two water molecules coming close together.

http://www.chemguide.co.uk/atoms/bonding/hbond.html (3 of 9)30/12/2004 11:05:42

intermolecular bonding - hydrogen bonds

The + hydrogen is so strongly attracted to the lone pair that it is almost


as if you were beginning to form a co-ordinate (dative covalent) bond. It
doesn't go that far, but the attraction is significantly stronger than an
ordinary dipole-dipole interaction.
Hydrogen bonds have about a tenth of the strength of an average
covalent bond, and are being constantly broken and reformed in liquid
water. If you liken the covalent bond between the oxygen and hydrogen
to a stable marriage, the hydrogen bond has "just good friends" status.
On the same scale, van der Waals attractions represent mere passing
acquaintances!
Water as a "perfect" example of hydrogen bonding
Notice that each water molecule can potentially form four hydrogen
bonds with surrounding water molecules. There are exactly the right
numbers of + hydrogens and lone pairs so that every one of them can
be involved in hydrogen bonding.
This is why the boiling point of water is higher than that of ammonia or
hydrogen fluoride. In the case of ammonia, the amount of hydrogen
bonding is limited by the fact that each nitrogen only has one lone pair. In
a group of ammonia molecules, there aren't enough lone pairs to go
around to satisfy all the hydrogens.
In hydrogen fluoride, the problem is a shortage of hydrogens. In water,
there are exactly the right number of each. Water could be considered as
the "perfect" hydrogen bonded system.

Note: You will find more discussion on the effect of


hydrogen bonding on the properties of water in the page on
molecular structures.

http://www.chemguide.co.uk/atoms/bonding/hbond.html (4 of 9)30/12/2004 11:05:42

intermolecular bonding - hydrogen bonds

More complex examples of hydrogen bonding


The hydration of negative ions
When an ionic substance dissolves in water, water molecules cluster
around the separated ions. This process is called hydration.
Water frequently attaches to positive ions by co-ordinate (dative
covalent) bonds. It bonds to negative ions using hydrogen bonds.

Note: If you are interested in the bonding in hydrated


positive ions, you could follow this link to co-ordinate (dative
covalent) bonding.

The diagram shows the potential hydrogen bonds formed to a chloride


ion, Cl-. Although the lone pairs in the chloride ion are at the 3-level and
wouldn't normally be active enough to form hydrogen bonds, in this case
they are made more attractive by the full negative charge on the chlorine.

However complicated the negative ion, there will always be lone pairs
that the hydrogen atoms from the water molecules can hydrogen bond to.

http://www.chemguide.co.uk/atoms/bonding/hbond.html (5 of 9)30/12/2004 11:05:42

intermolecular bonding - hydrogen bonds

Hydrogen bonding in alcohols


An alcohol is an organic molecule containing an -O-H group.
Any molecule which has a hydrogen atom attached directly to an oxygen
or a nitrogen is capable of hydrogen bonding. Such molecules will always
have higher boiling points than similarly sized molecules which don't
have an -O-H or an -N-H group. The hydrogen bonding makes the
molecules "stickier", and more heat is necessary to separate them.
Ethanol, CH3CH2-O-H, and methoxymethane, CH3-O-CH3, both have
the same molecular formula, C2H6O.

Note: If you haven't done any organic chemistry yet, don't


worry about the names.

http://www.chemguide.co.uk/atoms/bonding/hbond.html (6 of 9)30/12/2004 11:05:42

intermolecular bonding - hydrogen bonds

They have the same number of electrons, and a similar length to the
molecule. The van der Waals attractions (both dispersion forces and
dipole-dipole attractions) in each will be much the same.
However, ethanol has a hydrogen atom attached directly to an oxygen and that oxygen still has exactly the same two lone pairs as in a water
molecule. Hydrogen bonding can occur between ethanol molecules,
although not as effectively as in water. The hydrogen bonding is limited
by the fact that there is only one hydrogen in each ethanol molecule with
sufficient + charge.
In methoxymethane, the lone pairs on the oxygen are still there, but the
hydrogens aren't sufficiently + for hydrogen bonds to form. Except in
some rather unusual cases, the hydrogen atom has to be attached
directly to the very electronegative element for hydrogen bonding to
occur.
The boiling points of ethanol and methoxymethane show the dramatic
effect that the hydrogen bonding has on the stickiness of the ethanol
molecules:
ethanol (with hydrogen bonding)
methoxymethane (without hydrogen bonding)

78.5C
-24.8C

The hydrogen bonding in the ethanol has lifted its boiling point about 100
C.

It is important to realise that hydrogen bonding exists in addition to van


der Waals attractions. For example, all the following molecules contain
the same number of electrons, and the first two are much the same
length. The higher boiling point of the butan-1-ol is due to the additional
hydrogen bonding.

http://www.chemguide.co.uk/atoms/bonding/hbond.html (7 of 9)30/12/2004 11:05:42

intermolecular bonding - hydrogen bonds

Comparing the two alcohols (containing -OH groups), both boiling points
are high because of the additional hydrogen bonding due to the
hydrogen attached directly to the oxygen - but they aren't the same.
The boiling point of the 2-methylpropan-1-ol isn't as high as the butan-1ol because the branching in the molecule makes the van der Waals
attractions less effective than in the longer butan-1-ol.

Hydrogen bonding in organic molecules containing nitrogen


Hydrogen bonding also occurs in organic molecules containing N-H
groups - in the same sort of way that it occurs in ammonia. Examples
range from simple molecules like CH3NH2 (methylamine) to large
molecules like proteins and DNA. The two strands of the famous alphahelix in DNA are held together by hydrogen bonds involving N-H groups.

Where would you like to go now?


To look at van der Waals forces . . .
To the bonding menu . . .
To the atomic structure and bonding menu . . .
To Main Menu . . .

http://www.chemguide.co.uk/atoms/bonding/hbond.html (8 of 9)30/12/2004 11:05:42

intermolecular bonding - hydrogen bonds

Jim Clark 2000

http://www.chemguide.co.uk/atoms/bonding/hbond.html (9 of 9)30/12/2004 11:05:42

physical properties of molecular substances

MOLECULAR STRUCTURES

This page describes how the physical properties of substances having


molecular structures varies with the type of intermolecular attractions hydrogen bonding or van der Waals forces.

Important! There's not much point in reading this page


unless you are reasonably happy about the origin of
hydrogen bonding and van der Waals forces. Follow these
links first if you aren't sure about these.

The physical properties of molecular substances


Molecules are made of fixed numbers of atoms joined together by
covalent bonds, and can range from the very small (even down to single
atoms, as in the noble gases) to the very large (as in polymers, proteins
or even DNA).
The covalent bonds holding the molecules together are very strong, but
these are largely irrelevant to the physical properties of the substance.
Physical properties are governed by the intermolecular forces - forces
attracting one molecule to its neighbours - van der Waals attractions or
hydrogen bonds.
Melting and boiling points
Molecular substances tend to be gases, liquids or low melting point
solids, because the intermolecular forces of attraction are comparatively
weak. You don't have to break any covalent bonds in order to melt or boil
a molecular substance.

http://www.chemguide.co.uk/atoms/structures/molecular.html (1 of 9)30/12/2004 11:05:47

physical properties of molecular substances

Note: This is really important! You can make yourself look


extremely stupid if you imply in an exam that boiling water,
for example, splits it into hydrogen and oxygen by breaking
covalent bonds. Exactly the same water molecules are
present in ice, water and steam.

The size of the melting or boiling point will depend on the strength of the
intermolecular forces. The presence of hydrogen bonding will lift the
melting and boiling points. The larger the molecule the more van der
Waals attractions are possible - and those will also need more energy to
break.

Solubility in water
Most molecular substances are insoluble (or only very sparingly soluble)
in water. Those which do dissolve often react with the water, or else are
capable of forming hydrogen bonds with the water.
Why doesn't methane, CH4, dissolve in water?
The methane itself isn't the problem. Methane is a gas, and so its
molecules are already separate - the water doesn't need to pull them
apart from one another.
The problem is the hydrogen bonds between the water molecules. If
methane were to dissolve, it would have to force its way between water
molecules and so break hydrogen bonds. That costs a reasonable
amount of energy.
The only attractions possible between methane and water molecules are
the much weaker van der Waals forces - and not much energy is
released when these are set up. It simply isn't energetically profitable for
the methane and water to mix.
Why does ammonia, NH3, dissolve in water?
Ammonia has the ability to form hydrogen bonds. When the hydrogen
http://www.chemguide.co.uk/atoms/structures/molecular.html (2 of 9)30/12/2004 11:05:47

physical properties of molecular substances

bonds between water molecules are broken, they can be replaced by


equivalent bonds between water and ammonia molecules.
Some of the ammonia also reacts with the water to produce ammonium
ions and hydroxide ions.

The reversible arrows show that the reaction doesn't go to completion. At


any one time only about 1% of the ammonia has actually reacted to form
ammonium ions. The solubility of ammonia is mainly due to the hydrogen
bonding and not the reaction.
Other common substances which are freely soluble in water because
they can hydrogen bond with water molecules include ethanol (alcohol)
and sucrose (sugar).

Solubility in organic solvents


Molecular substances are often soluble in organic solvents - which are
themselves molecular. Both the solute (the substance which is
dissolving) and the solvent are likely to have molecules attracted to each
other by van der Waals forces. Although these attractions will be
disrupted when they mix, they are replaced by similar ones between the
two different sorts of molecules.

Electrical conductivity
Molecular substances won't conduct electricity. Even in cases where
electrons may be delocalised within a particular molecule, there isn't
sufficient contact between the molecules to allow the electrons to move
through the whole solid or liquid.

Some individual examples

http://www.chemguide.co.uk/atoms/structures/molecular.html (3 of 9)30/12/2004 11:05:47

physical properties of molecular substances

Iodine, I2
Iodine is a dark grey crystalline solid with a purple vapour. M.Pt: 114C.
B.Pt: 184C. It is very, very slightly soluble in water, but dissolves freely
in organic solvents.
Iodine is therefore a low melting point solid. The crystallinity suggests a
regular packing of the molecules.

The structure is described as face centred cubic - it is a cube of iodine


molecules with another molecule at the centre of each face.
The orientation of the iodine molecules within this structure is quite
difficult to draw (let alone remember!). If your syllabus and past exam
papers suggests that you need to remember it, look carefully at the next
sequence of diagrams showing the layers.

Note: If you haven't got a copy of your syllabus or copies of


recent past papers, follow this link to find out how to get them.

http://www.chemguide.co.uk/atoms/structures/molecular.html (4 of 9)30/12/2004 11:05:47

physical properties of molecular substances

Notice that as you look down on the cube, all the molecules on the left
and right hand sides are aligned the same way. The ones in the middle
are aligned in the opposite way.
All these diagrams show an "exploded" view of the crystal. The iodine
molecules are, of course, touching each other. Measurements of the
distances between the centres of the atoms in the crystal show two
different values:

The iodine atoms within each molecule are pulled closely together by the
covalent bond. The van der Waals attraction between the molecules is
much weaker, and you can think of the atoms in two separate molecules
as just loosely touching each other.

Ice
Ice is a good example of a hydrogen bonded solid.
There are lots of different ways that the water molecules can be arranged
in ice. This is one of them, but NOT the common one - I can't draw that in
any way that makes sense! The one below is known as "cubic ice", or
"ice Ic". It is based on the water molecules arranged in a diamond
structure.

http://www.chemguide.co.uk/atoms/structures/molecular.html (5 of 9)30/12/2004 11:05:47

physical properties of molecular substances

This is just a small part of a structure which extends over huge numbers
of molecules in three dimensions. In the diagram, the lines represent
hydrogen bonds. The lone pairs that the hydrogen atoms are attracted to
are left out for clarity.
Cubic ice is only stable at temperatures below -80C. The ice you are
familiar with has a different, hexagonal structure. It is called "ice Ih".

Note: Don't worry about this problem. If asked to draw ice in


a UK A level exam, don't try to be too clever. It is probably
best not to go beyond the top five molecules in the above
diagram. This will show the essential features of the bonding
in the structure without getting bogged down in stuff which is
far beyond this level.
If you are interested in following this up, try a Google search
(at the bottom of the Main Menu - link below) using the
search term ice structure hexagonal cubic (or something
similar). This will throw up lots of information together with an
assortment of fairly dreadful diagrams which I for one don't
have the visual imagination to unscramble!

http://www.chemguide.co.uk/atoms/structures/molecular.html (6 of 9)30/12/2004 11:05:47

physical properties of molecular substances

The unusual density behaviour of water


The hydrogen bonding forces a rather open structure on the ice - if you
made a model of it, you would find a significant amount of wasted space.
When ice melts, the structure breaks down and the molecules tend to fill
up this wasted space.
This means that the water formed takes up less space than the original
ice. Ice is a very unusual solid in this respect - most solids show an
increase in volume on melting.
When water freezes, the opposite happens - there is an expansion as the
hydrogen bonded structure establishes. Most liquids contract on freezing.
Remnants of the rigid hydrogen bonded structure are still present in very
cold liquid water, and don't finally disappear until 4C. The density of
water increases from 0C to 4C as the molecules free themselves from
the open structure and take up less space. After 4C, the thermal motion
of the molecules causes them to move apart and the density falls. That's
the normal behaviour with liquids on heating.

Note: You can find more about water (particularly its


abnormally high boiling point) in the page on hydrogen
bonding.

Polymers
Bonding in polymers
Polymers like poly(ethene) - commonly called polythene - consist of very
long molecules. Poly(ethene) molecules are made by joining up lots of
ethene molecules into chains of covalently bound carbon atoms with
hydrogens attached. There may be short branches along the main chain,
also consisting of carbon chains with attached hydrogens. The molecules
are attracted to each other in the solid by van der Waals dispersion
forces.
http://www.chemguide.co.uk/atoms/structures/molecular.html (7 of 9)30/12/2004 11:05:47

physical properties of molecular substances

By controlling the conditions under which ethene is polymerised, it is


possible to control the amount of branching to give two distinct types of
polythene.
High density polythene
High density polythene has virtually unbranched chains. The lack of
branching allows molecules to lie close together in a regular way which is
almost crystalline.
Because the molecules lie close together, dispersion forces are more
effective, and so the plastic is relatively strong and has a somewhat
higher melting point than low density polythene.
High density polythene is used for containers for household chemicals
like washing-up liquid, for example, or for bowls or buckets.
Low density polythene
Low density polythene has lots of short branches along the chain. These
branches prevent the chains from lying close together in a tidy
arrangement. As a result dispersion forces are less and the plastic is
weaker and has a lower melting point. Its density is lower, of course,
because of the wasted space within the unevenly packed structure.
Low density polythene is used for things like plastic bags.

Where would you like to go now?


To the structures menu . . .
To the atomic structure and bonding menu . . .
To Main Menu . . .

http://www.chemguide.co.uk/atoms/structures/molecular.html (8 of 9)30/12/2004 11:05:47

physical properties of molecular substances

Jim Clark 2000

http://www.chemguide.co.uk/atoms/structures/molecular.html (9 of 9)30/12/2004 11:05:47

Structures menu

Understanding Chemistry

STRUCTURES MENU

Ionic structures . . .
Takes NaCl as a typical example to show the relationship between
the physical properties of an ionic compound and its structure.
Giant covalent structures . . .
Looks at diamond, graphite and silicon dioxide and the way their
physical properties are affected by their structures.
Metallic structures . . .
Looks at the relationship between the physical properties of metals
and their structures.
Molecular structures . . .
Looks in detail at the structures of ice, iodine and poly(ethene) as
well as dealing in general with the sort of physical properties you
might expect of molecular substances.
How to decide . . .
Explains how you can decide which sort of structure a substance
has by looking at its physical properties.
Physical properties of Period 3 elements . . .
Explains how the structures of the elements from Na to Ar in the
Periodic Table affect their simple physical properties.
http://www.chemguide.co.uk/atoms/structsmenu.html (1 of 2)30/12/2004 11:05:49

Structures menu

Go to atomic structure and bonding menu . . .


Go to Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/atoms/structsmenu.html (2 of 2)30/12/2004 11:05:49

covalent bonding - double bonds

COVALENT BONDING - DOUBLE BONDS

This page explains how double covalent bonds arise. It starts with a
simple picture of double covalent bonding, and then takes a more
sophisticated view of the bonding in ethene.

Warning! This page assumes that you have already read


the page on single covalent bonds. If you have come straight
to this page via a search engine follow this link before you go
on.

A simple view of double covalent bonds


A double covalent bond is where two pairs of electrons are shared
between the atoms rather than just one pair.
Some simple molecules containing double bonds
Oxygen, O2
Two oxygen atoms can both achieve stable structures by sharing two
pairs of electrons as in the diagram.

The double bond is shown conventionally by two lines joining the atoms.
Each line represents one pair of shared electrons.
Carbon dioxide, CO2

http://www.chemguide.co.uk/atoms/bonding/doublebonds.html (1 of 6)30/12/2004 11:05:54

covalent bonding - double bonds

Ethene, C2H4
Ethene has a double bond between the two carbon atoms.

A more sophisticated view of the bonding in ethene


It is important to explore the bonding in ethene in more detail because it
has a direct impact on its chemistry. Unless you have some
understanding of the true nature of the double bond, you can't really
understand the way that ethene behaves.
An orbital view of the bonding in ethene
Ethene is built from hydrogen atoms (1s1) and carbon atoms
(1s22s22px12py1).
The carbon atom doesn't have enough unpaired electrons to form the
required number of bonds, so it needs to promote one of the 2s2 pair into
the empty 2pz orbital. This is exactly the same as happens whenever
carbon forms bonds - whatever else it ends up joined to.

http://www.chemguide.co.uk/atoms/bonding/doublebonds.html (2 of 6)30/12/2004 11:05:54

covalent bonding - double bonds

Now there's a difference, because each carbon is only joining to three


other atoms (another carbon and two hydrogens) rather than four - as,
for example, in methane.
The carbon re-organises the s orbital and two of the p orbitals to give
three new orbitals with exactly the same energy. This process is called
hybridisation. The other p orbital is left unchanged.

Note: You might wonder why it chooses to hybridise these


three orbitals rather than just use the three p orbitals which
already have the same energy. It's because it uses the
orbitals with the lowest energy first.

The new orbitals formed are called sp2


hybrids, because they are made by an s
orbital and two p orbitals reorganising
themselves. sp2 orbitals look rather like
sp3 orbitals that we discussed in the
bonding in methane in the page on
single bonds, except that they are
shorter and fatter. The three sp2 hybrid
orbitals arrange themselves as far apart as possible - which is at 120 to
each other in a plane. The remaining p orbital is at right angles to them.
The two carbon atoms and four hydrogen atoms would look like this
before they joined together:

http://www.chemguide.co.uk/atoms/bonding/doublebonds.html (3 of 6)30/12/2004 11:05:54

covalent bonding - double bonds

The various atomic orbitals which are pointing


towards each other now merge to give
molecular orbitals, each containing a bonding
pair of electrons. Molecular orbitals made by
end-to-end overlap of atomic orbitals are
called sigma bonds.

The p orbitals on each carbon aren't pointing towards each other, and so
we'll leave those for a moment. In the diagram, the black dots represent
the nuclei of the atoms.
Notice that the p orbitals are so close that they are overlapping sideways.
This sideways overlap also creates a
molecular orbital, but of a different kind. In
this one the electrons aren't held on the line
between the two nuclei, but above and below
the plane of the molecule. A bond formed in
this way is called a pi bond.

For clarity, the sigma bonds are shown using lines - each line
representing one pair of shared electrons. The various sorts of line show
the directions the bonds point in. An ordinary line represents a bond in
the plane of the screen (or the paper if you've printed it), a broken line is
a bond going back away from you, and a wedge shows a bond coming
out towards you.

http://www.chemguide.co.uk/atoms/bonding/doublebonds.html (4 of 6)30/12/2004 11:05:54

covalent bonding - double bonds

Note: The really interesting bond in ethene is the pi bond. In


almost all cases where you will draw the structure of ethene,
the sigma bonds will be shown as lines.

Be clear about what a pi bond is. It is a region of space in which you can
find the two electrons which make up the bond. Those two electrons can
live anywhere within that space. It would be quite misleading to think of
one living in the top and the other in the bottom.

Beyond A'level: This is a good example of the curious


behaviour of electrons. How do the electrons get from one
half of the pi bond to the other if they are never found in
between? It's an unanswerable question if you think of
electrons as particles. If you want to follow this up, you will
have to read some fairly high-powered stuff on the wave
nature of electrons.

Even if your syllabus doesn't expect you to know how a pi bond is


formed, it will expect you to know that it exists. The pi bond dominates
the chemistry of ethene. It is very vulnerable to attack - a very negative
region of space above and below the plane of the molecule. It is also
somewhat distant from the control of the nuclei and so is a weaker bond
than the sigma bond joining the two carbons.

Important! Check your syllabus! Find out whether you


actually need to know how a pi bond is formed. Don't forget
to look under ethene as well as in the bonding section of your
syllabus. If you don't need to know it, there's no point in
learning it! You will, however, need to know that a pi bond
exists - that the two bonds between the carbon atoms in
ethene aren't both the same.
If you haven't got a copy of your syllabus, find out how to
download one.

http://www.chemguide.co.uk/atoms/bonding/doublebonds.html (5 of 6)30/12/2004 11:05:54

covalent bonding - double bonds

All double bonds (whatever atoms they might be joining) will consist of a
sigma bond and a pi bond.
This orbital view of the double bond is only really important at A'level with
regard to organic compounds. If you want to read more about this, follow
the first link below which leads you to the menu for a section specifically
on organic bonding. You will find the description of ethene repeated, but
will also find information about the bonding in benzene and in the carbonoxygen double bond.

Where would you like to go now?


To explore more organic bonding . . .
To the bonding menu . . .
To the atomic structure and bonding menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/atoms/bonding/doublebonds.html (6 of 6)30/12/2004 11:05:54

Explaining the acidity of organic acids

THE ACIDITY OF ORGANIC ACIDS

This page explains the acidity of simple organic acids and looks at the
factors which affect their relative strengths.

Why are organic acids acidic?


Organic acids as weak acids
For the purposes of this topic, we are going to take the definition of an
acid as "a substance which donates hydrogen ions (protons) to other
things". We are going to get a measure of this by looking at how easily the
acids release hydrogen ions to water molecules when they are in solution
in water.
An acid in solution sets up this equilibrium:

Note: We are writing the acid as AH rather than HA, because,


in all the cases we shall be looking at, the hydrogen we are
interested in is at the right-hand end of a molecule.

http://www.chemguide.co.uk/basicorg/acidbase/acids.html (1 of 14)30/12/2004 11:06:05

Explaining the acidity of organic acids

A hydroxonium ion is formed together with the anion (negative ion) from
the acid.
This equilibrium is sometimes simplified by leaving out the water to
emphasise the ionisation of the acid.

If you write it like this, you must include the state symbols - "(aq)". Writing
H+(aq) implies that the hydrogen ion is attached to a water molecule as
H3O+. Hydrogen ions are always attached to something during chemical
reactions.
The organic acids are weak in the sense that this ionisation is very
incomplete. At any one time, most of the acid will be present in the
solution as un-ionised molecules. For example, in the case of dilute
ethanoic acid, the solution contains about 99% of ethanoic acid molecules
- at any instant, only about 1% have actually ionised. The position of
equilibrium therefore lies well to the left.
Comparing the strengths of weak acids
The strengths of weak acids are measured on the pKa scale. The smaller
the number on this scale, the stronger the acid is.
Three of the compounds we shall be looking at, together with their pKa
values are:

http://www.chemguide.co.uk/basicorg/acidbase/acids.html (2 of 14)30/12/2004 11:06:05

Explaining the acidity of organic acids

Remember - the smaller the number the stronger the acid. Comparing the
other two to ethanoic acid, you will see that phenol is very much weaker
with a pKa of 10.00, and ethanol is so weak with a pKa of about 16 that it
hardly counts as acidic at all!

Why are these acids acidic?


In each case, the same bond gets broken - the bond between the
hydrogen and oxygen in an -OH group. Writing the rest of the molecule as
"X":

http://www.chemguide.co.uk/basicorg/acidbase/acids.html (3 of 14)30/12/2004 11:06:05

Explaining the acidity of organic acids

Note: If you aren't sure about coordinate covalent (dative


covalent) bonding, you might like to follow this link. It isn't,
however, particularly important to the rest of the current page.
Use the BACK button on your browser to return to this page
later.

So . . . if the same bond is being broken in each case, why do these three
compounds have such widely different acid strengths?

Differences in acid strengths between carboxylic acids,


phenols and alcohols
The factors to consider
Two of the factors which influence the ionisation of an acid are:

the strength of the bond being broken,

the stability of the ions being formed.

In these cases, you seem to be breaking the same oxygen-hydrogen bond


each time, and so you might expect the strengths to be similar.

Note: You've got to be a bit careful about this. The bonds


won't be identically strong, because what's around them in the
molecule isn't the same in each case.

http://www.chemguide.co.uk/basicorg/acidbase/acids.html (4 of 14)30/12/2004 11:06:05

Explaining the acidity of organic acids

The most important factor in determining the relative acid strengths of


these molecules is the nature of the ions formed. You always get a
hydroxonium ion - so that's constant - but the nature of the anion (the
negative ion) varies markedly from case to case.

Ethanoic acid
Ethanoic acid has the structure:

The acidic hydrogen is the one attached to the oxygen. When ethanoic
acid ionises it forms the ethanoate ion, CH3COO-.
You might reasonably suppose that the structure of the ethanoate ion was
as below, but measurements of bond lengths show that the two carbonoxygen bonds are identical and somewhere in length between a single
and a double bond.

To understand why this is, you have to look in some detail at the bonding
in the ethanoate ion.

http://www.chemguide.co.uk/basicorg/acidbase/acids.html (5 of 14)30/12/2004 11:06:05

Explaining the acidity of organic acids

Warning! If you don't already understand about the bonding


in the carbon-oxygen double bond, you would be well advised
to skip this next bit - all the way down to the simplified
structure of the ethanoate ion towards the end of it. It goes
beyond anything that you are likely to want for UK A level
purposes.
If you do choose to follow this link, it will probably take you to
several other pages before you are ready to come back here
again. Use the BACK button (or HISTORY file or GO menu) on
your browser to return to this page.

Like any other double bond, a carbon-oxygen double bond is made up of


two different parts. One electron pair is found on the line between the two
nuclei - this is known as a sigma bond. The other electron pair is found
above and below the plane of the molecule in a pi bond.
Pi bonds are made by sideways overlap between p orbitals on the carbon
and the oxygen.
In an ethanoate ion, one of the lone pairs on the negative oxygen ends up
almost parallel to these p orbitals, and overlaps with them.

http://www.chemguide.co.uk/basicorg/acidbase/acids.html (6 of 14)30/12/2004 11:06:05

Explaining the acidity of organic acids

This leads to a delocalised pi system over the whole of the -COO- group,
rather like that in benzene.

All the oxygen lone pairs have been left out of this diagram to avoid
confusion.
Because the oxygens are more electronegative than the carbon, the
delocalised system is heavily distorted so that the electrons spend much
more time in the region of the oxygen atoms.
So where is the negative charge in all this? It has been spread around
over the whole of the -COO- group, but with the greatest chance of finding
it in the region of the two oxygen atoms.
Ethanoate ions can be drawn simply as:

The dotted line represents the delocalisation. The negative charge is


written centrally on that end of the molecule to show that it isn't localised
on one of the oxygen atoms.
The more you can spread charge around, the more stable an ion
becomes. In this case, if you delocalise the negative charge over several
atoms, it is going to be much less attractive to hydrogen ions - and so you
are less likely to re-form the ethanoic acid.

Phenol
Phenols have an -OH group attached directly to a benzene ring. Phenol
itself is the simplest of these with nothing else attached to the ring apart
http://www.chemguide.co.uk/basicorg/acidbase/acids.html (7 of 14)30/12/2004 11:06:05

Explaining the acidity of organic acids

from the -OH group.

When the hydrogen-oxygen bond in phenol breaks, you get a phenoxide


ion, C6H5O-.

Warning! You need to understand about the bonding in


benzene in order to make sense of this next bit.
If your syllabus says that you need to know about the acidity of
phenol, then you will have to understand the next few
paragraphs - which in turn means that you will have to
understand about benzene. If it doesn't mention phenol, skip it!
If you follow this link, you may have to explore several other
pages before you are ready to come back here again. Use the
BACK button (or HISTORY file or GO menu) on your browser
to return to this page.

Delocalisation also occurs in this ion. This time, one of the lone pairs on
the oxygen atom overlaps with the delocalised electrons on the benzene
ring.

http://www.chemguide.co.uk/basicorg/acidbase/acids.html (8 of 14)30/12/2004 11:06:05

Explaining the acidity of organic acids

This overlap leads to a delocalisation which extends from the ring out over
the oxygen atom. As a result, the negative charge is no longer entirely
localised on the oxygen, but is spread out around the whole ion.

Why then is phenol a much weaker acid than ethanoic acid?


In the ethanoate ion, the delocalised system was distorted towards the
two oxygen atoms, therefore essentially sharing the negative charge
between them.
Since each oxygen atom only has about half the charge it would have if
there wasn't any delocalisation, neither oxygen is going to be as attractive
towards hydrogen ions as it would otherwise be.
That means that the ethanoate ion won't take up a hydrogen ion as easily
as it might otherwise. If it stays ionised, the formation of the hydrogen ions
means that it is acidic.
In the phenoxide ion, the single oxygen atom is still the most
electronegative thing present, and the delocalised system will be heavily
distorted towards it. That still leaves the oxygen atom with most of its
negative charge.
What delocalisation there is makes the phenoxide ion more stable than it
would otherwise be, and so phenol is acidic to an extent.
However, the delocalisation hasn't shared the charge around very
effectively. There is still lots of negative charge around the oxygen to
which hydrogen ions will be attracted - and so the phenol will readily re-

http://www.chemguide.co.uk/basicorg/acidbase/acids.html (9 of 14)30/12/2004 11:06:05

Explaining the acidity of organic acids

form. Phenol is therefore only very weakly acidic.

Ethanol
Ethanol, CH3CH2OH, is so weakly acidic that you would hardly count it as
acidic at all. If the hydrogen-oxygen bond breaks to release a hydrogen
ion, an ethoxide ion is formed:

This has nothing at all going for it. There is no way of delocalising the
negative charge, which remains firmly on the oxygen atom. That intense
negative charge will be highly attractive towards hydrogen ions, and so
the ethanol will instantly re-form.
Since ethanol is very poor at losing hydrogen ions, it is hardly acidic at all.

Variations in acid strengths between different carboxylic


acids
You might think that all carboxylic acids would have the same strength
because each depends on the delocalisation of the negative charge
around the -COO- group to make the anion more stable, and so more
reluctant to re-combine with a hydrogen ion.
In fact, the carboxylic acids have widely different acidities. One obvious
difference is between methanoic acid, HCOOH, and the other simple
carboxylic acids:
pKa
HCOOH

http://www.chemguide.co.uk/basicorg/acidbase/acids.html (10 of 14)30/12/2004 11:06:05

3.75

Explaining the acidity of organic acids

CH3COOH

4.76

CH3CH2COOH

4.87

CH3CH2CH2COOH

4.82

Remember that the higher the value for pKa, the weaker the acid is.
Why is ethanoic acid weaker than methanoic acid? It again depends on
the stability of the anions formed - on how much it is possible to delocalise
the negative charge. The less the charge is delocalised, the less stable
the ion, and the weaker the acid.
The methanoate ion (from methanoic acid) is:

The only difference between this and the ethanoate ion is the presence of
the CH3 group in the ethanoate.
But that's important! Alkyl groups have a tendency to "push" electrons
away from themselves. That means that there will be a small amount of
extra negative charge built up on the -COO- group. Any build-up of charge
will make the ion less stable, and more attractive to hydrogen ions.
Ethanoic acid is therefore weaker than methanoic acid, because it will reform more easily from its ions.

The other alkyl groups have "electron-pushing" effects very similar to the
methyl group, and so the strengths of propanoic acid and butanoic acid

http://www.chemguide.co.uk/basicorg/acidbase/acids.html (11 of 14)30/12/2004 11:06:05

Explaining the acidity of organic acids

are very similar to ethanoic acid.

Note: If you want more information about the inductive effect


of alkyl groups, you could read about carbocations (carbonium
ions) in the mechanism section of this site.
Use the BACK button on your browser to return to this page if
you choose to follow this link.

The acids can be strengthened by pulling charge away from the -COOend. You can do this by attaching electronegative atoms like chlorine to
the chain.

As the next table shows, the more chlorines you can attach the better:
pKa
CH3COOH

4.76

CH2ClCOOH

2.86

CHCl2COOH

1.29

CCl3COOH

0.65

Trichloroethanoic acid is quite a strong acid.


Attaching different halogens also makes a difference. Fluorine is the most
electronegative and so you would expect it to be most successful at

http://www.chemguide.co.uk/basicorg/acidbase/acids.html (12 of 14)30/12/2004 11:06:05

Explaining the acidity of organic acids

pulling charge away from the -COO- end and so strengthening the acid.
pKa
CH2FCOOH

2.66

CH2ClCOOH

2.86

CH2BrCOOH

2.90

CH2ICOOH

3.17

The effect is there, but isn't as great as you might expect.


Finally, notice that the effect falls off quite quickly as the attached halogen
gets further away from the -COO- end. Here is what happens if you move
a chlorine atom along the chain in butanoic acid.
pKa
CH3CH2CH2COOH

4.82

CH3CH2CHClCOOH

2.84

CH3CHClCH2COOH

4.06

CH2ClCH2CH2COOH

4.52

The chlorine is effective at withdrawing charge when it is next-door to the COO- group, and much less so as it gets even one carbon further away.

Where would you like to go now?


To the acids and bases menu. . .
To menu of basic organic chemistry. . .
To Main Menu . . .
http://www.chemguide.co.uk/basicorg/acidbase/acids.html (13 of 14)30/12/2004 11:06:05

Explaining the acidity of organic acids

You might also be interested in:


The properties and reactions of:
carboxylic acids . . .
amino acids . . .
phenol . . .
alcohols . . .

Jim Clark 2000

http://www.chemguide.co.uk/basicorg/acidbase/acids.html (14 of 14)30/12/2004 11:06:05

bonding in carbonyl compounds - the carbon oxygen double bond

BONDING IN CARBONYL COMPOUNDS

Important! Don't try to read this unless you are sure that
you fully understand the orbital view of the bonding in
methane and the bonding in ethene.
You may also find it useful to read the article on orbitals if you
aren't sure about simple orbital theory.

The carbonyl group


The simple view of the bonding in carbon - oxygen double bonds
Where the carbon-oxygen double bond, C=O, occurs in
organic compounds it is called a carbonyl group. The
simplest compound containing this group is methanal.
We are going to look at the bonding in methanal, but it
would equally apply to any other compound containing
C=O. The interesting thing is the nature of the carbon-oxygen double
bond - not what it's attached to.

Note: Methanal is normally written as HCHO. If you wrote it


as HCOH, it looks as if it contains an -O-H group - and it
doesn't. Methanal is an aldehyde. All aldehydes contain the
CHO group.
Naming: methanal: meth counts 1 carbon atom, an means
no C=C, al says that it is an aldehyde and so contains CHO.

http://www.chemguide.co.uk/basicorg/bonding/carbonyl.html (1 of 5)30/12/2004 11:06:17

bonding in carbonyl compounds - the carbon oxygen double bond

An orbital view of the bonding in carbon - oxygen double bonds


The carbon atom
Just as in ethene or benzene, the carbon atom is joined to three other
atoms. The carbon's electrons rearrange themselves, and promotion and
hybridisation give sp2 hybrid orbitals.
Promotion gives:

Hybridisation of the 2s orbital and two of


the 2p orbitals means that the carbon
atom now looks like the diagram on the
right.
Three sp2 hybrid orbitals are formed and
these arrange themselves as far apart in
space as they can - at 120 to each
other. The remaining p orbital is at right
angles to them.
This is exactly the same as in ethene or in benzene.

Important! If this isn't really clear to you, you must go and


read the article about the bonding in ethene.

http://www.chemguide.co.uk/basicorg/bonding/carbonyl.html (2 of 5)30/12/2004 11:06:17

bonding in carbonyl compounds - the carbon oxygen double bond

The oxygen atom


Oxygen's electronic structure is 1s22s22px22py12pz1.
The 1s electrons are too deep inside the atom to be concerned with the
bonding and so we'll ignore them from now on. Hybridisation occurs in
the oxygen as well. It is easier to see this using "electrons-in-boxes".

This time two of the sp2 hybrid orbitals contain lone pairs of electrons.

Help! A "lone pair" of electrons is a pair of electrons at the


bonding level which isn't being used to bond on to another
atom.

The carbon atom and oxygen atom then bond in much the same way as
the two carbons do in ethene. In the next diagram, we are assuming that
the carbon will also bond to two hydrogens to make methanal - but it
could equally well bond to anything else.

http://www.chemguide.co.uk/basicorg/bonding/carbonyl.html (3 of 5)30/12/2004 11:06:17

bonding in carbonyl compounds - the carbon oxygen double bond

End-to-end overlap between the atomic orbitals that are pointing towards
each other produce sigma bonds.
Notice that the p orbitals are overlapping sideways.

This sideways overlap produces a pi bond. So just like C=C, C=O is


made up of a sigma bond and a pi bond.
Does that mean that the bonding is exactly the same as in ethene? No!
The distribution of electrons in the pi bond is heavily distorted towards
the oxygen end of the bond, because oxygen is much more
electronegative than carbon.

Help! You can read about the origins of electronegativity and


its effects in organic compounds in a separate article.

http://www.chemguide.co.uk/basicorg/bonding/carbonyl.html (4 of 5)30/12/2004 11:06:17

bonding in carbonyl compounds - the carbon oxygen double bond

This distortion in the pi bond causes major differences in the reactions of


compounds containing carbon-oxygen double bonds like methanal
compared with compounds containing carbon-carbon double bonds like
ethene.

Note: You can read about addition reactions or additionelimination reactions of carbonyl compounds elswhere on this
site.

Where would you like to go now?


To the organic bonding menu. . .
To menu of basic organic chemistry. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/basicorg/bonding/carbonyl.html (5 of 5)30/12/2004 11:06:17

Nucleophilic Addition Mechanisms Menu

Understanding Chemistry

NUCLEOPHILIC ADDITION MECHANISMS MENU

The addition of hydrogen cyanide to aldehydes and ketones . . .


Facts and mechanism for the nucleophilic addition of hydrogen
cyanide, HCN, to aldehydes and ketones.
The reduction of aldehydes and ketones . . .
Facts and a simplified mechanism for the reduction of aldehydes
and ketones using sodium tetrahydridoborate, NaBH4.

Go to menu of other types of mechanism. . .


Go to Main Menu . . .

You might also be interested in:


properties and reactions of aldehydes and ketones . . .
Covers all the physical and chemical properties of aldehydes and
ketones required by UK A level syllabuses.

http://www.chemguide.co.uk/mechanisms/nucaddmenu.html (1 of 2)30/12/2004 11:06:18

Nucleophilic Addition Mechanisms Menu

Jim Clark 2000 (modified 2004)

http://www.chemguide.co.uk/mechanisms/nucaddmenu.html (2 of 2)30/12/2004 11:06:18

nucleophilic addition - carbonyl compounds and hydrogen cyanide

THE NUCLEOPHILIC ADDITION OF HYDROGEN


CYANIDE TO ALDEHYDES AND KETONES

This page gives you the facts and simple, uncluttered mechanisms for
the nucleophilic addition reactions between carbonyl compounds
(specifically aldehydes and ketones) and hydrogen cyanide, HCN. If you
want the mechanisms explained to you in detail, there is a link at the
bottom of the page.
Aldehydes and ketones behave identically in their reaction with hydrogen
cyanide, and so will be considered together - although equations and
mechanisms will be given for both types of compounds for the sake of
completeness.

The reaction of aldehydes and ketones with hydrogen


cyanide
The facts
Hydrogen cyanide adds across the carbon-oxygen double bond in
aldehydes and ketones to produce compounds known as hydroxynitriles.
For example, with ethanal (an aldehyde) you get 2-hydroxypropanenitrile:

With propanone (a ketone) you get 2-hydroxy-2-methylpropanenitrile:

http://www.chemguide.co.uk/mechanisms/nucadd/hcn.html (1 of 6)30/12/2004 11:06:28

nucleophilic addition - carbonyl compounds and hydrogen cyanide

Note: When you are naming these compounds, don't forget


that the longest carbon chain has to include the carbon in the
-CN group. In both of the above examples, the longest
carbon chain is 3 carbons - hence the "prop" in both names.
The carbon with the nitrogen attached is always counted as
the number 1 carbon in the chain.

The reaction isn't normally done using hydrogen cyanide itself, because
this is an extremely poisonous gas. Instead, the aldehyde or ketone is
mixed with a solution of sodium or potassium cyanide in water to which a
little sulphuric acid has been added. The pH of the solution is adjusted to
about 4 - 5, because this gives the fastest reaction.
The solution will contain hydrogen cyanide (from the reaction between
the sodium or potassium cyanide and the sulphuric acid), but still
contains some free cyanide ions. This is important for the mechanism.

Note: If the reaction is done using hydrogen cyanide itself, a


little sodium hydroxide solution is added to produce some
cyanide ions from the weakly acidic HCN. Again the pH of the
solution is adjusted to around pH 5 - in other words, the
sodium hydroxide is not added to excess. The rate of the
reaction falls if the pH is any higher.
Whichever set of reagents you use, the reaction contains the
same mixture of hydrogen cyanide and cyanide ions.

The mechanisms
These are examples of nucleophilic addition.
The carbon-oxygen double bond is highly polar, and
the slightly positive carbon atom is attacked by the
cyanide ion acting as a nucleophile.

http://www.chemguide.co.uk/mechanisms/nucadd/hcn.html (2 of 6)30/12/2004 11:06:28

nucleophilic addition - carbonyl compounds and hydrogen cyanide

Nucleophile: A species (molecule or ion) which attacks a


positive site in something else. Nucleophiles are either fully
negative ions or contain a fairly negative region somewhere
in a molecule. All nucleophiles have at least one active lone
pair of electrons. When you write mechanisms for reactions
involving nucleophiles, you must show that lone pair.

The mechanism for the addition of HCN to propanone


In the first stage, there is a nucleophilic attack by the cyanide ion on the
slightly positive carbon atom.

The negative ion formed then picks up a hydrogen ion from somewhere for example, from a hydrogen cyanide molecule.

The hydrogen ion could also come from the water or the H3O+ ions
present in the slightly acidic solution. You don't need to remember all of
these. One equation is perfectly adequate.

Note: The product molecule here has been drawn differently


from the one in the equation further up this page. It has been
rotated through 90. There is no reason why you can't do that
if it makes the appearance of the mechanism easier to follow.

http://www.chemguide.co.uk/mechanisms/nucadd/hcn.html (3 of 6)30/12/2004 11:06:28

nucleophilic addition - carbonyl compounds and hydrogen cyanide

The mechanism for the addition of HCN to ethanal


As before, the reaction starts with a nucleophilic attack by the cyanide
ion on the slightly positive carbon atom.

It is completed by the addition of a hydrogen ion from, for example, a


hydrogen cyanide molecule.

Note: Again, the product molecule looks different from the


one in the equation further up this page. The central carbon
atom still has the same four groups attached, but to make the
mechanism easier to follow, they are simply arranged
differently. That's not a problem - we're still talking about the
same substance.

Optical isomerism in 2-hydroxypropanenitrile


When 2-hydroxypropanenitrile is made in this last mechanism, it occurs
as a racemic mixture - a 50/50 mixture of two optical isomers. It is
possible that you might be exected to explain how this arises.

http://www.chemguide.co.uk/mechanisms/nucadd/hcn.html (4 of 6)30/12/2004 11:06:28

nucleophilic addition - carbonyl compounds and hydrogen cyanide

Note: You almost certainly won't be able to tell whether or


not you need to know this from the syllabus. You need to
refer to recent exam papers and mark schemes. If you
haven't already got these, you can obtain them from your
Exam Board via links on the syllabuses page.

Optical isomerism occurs in compounds which have four different groups


attached to a single carbon atom. In this case, the product molecule
contains a CH3, a CN, an H and an OH all attached to the central carbon
atom.
The reason for the formation of equal amounts of
two isomers lies in the way the ethanal gets
attacked.
Ethanal is a planar molecule, and attack by a
cyanide ion will either be from above the plane of the molecule, or from
below. There is an equal chance of either happening.

Attack from one side will lead to one of the two isomers, and attack from
the other side will lead to the other.

Note: This is probably as much as you need to know for


exam purposes, but a full explanation of this is given on the
"talk through" page. Follow the link below.

http://www.chemguide.co.uk/mechanisms/nucadd/hcn.html (5 of 6)30/12/2004 11:06:28

nucleophilic addition - carbonyl compounds and hydrogen cyanide

All aldehydes will form a racemic mixture in this way. Unsymmetrical


ketones will as well. (A ketone can be unsymmetrical in the sense that
there is a different alkyl group either side of the carbonyl group.) What
matters is that the product molecule must have four different groups
attached to the carbon which was originally part of the carbon-oxygen
double bond.

Where would you like to go now?


Help! Talk me through these mechanisms . . .
To menu of nucleophilic addition reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/nucadd/hcn.html (6 of 6)30/12/2004 11:06:28

Explaining the nucleophilic addition of hydrogen cyanide to carbonyl compounds

EXPLAINING THE NUCLEOPHILIC ADDITION OF


HYDROGEN CYANIDE TO ALDEHYDES AND
KETONES
This page explains the mechanism for the nucleophilic addition reaction
between carbonyl compounds (specifically aldehydes and ketones) and
hydrogen cyanide. It also looks in some detail at why a racemic mixture
is formed when hydrogen cyanide reacts with an aldehyde like ethanal.

Background
Bonding in the carbonyl group: the carbon-oxygen double bond
Oxygen is far more electronegative than carbon and
so has a strong tendency to pull electrons in a
carbon-oxygen bond towards itself. One of the two
pairs of electrons that make up a carbon-oxygen
double bond is even more easily pulled towards the oxygen. That makes
the carbon-oxygen double bond very highly polar.

Note: If you aren't sure about electronegativity and bond


polarity follow this link before you go on.
If you are interested in really understanding the bonding in
the carbon-oxygen double bond you could explore it in detail
by following this link. You don't need to know about this to
understand the rest of this page. You might find that following
this link will take you some time, because you will probably
have to explore several pages of background material as well.

http://www.chemguide.co.uk/mechanisms/nucadd/hcntt.html (1 of 9)30/12/2004 11:06:33

Explaining the nucleophilic addition of hydrogen cyanide to carbonyl compounds

The cyanide ion as a nucleophile


A nucleophile is a species (either a negatively charged ion or a negative
region in a polar molecule) which is attracted to a positive site in another
substance. All nucleophiles contain an active lone pair of electrons.
The cyanide ion comes from hydrogen cyanide, which is a covalent
molecule. Hydrogen cyanide is very weakly acidic, which means that it
can lose a hydrogen ion - although not very easily.

Notice that when the hydrogen is lost, it leaves its electron behind on the
carbon. That leaves a lone pair of electrons on that carbon, together with
a negative charge. It's essential to realise that in the cyanide ion the
active lone pair and the charge are on the carbon atom and not the
nitrogen.

Note: There is also a lone pair on the nitrogen atom (not


shown to avoid confusion), but that isn't important because
that's not where the negative charge is.

Explaining the conditions for the reaction


Remember that the reaction is done by reacting the aldehyde or ketone
with a solution of sodium or potassium cyanide to which enough dilute
sulphuric acid is added to give a pH of around 4 - 5.
Hydrogen cyanide as a very weak acid
The initial attack on the carbonyl group is by a cyanide ion. Although the
reaction overall adds hydrogen cyanide across the double bond, using
hydrogen cyanide as the reagent isn't successful because hydrogen
cyanide is such a weak acid. It produces very, very few hydrogen ions
and cyanide ions in solution. The number of cyanide ions available to
attack the slightly positive carbon is extremely small and so the reaction
would be very slow.
http://www.chemguide.co.uk/mechanisms/nucadd/hcntt.html (2 of 9)30/12/2004 11:06:33

Explaining the nucleophilic addition of hydrogen cyanide to carbonyl compounds

Why is the potassium cyanide acidified slightly?


The presence of an acid in solution helps to strengthen the polarity of the
carbon-oxygen double bond. The electrons in that bond are strongly
attracted towards the hydrogen ions in the solution.

Why isn't a lot more acid added to give a really low pH?
You might think that the more acid you added, the more you could
increase the polarity of the carbon-oxygen double bond. Unfortunately,
there is a competing effect.
The more acid you add, the more the cyanide ions get converted into
hydrogen cyanide. Since cyanide ions are what actually attacks the
slightly positive carbon, removing them isn't helpful!
A pH of 4 - 5 is found experimentally to give you the best rate of reaction.
It increases the polarity of the double bond by a useful amount, but
without removing too many of the cyanide ions as HCN.

The mechanisms
Mechanisms for the addition of hydrogen cyanide to aldehydes and
ketones are given separately for completeness, but there is no absolutely
no difference between them. They are examples of nucleophilic
addition.
The mechanism for the addition of HCN to propanone
As the cyanide ion approaches the slightly positive carbon atom, the lone

http://www.chemguide.co.uk/mechanisms/nucadd/hcntt.html (3 of 9)30/12/2004 11:06:33

Explaining the nucleophilic addition of hydrogen cyanide to carbonyl compounds

pair of electrons is attracted towards the carbon and forms a bond with it.
At the same time the two electrons in one of the bonds joining the carbon
to the oxygen are repelled until they end up entirely on the oxygen giving it a negative charge.

Note: If you aren't happy about the use of curly arrows in


mechanisms, follow this link before you go on. Use the BACK
button on your browser to return to this page.

The negative ion formed then picks up a hydrogen ion from somewhere for example, from a hydrogen cyanide molecule.

The hydrogen ion could also come from the water or the H3O+ ions (often
written as H+(aq)) present in the slightly acidic solution. You don't need to
remember all of these. One equation is perfectly adequate. In fact, most
sources show this final stage as a reaction with just H+.

http://www.chemguide.co.uk/mechanisms/nucadd/hcntt.html (4 of 9)30/12/2004 11:06:33

Explaining the nucleophilic addition of hydrogen cyanide to carbonyl compounds

Note: Although most people (including examiners) probably


take this short cut, it doesn't mean that it's right. It is sloppy
because it suggests the possibility of free hydrogen ions in a
chemical reaction - hydrogen ions are simply protons, and
are always attached to something else. It also denies you the
satisfaction of writing an equation which shows the
production of a new cyanide ion which can go on to react with
another molecule of the carbonyl compound.
If you can't be bothered to do it properly, then do at least
write the hydrogen ion as H+(aq) - not just as H+.

The mechanism for the addition of HCN to ethanal


Exactly as before, as the cyanide ion approaches the slightly positive
carbon atom, the lone pair of electrons is attracted towards the carbon
and forms a bond with it. At the same time the two electrons in one of the
bonds joining the carbon to the oxygen are repelled until they end up
entirely on the oxygen - giving it a negative charge.

The negative ion formed then picks up a hydrogen ion from somewhere for example, from a hydrogen cyanide molecule.

Once again, most sources show this final stage as a reaction with just H
http://www.chemguide.co.uk/mechanisms/nucadd/hcntt.html (5 of 9)30/12/2004 11:06:33

Explaining the nucleophilic addition of hydrogen cyanide to carbonyl compounds

+.

If you must do it that way, then write the hydrogen ion as H+(aq) - not

just as H+.

Optical isomerism in 2-hydroxypropanenitrile


2-hydroxypropanenitrile is the name of the product when ethanal reacts
with hydrogen cyanide. It is formed in this reaction as an exactly equal
mixture of two optical isomers, known as a racemic mixture.

Note: For a full discussion of optical isomerism follow this


link. Use the BACK button on your browser to return to this
page.

Optical isomerism occurs in compounds which have four different groups


attached to a single carbon atom. In this case, the product molecule
contains a CH3, a CN, an H and an OH all attached to the central carbon
atom.
The reason for the formation of equal amounts of
two isomers lies in the way the ethanal gets
attacked.
Ethanal is a planar molecule, and attack by a
cyanide ion will either be from above the plane of the molecule, or from
below. There is an equal chance of either happening.

Attack from above will lead to one of the two isomers, and attack from
below will lead to the other.
http://www.chemguide.co.uk/mechanisms/nucadd/hcntt.html (6 of 9)30/12/2004 11:06:33

Explaining the nucleophilic addition of hydrogen cyanide to carbonyl compounds

The next diagram shows what happens if attack is from above the plane
of the molecule. Notice that the existing groups get forced down away
from the approaching cyanide ion. The diagram shows the final product
after it has gained a hydrogen ion.

Attack from below forces the existing groups upwards.

Why are these two product molecules different from each other?
Everything seems to be attached in the same way, but look what
happens if you rotate the second molecule in space so that the cyanide
group is at the top.

Now compare that with the molecule formed by attack from above.

http://www.chemguide.co.uk/mechanisms/nucadd/hcntt.html (7 of 9)30/12/2004 11:06:33

Explaining the nucleophilic addition of hydrogen cyanide to carbonyl compounds

They aren't the same! Although the CN and H are lined up the same way,
the CH3 and OH are reversed. There is no way that you can rotate one
molecule in space to make it look the same as the other one. These are
therefore isomers.
The relationship between them is that they are mirror images of each
other. If one isomer were to look in a mirror, what it would see would be
the other one. Optical isomers are described as "non-superimposable
mirror images".
Because there is an equal chance of the attack coming from above or
below the plane of the molecule, then you will get equal amounts of the
two isomers formed - a racemic mixture.
This argument applies to all aldehydes, and to ketones as long as they
are unsymmetric - with a different alkyl group either side of the carbonyl
group.
A symmetric ketone like propanone, CH3COCH3, will only produce a
single product - not a mixture of isomers. The product doesn't have four
different groups around the central carbon atom, and so won't have
optical isomers. If you followed the above argument through with
propanone, you would find that you ended up with two molecules which
could be rotated in space so that they were identical. Convince yourself
by trying it!

Where would you like to go now?


To menu of nucleophilic addition reactions. . .
To menu of other types of mechanism. . .
http://www.chemguide.co.uk/mechanisms/nucadd/hcntt.html (8 of 9)30/12/2004 11:06:33

Explaining the nucleophilic addition of hydrogen cyanide to carbonyl compounds

To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/nucadd/hcntt.html (9 of 9)30/12/2004 11:06:33

optical isomerism

STEREOISOMERISM - OPTICAL ISOMERISM

Optical isomerism is a form of stereoisomerism. This page explains what


stereoisomers are and how you recognise the possibility of optical
isomers in a molecule.

What is stereoisomerism?
What are isomers?
Isomers are molecules that have the same molecular formula, but have a
different arrangement of the atoms in space. That excludes any different
arrangements which are simply due to the molecule rotating as a whole,
or rotating about particular bonds.
Where the atoms making up the various isomers are joined up in a
different order, this is known as structural isomerism. Structural
isomerism is not a form of stereoisomerism, and is dealt with on a
separate page.

Note: If you aren't sure about structural isomerism, it might


be worth reading about it before you go on with this page.

http://www.chemguide.co.uk/basicorg/isomerism/optical.html (1 of 11)30/12/2004 11:06:42

optical isomerism

What are stereoisomers?


In stereoisomerism, the atoms making up the isomers are joined up in
the same order, but still manage to have a different spatial arrangement.
Optical isomerism is one form of stereoisomerism.

Optical isomerism
Why optical isomers?
Optical isomers are named like this because of their effect on plane
polarised light.

Help! If you don't understand about plane polarised light,


follow this link before you go on with this page.

Simple substances which show optical isomerism exist as two isomers


known as enantiomers.

A solution of one enantiomer rotates the plane of polarisation in a


clockwise direction. This enantiomer is known as the (+) form.
For example, one of the optical isomers (enantiomers) of the
amino acid alanine is known as (+)alanine.

A solution of the other enantiomer rotates the plane of polarisation


in an anti-clockwise direction. This enantiomer is known as the (-)
form. So the other enantiomer of alanine is known as or (-)alanine.
If the solutions are equally concentrated the amount of rotation
caused by the two isomers is exactly the same - but in opposite
directions.
When optically active substances are made in the lab, they often
occur as a 50/50 mixture of the two enantiomers. This is known as

http://www.chemguide.co.uk/basicorg/isomerism/optical.html (2 of 11)30/12/2004 11:06:42

optical isomerism

a racemic mixture or racemate. It has no effect on plane


polarised light.

Note: One of the worrying things about optical isomerism is


the number of obscure words that suddenly get thrown at
you. Bear with it - things are soon going to get more visual!
There is an alternative way of describing the (+) and (-) forms
which is potentially very confusing. This involves the use of
the lowercase letters d- and l-, standing for dextrorotatory
and laevorotatory respectively. Unfortunately, there is
another different use of the capital letters D- and L- in this
topic. This is totally confusing! Stick with (+) and (-).

How optical isomers arise


The examples of organic optical isomers required at A' level all contain a
carbon atom joined to four different groups. These two models each have
the same groups joined to the central carbon atom, but still manage to be
different:

Obviously as they are drawn, the orange and blue groups aren't aligned
the same way. Could you get them to align by rotating one of the
molecules? The next diagram shows what happens if you rotate
molecule B.

http://www.chemguide.co.uk/basicorg/isomerism/optical.html (3 of 11)30/12/2004 11:06:42

optical isomerism

They still aren't the same - and there is no way that you can rotate them
so that they look exactly the same. These are isomers of each other.
They are described as being non-superimposable in the sense that (if
you imagine molecule B being turned into a ghostly version of itself) you
couldn't slide one molecule exactly over the other one. Something would
always be pointing in the wrong direction.

Note: Unless your visual imagination is reasonably good,


this is much easier to understand if you have actually got
some models to play with. If your school or college hasn't
given you the opportunity to play around with molecular
models, you might consider getting hold of a cheap set. The
models made by molymod are both cheap and easy to use.
An introductory organic set is more than adequate. Find them
at www.molymod.com.
Alternatively , get hold of some coloured Plasticene and
some used matches and make your own.

http://www.chemguide.co.uk/basicorg/isomerism/optical.html (4 of 11)30/12/2004 11:06:42

optical isomerism

What happens if two of the groups attached to the central carbon atom
are the same? The next diagram shows this possibility.

The two models are aligned exactly as before, but the orange group has
been replaced by another pink one.
Rotating molecule B this time shows that it is exactly the same as
molecule A. You only get optical isomers if all four groups attached to the
central carbon are different.

Chiral and achiral molecules


The essential difference between the two examples we've looked at lies
in the symmetry of the molecules.
If there are two groups the same attached to the central carbon atom, the
molecule has a plane of symmetry. If you imagine slicing through the
molecule, the left-hand side is an exact reflection of the right-hand side.
Where there are four groups attached, there is no symmetry anywhere in
the molecule.

http://www.chemguide.co.uk/basicorg/isomerism/optical.html (5 of 11)30/12/2004 11:06:42

optical isomerism

A molecule which has no plane of symmetry is described as chiral. The


carbon atom with the four different groups attached which causes this
lack of symmetry is described as a chiral centre or as an asymmetric
carbon atom.
The molecule on the left above (with a plane of symmetry) is described
as achiral.
Only chiral molecules have optical isomers.

The relationship between the enantiomers


One of the enantiomers is simply a non-superimposable mirror image
of the other one.
In other words, if one isomer looked in a mirror, what it would see is the
other one. The two isomers (the original one and its mirror image) have a
different spatial arrangement, and so can't be superimposed on each
other.

If an achiral molecule (one with a plane of symmetry) looked in a mirror,


http://www.chemguide.co.uk/basicorg/isomerism/optical.html (6 of 11)30/12/2004 11:06:42

optical isomerism

you would always find that by rotating the image in space, you could
make the two look identical. It would be possible to superimpose the
original molecule and its mirror image.

Some real examples of optical isomers


Butan-2-ol
The asymmetric carbon atom in a compound (the one with four different
groups attached) is often shown by a star.

It's extremely important to draw the isomers correctly. Draw one of them
using standard bond notation to show the 3-dimensional arrangement
around the asymmetric carbon atom. Then draw the mirror to show the
examiner that you know what you are doing, and then the mirror image.

Help! If you don't understand this bond notation, follow this


link to drawing organic molecules before you go on with this
page.

http://www.chemguide.co.uk/basicorg/isomerism/optical.html (7 of 11)30/12/2004 11:06:42

optical isomerism

Notice that you don't literally draw the mirror images of all the letters and
numbers! It is, however, quite useful to reverse large groups - look, for
example, at the ethyl group at the top of the diagram.
It doesn't matter in the least in what order you draw the four groups
around the central carbon. As long as your mirror image is drawn
accurately, you will automatically have drawn the two isomers.
So which of these two isomers is (+)butan-2-ol and which is (-)butan-2ol? There is no simple way of telling that. For A'level purposes, you can
just ignore that problem - all you need to be able to do is to draw the two
isomers correctly.
2-hydroxypropanoic acid (lactic acid)
Once again the chiral centre is shown by a star.

The two enantiomers are:

It is important this time to draw the COOH group backwards in the mirror
image. If you don't there is a good chance of you joining it on to the
central carbon wrongly.

http://www.chemguide.co.uk/basicorg/isomerism/optical.html (8 of 11)30/12/2004 11:06:42

optical isomerism

If you draw it like this in an exam, you won't get the mark for that isomer
even if you have drawn everything else perfectly.
2-aminopropanoic acid (alanine)
This is typical of naturally-occurring amino acids. Structurally, it is just
like the last example, except that the -OH group is replaced by -NH2

The two enantiomers are:

Only one of these isomers occurs naturally: the (+) form. You can't tell
just by looking at the structures which this is.
It has, however, been possible to work out which of these structures is
which. Naturally occurring alanine is the right-hand structure, and the
way the groups are arranged around the central carbon atom is known
as an L- configuration. Notice the use of the capital L. The other
configuration is known as D-.
http://www.chemguide.co.uk/basicorg/isomerism/optical.html (9 of 11)30/12/2004 11:06:42

optical isomerism

So you may well find alanine described as L-(+)alanine.


That means that it has this particular structure and rotates the plane of
polarisation clockwise.
Even if you know that a different compound has an arrangement of
groups similar to alanine, you still can't say which way it will rotate the
plane of polarisation.
The other amino acids, for example, have the same arrangement of
groups as alanine does (all that changes is the CH3 group), but some are
(+) forms and others are (-) forms.
It's quite common for natural systems to only work with one of the
enantiomers of an optically active substance. It isn't too difficult to see
why that might be. Because the molecules have different spatial
arrangements of their various groups, only one of them is likely to fit
properly into the active sites on the enzymes they work with.
In the lab, it is quite common to produce equal amounts of both forms of
a compound when it is synthesised. This happens just by chance, and
you tend to get racemic mixtures.

Note: For a detailed discussion of this, you could have a


look at the page on the addition of HCN to aldehydes

Where would you like to go now?


To the isomerism menu. . .
To menu of basic organic chemistry. . .
To Main Menu . . .

http://www.chemguide.co.uk/basicorg/isomerism/optical.html (10 of 11)30/12/2004 11:06:42

optical isomerism

Jim Clark 2000 (modified 2004)

http://www.chemguide.co.uk/basicorg/isomerism/optical.html (11 of 11)30/12/2004 11:06:42

structural isomerism

STRUCTURAL ISOMERISM

This page explains what structural isomerism is, and looks at some of the
various ways that structural isomers can arise.

What is structural isomerism?


What are isomers?
Isomers are molecules that have the same molecular formula, but have a
different arrangement of the atoms in space. That excludes any different
arrangements which are simply due to the molecule rotating as a whole,
or rotating about particular bonds.
For example, both of the following are the same molecule. They are not
isomers. Both are butane.

There are also endless other possible ways that this molecule could twist
itself. There is completely free rotation around all the carbon-carbon
single bonds.

http://www.chemguide.co.uk/basicorg/isomerism/structural.html (1 of 7)30/12/2004 11:06:49

structural isomerism

Note: Isomerism is much easier to understand if you have


actually got some models to play with. If your school or
college hasn't given you the opportunity to play around with
molecular models in the early stages of your organic
chemistry course, you might consider getting hold of a cheap
set. The models made by molymod are both cheap and easy
to use. An introductory organic set is more than adequate.
Find them at www.molymod.com.
Alternatively , get hold of some coloured Plasticene and
some used matches and make your own.

If you had a model of a molecule in front of you, you would have to take it
to pieces and rebuild it if you wanted to make an isomer of that molecule.
If you can make an apparently different molecule just by rotating single
bonds, it's not different - it's still the same molecule.

Note: It's really important that you understand this. If you


aren't sure, then you must get hold of (or make) some
models.

What are structural isomers?


In structural isomerism, the atoms are arranged in a completely different
order. This is easier to see with specific examples.
What follows looks at some of the ways that structural isomers can arise.
The names of the various forms of structural isomerism probably don't
matter all that much, but you must be aware of the different possibilities
when you come to draw isomers.

Types of structural isomerism

http://www.chemguide.co.uk/basicorg/isomerism/structural.html (2 of 7)30/12/2004 11:06:49

structural isomerism

Chain isomerism
These isomers arise because of the possibility of branching in carbon
chains. For example, there are two isomers of butane, C4H10. In one of
them, the carbon atoms lie in a "straight chain" whereas in the other the
chain is branched.

Note: Although the chain is drawn as straight, in reality it's


anything but straight. If you aren't happy about the ways of
drawing organic molecules, follow this link.
Use the BACK button on your browser to return to this page.

Be careful not to draw "false" isomers which are just


twisted versions of the original molecule. For example,
this structure is just the straight chain version of butane rotated about the
central carbon-carbon bond.
You could easily see this with a model. This is the example we've already
used at the top of this page.

http://www.chemguide.co.uk/basicorg/isomerism/structural.html (3 of 7)30/12/2004 11:06:49

structural isomerism

Pentane, C5H12, has three chain isomers. If you think you can find any
others, they are simply twisted versions of the ones below. If in doubt
make some models.

Position isomerism
In position isomerism, the basic carbon skeleton remains unchanged, but
important groups are moved around on that skeleton.
For example, there are two structural isomers with the molecular formula
C3H7Br. In one of them the bromine atom is on the end of the chain,
whereas in the other it's attached in the middle.

If you made a model, there is no way that you could twist one molecule
to turn it into the other one. You would have to break the bromine off the
end and re-attach it in the middle. At the same time, you would have to
move a hydrogen from the middle to the end.
Another similar example occurs in alcohols such as C4H9OH

http://www.chemguide.co.uk/basicorg/isomerism/structural.html (4 of 7)30/12/2004 11:06:49

structural isomerism

These are the only two possibilities provided you keep to a four carbon
chain, but there is no reason why you should do that. You can easily
have a mixture of chain isomerism and position isomerism - you aren't
restricted to one or the other.
So two other isomers of butanol are:

Note: It's essential if you are asked to draw isomers in an


exam not to restrict yourself to chain isomers or position
isomers. You must be aware of all the possibilities.

You can also get position isomers on benzene rings. Consider the
molecular formula C7H8Cl. There are four different isomers you could
make depending on the position of the chlorine atom. In one case it is
attached to the side-group carbon atom, and then there are three other
possible positions it could have around the ring - next to the CH3 group,
next-but-one to the CH3 group, or opposite the CH3 group.

Functional group isomerism


http://www.chemguide.co.uk/basicorg/isomerism/structural.html (5 of 7)30/12/2004 11:06:49

structural isomerism

In this variety of structural isomerism, the isomers contain different


functional groups - that is, they belong to different families of compounds
(different homologous series).
For example, a molecular formula C3H6O could be either propanal (an
aldehyde) or propanone (a ketone).

There are other possibilities as well for this same molecular formula - for
example, you could have a carbon-carbon double bond (an alkene) and
an -OH group (an alcohol) in the same molecule.

Another common example is illustrated by the molecular formula


C3H6O2. Amongst the several structural isomers of this are propanoic
acid (a carboxylic acid) and methyl ethanoate (an ester).

http://www.chemguide.co.uk/basicorg/isomerism/structural.html (6 of 7)30/12/2004 11:06:49

structural isomerism

Note: To repeat the warning given earlier: If you are asked


to draw the structural isomers from a given molecular
formula, don't forget to think about all the possibilities. Can
you branch the carbon chain? Can you move a group around
on that chain? Is it possible to make more than one type of
compound?
Be careful though! If you are asked to draw the structures of
esters with the molecular formula C3H6O2, you aren't going to
get a lot of credit for drawing propanoic acid, even if it is a
valid isomer.

Where would you like to go now?


To the isomerism menu. . .
To menu of basic organic chemistry. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/basicorg/isomerism/structural.html (7 of 7)30/12/2004 11:06:49

Isomerism Menu

Understanding Chemistry

ISOMERISM MENU

Structural isomerism . . .
Explains what structural isomerism is, together with examples of
the various ways that these isomers can arise.
Geometric isomerism . . .
Explains how geometric (cis/trans) isomerism arises in simple
organic compounds containing carbon-carbon double bonds.
Optical isomerism . . .
Explains how to recognise optical isomerism in simple cases, and
how to draw the isomers.

Go to menu of basic organic chemistry. . .


Go to Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/basicorg/isomermenu.html30/12/2004 11:06:51

geometric (cis / trans) isomerism

STEREOISOMERISM - GEOMETRIC ISOMERISM

Geometric isomerism (also known as cis / trans isomerism) is a form of


stereoisomerism. This page explains what stereoisomers are and how
you recognise the possibility of geometric isomers in a molecule.

What is stereoisomerism?
What are isomers?
Isomers are molecules that have the same molecular formula, but have a
different arrangement of the atoms in space. That excludes any different
arrangements which are simply due to the molecule rotating as a whole,
or rotating about particular bonds.
Where the atoms making up the various isomers are joined up in a
different order, this is known as structural isomerism. Structural
isomerism is not a form of stereoisomerism, and is dealt with on a
separate page.

Note: If you aren't sure about structural isomerism, it might


be worth reading about it before you go on with this page.

http://www.chemguide.co.uk/basicorg/isomerism/geometric.html (1 of 11)30/12/2004 11:06:57

geometric (cis / trans) isomerism

What are stereoisomers?


In stereoisomerism, the atoms making up the isomers are joined up in
the same order, but still manage to have a different spatial arrangement.
Geometric isomerism is one form of stereoisomerism.

Geometric (cis / trans) isomerism


How geometric isomers arise
These isomers occur where you have restricted rotation somewhere in a
molecule. The only examples of this required at A' level involve the
carbon-carbon double bond.
Think about what happens in molecules where there is unrestricted
rotation about carbon bonds - in other words where the carbon-carbon
bonds are all single. The next diagram shows two possible configurations
of 1,2-dichloroethane.

These two models represent exactly the same molecule. You can get
from one to the other just by twisting around the carbon-carbon single
bond. These molecules are not isomers.
If you draw a structural formula instead of using models, you have to
bear in mind the possibility of this free rotation about single bonds. You
must accept that these two structures represent the same molecule:

http://www.chemguide.co.uk/basicorg/isomerism/geometric.html (2 of 11)30/12/2004 11:06:57

geometric (cis / trans) isomerism

But what happens if you have a carbon-carbon double bond - as in 1,2dichloroethene?

These two molecules aren't the same. The carbon-carbon double bond
won't rotate and so you would have to take the models to pieces in order
to convert one structure into the other one. That is a simple test for
isomers. If you have to take a model to pieces to convert it into another
one, then you've got isomers. If you merely have to twist it a bit, then you
haven't!

Note: In the model, the reason that you can't rotate a carboncarbon double bond is that there are two links joining the
carbons together. In reality, the reason is that you would
have to break the pi bond. Pi bonds are formed by the
sideways overlap between p orbitals. If you tried to rotate the
carbon-carbon bond, the p orbitals won't line up any more
and so the pi bond is disrupted. This costs energy and only
happens if the compound is heated strongly.
If you are interested in the bonding in carbon-carbon double
bonds, follow this link. Be warned, though, that you might
have to read several pages of background material and it
could all take a long time. It isn't necessary for understanding
the rest of this page.

http://www.chemguide.co.uk/basicorg/isomerism/geometric.html (3 of 11)30/12/2004 11:06:57

geometric (cis / trans) isomerism

Drawing structural formulae for the last pair of models gives two possible
isomers.
In one, the two chlorine atoms are locked on opposite sides of the double
bond. This is known as the trans isomer. (trans : from latin meaning
"across" - as in transatlantic).
In the other, the two chlorine atoms are locked on the same side of the
double bond. This is know as the cis isomer. (cis : from latin meaning "on
this side")

The most likely example of geometric isomerism you will meet at A' level
is but-2-ene. In one case, the CH3 groups are on opposite sides of the
double bond, and in the other case they are on the same side.

The importance of drawing geometric isomers properly


It's very easy to miss geometric isomers in exams if you take short-cuts
in drawing the structural formulae. For example, it is very tempting to
draw but-2-ene as
CH3CH=CHCH3

http://www.chemguide.co.uk/basicorg/isomerism/geometric.html (4 of 11)30/12/2004 11:06:57

geometric (cis / trans) isomerism

If you write it like this, you will almost certainly miss the fact that there are
geometric isomers. If there is even the slightest hint in a question that
isomers might be involved, always draw compounds containing carboncarbon double bonds showing the correct bond angles (120 ) around the
carbon atoms at the ends of the bond. In other words, use the format
shown in the last diagrams above.

How to recognise the possibility of geometric isomerism


You obviously need to have restricted rotation somewhere in the
molecule. For A' level purposes, that means a carbon-carbon double
bond. If you have a carbon-carbon double bond, then think carefully
about the possibility of geometric isomers.
What needs to be attached to the carbon-carbon double bond?

Note: This is much easier to understand if you have actually


got some models to play with. If your school or college hasn't
given you the opportunity to play around with molecular
models in the early stages of your organic chemistry course,
you might consider getting hold of a cheap set. The models
made by molymod are both cheap and easy to use. An
introductory organic set is more than adequate. Find them at
www.molymod.com.
Alternatively , get hold of some coloured Plasticene and
some used matches and make your own.

http://www.chemguide.co.uk/basicorg/isomerism/geometric.html (5 of 11)30/12/2004 11:06:57

geometric (cis / trans) isomerism

Think about this case:

Although we've swapped the right-hand groups around, these are still the
same molecule. To get from one to the other, all you would have to do is
to turn the whole model over.
You won't have geometric isomers if there are two groups the same on
one end of the bond - in this case, the two pink groups on the left-hand
end.
So . . . there must be two different groups on the left-hand carbon and
two different groups on the right-hand one. The cases we've been
exploring earlier are like this:

But you could make things even more different and still have geometric
isomers:

Here, the blue and green groups are either on the same side of the bond
http://www.chemguide.co.uk/basicorg/isomerism/geometric.html (6 of 11)30/12/2004 11:06:57

geometric (cis / trans) isomerism

or the opposite side.


Or you could go the whole hog and make everything different. You still
get geometric isomers, but this time the words cis and trans are
meaningless. Don't worry about this - you would never have to name a
complex case like this in an exam.

Summary
To get geometric isomers you must have:

restricted rotation (involving a carbon-carbon double bond for A'


level purposes);
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.

Note: You might not need this next bit if you are meeting
geometric isomerism for the first time. Don't get too bogged
down in it if you don't have to!

http://www.chemguide.co.uk/basicorg/isomerism/geometric.html (7 of 11)30/12/2004 11:06:57

geometric (cis / trans) isomerism

The effect of geometric isomerism on physical properties


The table shows the melting point and boiling point of the cis and trans
isomers of 1,2-dichloroethene.

isomer

melting point (C)

boiling point (C)

cis

-80

60

trans

-50

48

In each case, the higher melting or boiling point is shown in red.


You will notice that:

the trans isomer has the higher melting point;

the cis isomer has the higher boiling point.

This is common. You can see the same effect with the cis and trans
isomers of but-2-ene:

isomer

melting point (C)

boiling point (C)

cis-but-2-ene

-139

trans-but-2-ene

-106

Why is the boiling point of the cis isomers higher?


There must be stronger intermolecular forces between the molecules of
the cis isomers than between trans isomers.
Taking 1,2-dichloroethene as an example:
http://www.chemguide.co.uk/basicorg/isomerism/geometric.html (8 of 11)30/12/2004 11:06:57

geometric (cis / trans) isomerism

Both of the isomers have exactly the same atoms joined up in exactly the
same order. That means that the van der Waals dispersion forces
between the molecules will be identical in both cases.
The difference between the two is that the cis isomer is a polar molecule
whereas the trans isomer is non-polar.

Note: If you aren't sure about intermolecular forces (and also


about bond polarity), it is essential that you follow this link
before you go on. You need to know about van der Waals
dispersion forces and dipole-dipole interactions, and to follow
the link on that page to another about bond polarity if you
need to.
Use the BACK button on your browser to return to this page.

Both molecules contain polar chlorine-carbon bonds, but in the cis


isomer they are both on the same side of the molecule. That means that
one side of the molecule will have a slight negative charge while the
other is slightly positive. The molecule is therefore polar.

Because of this, there will be dipole-dipole interactions as well as


dispersion forces - needing extra energy to break. That will raise the
boiling point.
A similar thing happens where there are CH3 groups attached to the
carbon-carbon double bond, as in cis-but-2-ene.
Alkyl groups like methyl groups tend to "push" electrons away from
themselves. You again get a polar molecule, although with a reversed
polarity from the first example.
http://www.chemguide.co.uk/basicorg/isomerism/geometric.html (9 of 11)30/12/2004 11:06:57

geometric (cis / trans) isomerism

Note: The term "electron pushing" is only to help remember


what happens. The alkyl group doesn't literally "push" the
electrons away - the other end of the bond attracts them
more strongly. The arrows with the cross on (representing the
more positive end of the bond) are a conventional way of
showing this electron pushing effect.

By contrast, although there will still be polar bonds in the trans isomers,
overall the molecules are non-polar.

The slight charge on the top of the molecule (as drawn) is exactly
balanced by an equivalent charge on the bottom. The slight charge on
the left of the molecule is exactly balanced by the same charge on the
right.
This lack of overall polarity means that the only intermolecular attractions
these molecules experience are van der Waals dispersion forces. Less
energy is needed to separate them, and so their boiling points are lower.

Why is the melting point of the cis isomers lower?


http://www.chemguide.co.uk/basicorg/isomerism/geometric.html (10 of 11)30/12/2004 11:06:57

geometric (cis / trans) isomerism

You might have thought that the same argument would lead to a higher
melting point for cis isomers as well, but there is another important factor
operating.
In order for the intermolecular forces to work well, the molecules must be
able to pack together efficiently in the solid.
Trans isomers pack better than cis isomers. The "U" shape of the cis
isomer doesn't pack as well as the straighter shape of the trans isomer.
The poorer packing in the cis isomers means that the intermolecular
forces aren't as effective as they should be and so less energy is needed
to melt the molecule - a lower melting point.

Where would you like to go now?


To the isomerism menu. . .
To menu of basic organic chemistry. . .
To Main Menu . . .

Jim Clark 2000 (modified 2004)

http://www.chemguide.co.uk/basicorg/isomerism/geometric.html (11 of 11)30/12/2004 11:06:57

plane polarised light

PLANE POLARISED LIGHT

This page gives a simple explanation of what plane polarised light is and
the way it is affected by optically active compounds.

A simple analogy - "plane polarised string"


Imagine tying a piece of thick string to a hook in a wall, and then shaking
the string vigorously. The string will be vibrating in all possible directions
- up-and-down, side-to-side, and all the directions in-between - giving it a
really complex overall motion.

Now, suppose you passed the string through a vertical slit. The string is a
really snug fit in the slit. The only vibrations still happening the other side
of the slit will be vertical ones. All the others will have been prevented by
the slit.

http://www.chemguide.co.uk/basicorg/isomerism/polarised.html (1 of 6)30/12/2004 11:07:02

plane polarised light

What emerges from the slit could be described as "plane polarised


string", because the vibrations are only in a single (vertical) plane.
Now look at the possibility of putting a second slit on the string. If it is
aligned the same way as the first one, the vibrations will still get through.

But if the second slit is at 90 to the first one, the string will stop vibrating
entirely to the right of the second slit. The second slit will only let through
horizontal vibrations - and there aren't any.

http://www.chemguide.co.uk/basicorg/isomerism/polarised.html (2 of 6)30/12/2004 11:07:02

plane polarised light

The real thing - plane polarised light


Light is also made up of vibrations - this time, electromagnetic ones.
Some materials have the ability to screen out all the vibrations apart from
those in one plane and so produce plane polarised light.
The most familiar example of this is the material that Polaroid sunglasses
are made of. If you wear one pair of Polaroid sunglasses and hold
another pair up in front of them so that the glasses are held vertically
rather than horizontally, you'll find that no light gets through - you will just
see darkness. This is equivalent to the two slits at right angles in the
string analogy. The polaroids are described as being "crossed".

Care! It is important not to take the analogy too far. The


polaroid material doesn't consist of "slits" in any sense of the
word. The way it actually polarises the light is quite different
(and irrelevant to us here!).

Optically active substances


An optically active substance is one which can rotate the plane of
polarisation of plane polarised light. if you shine a beam of polarised
monochromatic light (light of only a single frequency - in other words a
single colour) through a solution of an optically active substance, when
the light emerges, its plane of polarisation is found to have rotated.
The rotation may be either clockwise or anti-clockwise. Assuming the
original plane of polarisation was vertical, you might get either of these
results.

http://www.chemguide.co.uk/basicorg/isomerism/polarised.html (3 of 6)30/12/2004 11:07:02

plane polarised light

How can you tell that the plane of polarisation has been rotated?
You use a polarimeter.

The polariser and analyser are both made of polaroid material.


The polarimeter is originally set up with water in the tube. Water isn't
optically active - it has no effect on the plane of polarisation. The
analyser is rotated until you can't see any light coming through the
instrument. The polaroids are then "crossed".

http://www.chemguide.co.uk/basicorg/isomerism/polarised.html (4 of 6)30/12/2004 11:07:02

plane polarised light

Now you put a solution of an optically active substance into the tube. It
rotates the plane of polarisation of the light, and so the analyser won't be
at right-angles to it any longer and some light will get through. You would
have to rotate the analyser in order to cut the light off again.

You can easily tell whether the plane of polarisation has been rotated
clockwise or anti-clockwise, and by how much.

http://www.chemguide.co.uk/basicorg/isomerism/polarised.html (5 of 6)30/12/2004 11:07:02

plane polarised light

Note: It is very unlikely that you will need to learn much of


this. You won't be able to tell from your syllabus whether any
of this is required. Look instead at past exam papers. If you
haven't already got these, you can get them from your Exam
Board by following the appropriate link on the syllabuses
page.
Probably all you need to be able to do is to understand the
expression "rotates the plane of polarisation of plane
polarised light" so that you can use it sensibly.

Where would you like to go now?


Back to optical isomerism. . .
To the isomerism menu. . .
To menu of basic organic chemistry. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/basicorg/isomerism/polarised.html (6 of 6)30/12/2004 11:07:02

reduction of carbonyl compounds using sodium tetrahydridoborate

THE REDUCTION OF ALDEHYDES AND KETONES

This page gives you the facts and mechanisms for the reduction of
carbonyl compounds (specifically aldehydes and ketones) using sodium
tetrahydridoborate (sodium borohydride) as the reducing agent.
Only one UK A level Exam Board (AQA) is likely to ask for these
mechanisms, and they are happy with a simplified version of what is quite
a complex mechanism. Because of that simplification, these reactions are
dealt with entirely on this page - without the "talk through" page that you
will find for other mechanisms on this site.

The reduction of aldehydes and ketones by sodium


tetrahydridoborate
The facts
Sodium tetrahydridoborate (previously known as sodium borohydride) has
the formula NaBH4, and contains the BH4- ion. That ion acts as the
reducing agent.
There are several quite different ways of carrying out this reaction. Two
possible variants (there are several others!) are:

The reaction is carried out in solution in water to which some


sodium hydroxide has been added to make it alkaline. The reaction
produces an intermediate which is converted into the final product
by addition of a dilute acid like sulphuric acid.
The reaction is carried out in solution in an alcohol like methanol,
ethanol or propan-2-ol. This produces an intermediate which can be
converted into the final product by boiling it with water.

In each case, reduction essentially involves the addition of a hydrogen


atom to each end of the carbon-oxygen double bond to form an alcohol.
Reduction of aldehydes and ketones lead to two different sorts of alcohol.
http://www.chemguide.co.uk/mechanisms/nucadd/reduce.html (1 of 6)30/12/2004 11:07:07

reduction of carbonyl compounds using sodium tetrahydridoborate

The reduction of an aldehyde


For example, with ethanal you get ethanol:

Notice that this is a simplified equation - perfectly acceptable to


examiners. The H in square brackets means "hydrogen from a reducing
agent".
In general terms, reduction of an aldehyde leads to a primary alcohol. A
primary alcohol is one which only has one alkyl group attached to the
carbon with the -OH group on it. They all contain the grouping -CH2OH.

Note: There is one exception to this. Methanol CH3OH is also


a primary alcohol. Think of this as H-CH2OH.

The reduction of a ketone


For example, with propanone you get propan-2-ol:

Reduction of a ketone leads to a secondary alcohol. A secondary


alcohol is one which has two alkyl groups attached to the carbon with the OH group on it. They all contain the grouping -CHOH.

http://www.chemguide.co.uk/mechanisms/nucadd/reduce.html (2 of 6)30/12/2004 11:07:07

reduction of carbonyl compounds using sodium tetrahydridoborate

Beware! The following mechanisms are simplified for UK A


level purposes to the point that they are wrong! If you are
working outside the UK A level system, please don't read any
further!

The simplified mechanisms


The BH4- ion is essentially a source of hydride ions, H-. The simplification
used is to write H- instead of BH4-.
Doing this not only makes the initial attack easier to write, but avoids you
getting involved with some quite complicated boron compounds that are
formed as intermediates.
The reduction is an example of nucleophilic
addition.
The carbon-oxygen double bond is highly polar, and
the slightly positive carbon atom is attacked by the
hydride ion acting as a nucleophile. A hydride ion is a hydrogen atom with
an extra electron - hence the lone pair.

Nucleophile: A species (molecule or ion) which attacks a


positive site in something else. Nucleophiles are either fully
negative ions or contain a fairly negative region somewhere in
a molecule. All nucleophiles have at least one active lone pair
of electrons. When you write mechanisms for reactions
involving nucleophiles, you must show that lone pair.

http://www.chemguide.co.uk/mechanisms/nucadd/reduce.html (3 of 6)30/12/2004 11:07:07

reduction of carbonyl compounds using sodium tetrahydridoborate

The mechanism for the reduction of ethanal


In the first stage, there is a nucleophilic attack by the hydride ion on the
slightly positive carbon atom. The lone pair of electrons on the hydride ion
forms a bond with the carbon, and the electrons in one of the carbonoxygen bonds are repelled entirely onto the oxygen, giving it a negative
charge.

What happens now depends on whether you add an acid or water to


complete the reaction.
Adding an acid:
When the acid is added, the negative ion formed picks up a hydrogen ion
to give an alcohol.

Note: You may find that other sources write the hydrogen ion
simply as H+. That's not good practice, because it suggests a
free hydrogen ion. The hydrogen ion is actually attached to a
water molecule as H3O+. Writing that makes the equation look
more complicated. H+(aq) is a happy compromise.

http://www.chemguide.co.uk/mechanisms/nucadd/reduce.html (4 of 6)30/12/2004 11:07:07

reduction of carbonyl compounds using sodium tetrahydridoborate

Adding water:
This time, the negative ion takes a hydrogen ion from a water molecule.

The mechanism for the reduction of propanone


As before, the reaction starts with a nucleophilic attack by the hydride ion
on the slightly positive carbon atom.

Again, what happens next depends on whether you add an acid or water
to complete the reaction.
Adding an acid:
The negative ion reacts with a hydrogen ion from the acid added in the
second stage of the reaction.

Adding water:
This time, the negative ion takes a hydrogen ion from a water molecule.

http://www.chemguide.co.uk/mechanisms/nucadd/reduce.html (5 of 6)30/12/2004 11:07:07

reduction of carbonyl compounds using sodium tetrahydridoborate

Important!
Remember that the equations and mechanisms given on this page are not
the truth - they are merely simplifications to suit the demands of a
particular A' level syllabus.

Where would you like to go now?


To menu of nucleophilic addition reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000 (modified 2004)

http://www.chemguide.co.uk/mechanisms/nucadd/reduce.html (6 of 6)30/12/2004 11:07:07

Aldehydes and Ketones Menu

Understanding Chemistry

ALDEHYDES AND KETONES MENU

Background . . .
An introduction to aldehydes and ketones, their reactivity and their
physical properties.
Preparation of aldehydes and ketones . . .
Their preparation by the oxidation of primary or secondary
alcohols.
Simple addition reactions . . .
The addition of hydrogen cyanide and of sodium hydrogensulphite
(sodium bisulphite) to aldehydes and ketones.
Reduction of aldehydes and ketones . . .
The reduction of aldehydes and ketones using sodium
tetrahydridoborate(III) or lithium tetrahydridoaluminate(III) (sodium
borohydride or lithium aluminium hydride).
Reactions with Grignard reagents . . .
The reactions of aldehydes and ketones with Grignard reagents as
a way of making complicated alcohols.
Oxidation of aldehydes and ketones . . .
Covers the main ways of distinguishing between aldehydes and
ketones using, for example, Tollens' reagent, Fehling's solution or
http://www.chemguide.co.uk/organicprops/carbonylmenu.html (1 of 2)30/12/2004 11:07:08

Aldehydes and Ketones Menu

Benedict's solution.
Addition-elimination reactions . . .
Looks at the test for aldehydes and ketones using 2,4dinitrophenylhydrazine (Brady's reagent), plus a quick look at
some similar reactions.
The triiodomethane (iodoform) reaction . . .
The use of this reaction to detect the presence of the CH3CO
group in aldehydes and ketones.

Go to menu of other organic compounds . . .


Go to Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/carbonylmenu.html (2 of 2)30/12/2004 11:07:08

an introduction to aldehydes and ketones

INTRODUCING ALDEHYDES AND KETONES

This page explains what aldehydes and ketones are, and looks at the
way their bonding affects their reactivity. It also considers their simple
physical properties such as solubility and boiling points. Details of the
chemical reactions of aldehydes and ketones are described on separate
pages.

What are aldehydes and ketones?


Aldehydes and ketones as carbonyl compounds
Aldehydes and ketones are simple compounds which contain a carbonyl
group - a carbon-oxygen double bond. They are simple in the sense that
they don't have other reactive groups like -OH or -Cl attached directly to
the carbon atom in the carbonyl group - as you might find, for example, in
carboxylic acids containing -COOH.
Examples of aldehydes
In aldehydes, the carbonyl group has a hydrogen atom attached to it
together with either

a second hydrogen atom


or, more commonly, a hydrocarbon group which might be an alkyl
group or one containing a benzene ring.

For the purposes of this section, we shall ignore those containing


benzene rings.

http://www.chemguide.co.uk/organicprops/carbonyls/background.html (1 of 8)30/12/2004 11:07:12

an introduction to aldehydes and ketones

Note: There is no very significant reason for this. It is just


that if you are fairly new to organic chemistry you might not
have come across any compounds with benzene rings in
them yet. I'm just trying to avoid adding to your confusion!

Notice that these all have exactly the same end to the molecule. All that
differs is the complexity of the other group attached.
When you are writing formulae for these, the aldehyde group (the
carbonyl group with the hydrogen atom attached) is always written as CHO - never as COH. That could easily be confused with an alcohol.
Ethanal, for example, is written as CH3CHO; methanal as HCHO.
The name counts the total number of carbon atoms in the longest chain including the one in the carbonyl group. If you have side groups attached
to the chain, notice that you always count from the carbon atom in the
carbonyl group as being number 1.
Note: If you aren't confident about naming organic
compounds, then you really ought to follow this link before
you go on.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/carbonyls/background.html (2 of 8)30/12/2004 11:07:12

an introduction to aldehydes and ketones

Examples of ketones
In ketones, the carbonyl group has two hydrocarbon groups
attached. Again, these can be either alkyl groups or ones containing
benzene rings. Again, we'll concentrated on those containing alkyl
groups just to keep things simple.
Notice that ketones never have a hydrogen atom attached to the
carbonyl group.

Propanone is normally written CH3COCH3. Notice the need for


numbering in the longer ketones. In pentanone, the carbonyl group could
be in the middle of the chain or next to the end - giving either pentan-3one or pentan-2-one.

Bonding and reactivity


Bonding in the carbonyl group
Oxygen is far more electronegative than carbon and
so has a strong tendency to pull electrons in a
carbon-oxygen bond towards itself. One of the two
pairs of electrons that make up a carbon-oxygen
double bond is even more easily pulled towards the oxygen. That makes
the carbon-oxygen double bond very highly polar.

http://www.chemguide.co.uk/organicprops/carbonyls/background.html (3 of 8)30/12/2004 11:07:12

an introduction to aldehydes and ketones

Note: If you aren't sure about electronegativity and bond


polarity follow this link before you go on.
If you are interested in really understanding the bonding in
the carbon-oxygen double bond you could explore it in detail
by following this link. You might find that following this link will
take you some time, because you will probably have to
explore several pages of background material as well.
Use the BACK button (or HISTORY file or GO menu) on your
browser to return to this page.

Important reactions of the carbonyl group


The slightly positive carbon atom in the carbonyl group can be attacked
by nucleophiles. A nucleophile is a negatively charged ion (for example,
a cyanide ion, CN-), or a slightly negatively charged part of a molecule
(for example, the lone pair on a nitrogen atom in ammonia, NH3).
During the reaction, the carbon-oxygen double bond gets broken. The
net effect of all this is that the carbonyl group undergoes addition
reactions, often followed by the loss of a water molecule. This gives a
reaction known as addition-elimination or condensation. You will find
examples of simple addition reactions and addition-elimination if you
explore the aldehydes and ketones menu (link at the bottom of the page).
Both aldehydes and ketones contain a carbonyl group. That means that
their reactions are very similar in this respect.
Where aldehydes and ketones differ
An aldehyde differs from a ketone by having a hydrogen atom attached
to the carbonyl group. This makes the aldehydes very easy to oxidise.
For example, ethanal, CH3CHO, is very easily oxidised to either ethanoic
acid, CH3COOH, or ethanoate ions, CH3COO-.

http://www.chemguide.co.uk/organicprops/carbonyls/background.html (4 of 8)30/12/2004 11:07:12

an introduction to aldehydes and ketones

Ketones don't have that hydrogen atom and are resistant to oxidation.
They are only oxidised by powerful oxidising agents which have the
ability to break carbon-carbon bonds.
You will find the oxidation of aldehydes and ketones discussed if you
follow a link from the aldehydes and ketones menu (see the bottom of
this page).

Physical properties
Boiling points
Methanal is a gas (boiling point -21C), and ethanal has a boiling point of
+21C. That means that ethanal boils at close to room temperature.
The other aldehydes and the ketones are liquids, with boiling points
rising as the molecules get bigger.
The size of the boiling point is governed by the strengths of the
intermolecular forces.
van der Waals dispersion forces
These attractions get stronger as the molecules get longer and have
more electrons. That increases the sizes of the temporary dipoles that
are set up. This is why the boiling points increase as the number of
carbon atoms in the chains increases - irrespective of whether you are
talking about aldehydes or ketones.

Note: If you aren't happy about intermolecular forces


(particularly van der Waals dispersion forces and dipoledipole interactions) then you really ought to follow this link
before you go on.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/carbonyls/background.html (5 of 8)30/12/2004 11:07:12

an introduction to aldehydes and ketones

van der Waals dipole-dipole attractions


Both aldehydes and ketones are polar molecules because of the
presence of the carbon-oxygen double bond. As well as the dispersion
forces, there will also be attractions between the permanent dipoles on
nearby molecules.
That means that the boiling points will be higher than those of similarly
sized hydrocarbons - which only have dispersion forces.
It is interesting to compare three similarly sized molecules. They have
similar lengths, and similar (although not identical) numbers of electrons.

molecule

type

boiling point (C)

CH3CH2CH3

alkane

-42

CH3CHO

aldehyde

+21

CH3CH2OH

alcohol

+78

Notice that the aldehyde (with dipole-dipole attractions as well as


dispersion forces) has a boiling point higher than the similarly sized
alkane which only has dispersion forces.
However, the aldehyde's boiling point isn't as high as the alcohol's. In the
alcohol, there is hydrogen bonding as well as the other two kinds of
intermolecular attraction.
Although the aldehydes and ketones are highly polar molecules, they
don't have any hydrogen atoms attached directly to the oxygen, and so
they can't hydrogen bond with each other.

http://www.chemguide.co.uk/organicprops/carbonyls/background.html (6 of 8)30/12/2004 11:07:12

an introduction to aldehydes and ketones

Note: If you aren't sure about hydrogen bonding then it


would be a good idea to follow this link before you go on.
Use the BACK button on your browser to return to this page.

Solubility in water
The small aldehydes and ketones are freely soluble in water but solubility
falls with chain length. For example, methanal, ethanal and propanone the common small aldehydes and ketones - are miscible with water in all
proportions.
The reason for the solubility is that although aldehydes and ketones can't
hydrogen bond with themselves, they can hydrogen bond with water
molecules.
One of the slightly positive hydrogen atoms in a water molecule can be
sufficiently attracted to one of the lone pairs on the oxygen atom of an
aldehyde or ketone for a hydrogen bond to be formed.

There will also, of course, be dispersion forces and dipole-dipole


attractions between the aldehyde or ketone and the water molecules.
Forming these attractions releases energy which helps to supply the
energy needed to separate the water molecules and aldehyde or ketone
molecules from each other before they can mix together.
As chain lengths increase, the hydrocarbon "tails" of the molecules (all
the hydrocarbon bits apart from the carbonyl group) start to get in the
http://www.chemguide.co.uk/organicprops/carbonyls/background.html (7 of 8)30/12/2004 11:07:12

an introduction to aldehydes and ketones

way.
By forcing themselves between water molecules, they break the
relatively strong hydrogen bonds between water molecules without
replacing them by anything as good. This makes the process
energetically less profitable, and so solubility decreases.

Where would you like to go now?


To the aldehydes and ketones menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/carbonyls/background.html (8 of 8)30/12/2004 11:07:12

Understanding the names of organic compounds

THE NAMES OF ORGANIC COMPOUNDS

This page explains how to write the formula for an organic compound
given its name - and vice versa. It covers alkanes, cycloalkanes, alkenes,
simple compounds containing halogens, alcohols, aldehydes and
ketones. At the bottom of the page, you will find links to other types of
compound.

Background
How this page is going to tackle the problem
There are two skills you have to develop in this area:

You need to be able to translate the name of an organic


compound into its structural formula.
You need to be able to name a compound from its given formula.

The first of these is more important (and also easier!) than the second. In
an exam, if you can't write a formula for a given compound, you aren't
going to know what the examiner is talking about and could lose lots of
marks. However, you might only be asked to write a name for a given
formula once in a whole exam - in which case you only risk 1 mark.
So, we're going to look mainly at how you decode names and turn them
into formulae. In the process you will also pick up tips about how to
produce names yourself.
In the early stages of an organic chemistry course people frequently get
confused and daunted by the names because they try to do too much at
once. Don't try to read all these pages in one go. Just go as far as the
compounds you are interested in at the moment and ignore the rest.
Come back to them as they arise during the natural flow of your course.
Cracking the code

http://www.chemguide.co.uk/basicorg/conventions/names.html (1 of 14)30/12/2004 11:07:30

Understanding the names of organic compounds

A modern organic name is simply a code. Each part of the name gives
you some useful information about the compound.
For example, to understand the name 2-methylpropan-1-ol you need to
take the name to pieces.
The prop in the middle tells you how many carbon atoms there are in the
longest chain (in this case, 3). The an which follows the "prop" tells you
that there aren't any carbon-carbon double bonds.
The other two parts of the name tell you about interesting things which
are happening on the first and second carbon atom in the chain. Any
name you are likely to come across can be broken up in this same way.
Counting the carbon atoms
You will need to remember the codes for the number of carbon atoms in
a chain up to 6 carbons. There is no easy way around this - you have got
to learn them. If you don't do this properly, you won't be able to name
anything!
code

no of carbons

meth

eth

prop

but

pent

hex

Types of carbon-carbon bonds


Whether or not the compound contains a carbon-carbon double bond is
shown by the two letters immediately after the code for the chain length.
code

means

http://www.chemguide.co.uk/basicorg/conventions/names.html (2 of 14)30/12/2004 11:07:30

Understanding the names of organic compounds

an

only carbon-carbon single bonds

en

contains a carbon-carbon double


bond

For example, butane means four carbons in a chain with no double bond.
Propene means three carbons in a chain with a double bond between
two of the carbons.
Alkyl groups
Compounds like methane, CH4, and ethane, CH3CH3, are members of a
family of compounds called alkanes. If you remove a hydrogen atom
from one of these you get an alkyl group.
For example:

A methyl group is CH3.

An ethyl group is CH3CH2.

These groups must, of course, always be attached to something else.

Types of compounds
The alkanes
Example 1: Write the structural formula for 2-methylpentane.
Start decoding the name from the bit that counts the number of carbon
atoms in the longest chain - pent counts 5 carbons.
Are there any carbon-carbon double bonds? No - an tells you there aren't
any.
Now draw this carbon skeleton:
http://www.chemguide.co.uk/basicorg/conventions/names.html (3 of 14)30/12/2004 11:07:30

Understanding the names of organic compounds

Put a methyl group on the number 2 carbon atom:

Does it matter which end you start counting from? No - if you counted
from the other end, you would draw the next structure. That's exactly the
same as the first one, except that it has been flipped over.

Finally, all you have to do is to put in the correct number of hydrogen


atoms on each carbon so that each carbon is forming four bonds.

If you had to name this yourself:

Count the longest chain of carbons that you can find. Don't
assume that you have necessarily drawn that chain horizontally. 5
carbons means pent.
Are there any carbon-carbon double bonds? No - therefore
pentane.
There's a methyl group on the number 2 carbon - therefore 2methylpentane. Why the number 2 as opposed to the number 4
carbon? In other words, why do we choose to number from this
particluar end? The convention is that you number from the end
which produces the lowest numbers in the name - hence 2- rather
than 4-.

Example 2: Write the structural formula for 2,3-dimethylbutane.

http://www.chemguide.co.uk/basicorg/conventions/names.html (4 of 14)30/12/2004 11:07:30

Understanding the names of organic compounds

Start with the carbon backbone. There are 4 carbons in the longest chain
(but) with no carbon-carbon double bonds (an).

This time there are two methyl groups (di) on the number 2 and number
3 carbon atoms.

Completing the formula by filling in the hydrogen atoms gives:

Note: Does it matter whether you draw the two methyl


groups one up and one down, or both up, or both down? Not
in the least! If you aren't sure about drawing organic
molecules, follow this link before you go on. Use the BACK
button on your browser to return to this page.

Example 3: Write the structural formula for 2,2-dimethylbutane.


This is exactly like the last example, except that both methyl groups are
on the same carbon atom. Notice that the name shows this by using 2,2as well as di. The structure is worked out as before:

http://www.chemguide.co.uk/basicorg/conventions/names.html (5 of 14)30/12/2004 11:07:30

Understanding the names of organic compounds

Example 4: Write the structural formula for 3-ethyl-2-methylhexane.


hexan shows a 6 carbon chain with no carbon-carbon double bonds.

This time there are two different alkyl groups attached - an ethyl group on
the number 3 carbon atom and a methyl group on number 2.

Filling in the hydrogen atoms gives:

Note: Once again it doesn't matter whether the ethyl and


methyl groups point up or down. You might also have chosen
to start numbering from the right-hand end of the chain.
These would all be perfectly valid structures. All you would
have done is to rotate the whole molecule in space, or rotate
it around particular bonds.
If you aren't sure about this, then you must read about
drawing organic molecules before you go on.

http://www.chemguide.co.uk/basicorg/conventions/names.html (6 of 14)30/12/2004 11:07:30

Understanding the names of organic compounds

If you had to name this yourself:


How do you know what order to write the different alkyl groups at the
beginning of the name? The convention is that you write them in
alphabetical order - hence ethyl comes before methyl which in turn
comes before propyl.

The cycloalkanes
In a cycloalkane the carbon atoms are joined up in a ring - hence cyclo.
Example: Write the structural formula for cyclohexane.
hexan shows 6 carbons with no carbon-carbon double bonds. cyclo
shows that they are in a ring. Drawing the ring and putting in the correct
number of hydrogens to satisfy the bonding requirements of the carbons
gives:

The alkenes
Example 1: Write the structural formula for propene.
prop counts 3 carbon atoms in the longest chain. en tells you that there
is a carbon-carbon double bond. That means that the carbon skeleton
looks like this:

Putting in the hydrogens gives you:

http://www.chemguide.co.uk/basicorg/conventions/names.html (7 of 14)30/12/2004 11:07:30

Understanding the names of organic compounds

Example 2: Write the structural formula for but-1-ene.


but counts 4 carbon atoms in the longest chain and en tells you that
there is a carbon-carbon double bond. The number in the name tells you
where the double bond starts.
No number was necessary in the propene example above because the
double bond has to start on one of the end carbon atoms. In the case of
butene, though, the double bond could either be at the end of the chain
or in the middle - and so the name has to code for the its position.
The carbon skeleton is:

And the full structure is:

Incidentally, you might equally well have decided that the right-hand
carbon was the number 1 carbon, and drawn the structure as:

Example 3: Write the structural formula for 3-methylhex-2-ene.


The longest chain has got 6 carbon atoms (hex) with a double bond
starting on the second one (-2-en).
But this time there is a methyl group attached to the chain on the number
3 carbon atom, giving you the underlying structure:

Adding the hydrogens gives the final structure:

http://www.chemguide.co.uk/basicorg/conventions/names.html (8 of 14)30/12/2004 11:07:30

Understanding the names of organic compounds

Be very careful to count the bonds around each carbon atom when you
put the hydrogens in. It would be very easy this time to make the mistake
of writing an H after the third carbon - but that would give that carbon a
total of 5 bonds.

Compounds containing halogens


Example 1: Write the structural formula for 1,1,1-trichloroethane.
This is a two carbon chain (eth) with no double bonds (an). There are
three chlorine atoms all on the first carbon atom.

Example 2: Write the structural formula for 2-bromo-2-methylpropane.


First sort out the carbon skeleton. It's a three carbon chain with no
double bonds and a methyl group on the second carbon atom.

Draw the bromine atom which is also on the second carbon.

And finally put the hydrogen atoms in.

http://www.chemguide.co.uk/basicorg/conventions/names.html (9 of 14)30/12/2004 11:07:30

Understanding the names of organic compounds

If you had to name this yourself:


Notice that the whole of the hydrocarbon part of the name is written
together - as methylpropane - before you start adding anything else on to
the name.
Example 2: Write the structural formula for 1-iodo-3-methylpent-2-ene.
This time the longest chain has 5 carbons (pent), but has a double bond
starting on the number 2 carbon. There is also a methyl group on the
number 3 carbon.

Now draw the iodine on the number 1 carbon.

Giving a final structure:

Note: You could equally well draw this molecule the other
way round, but normally where you have, say, 1-bromosomething, you tend to write the bromine (or other halogen)
on the right-hand end of the structure.

http://www.chemguide.co.uk/basicorg/conventions/names.html (10 of 14)30/12/2004 11:07:30

Understanding the names of organic compounds

Alcohols
All alcohols contain an -OH group. This is shown in a name by the
ending ol.
Example 1: Write the structural formula for methanol.
This is a one carbon chain with no carbon-carbon double bond
(obviously!). The ol ending shows it's an alcohol and so contains an -OH
group.

Example 2: Write the structural formula for 2-methylpropan-1-ol.


The carbon skeleton is a 3 carbon chain with no carbon-carbon double
bonds, but a methyl group on the number 2 carbon.

The -OH group is attached to the number 1 carbon.

The structure is therefore:

Example 3: Write the structural formula for ethane-1,2-diol.


This is a two carbon chain with no double bond. The diol shows 2 -OH
groups, one on each carbon atom.

http://www.chemguide.co.uk/basicorg/conventions/names.html (11 of 14)30/12/2004 11:07:30

Understanding the names of organic compounds

Note: There's no particular significance in the fact that this


formula has the carbon chain drawn vertically. If you draw it
horizontally, unless you stretch the carbon-carbon bond a lot,
the -OH groups look very squashed together. Drawing it
vertically makes it look tidier!

Aldehydes
All aldehydes contain the group:

If you are going to write this in a condensed form, you write it as -CHO never as -COH, because that looks like an alcohol.
The names of aldehydes end in al.
Example 1: Write the structural formula for propanal.
This is a 3 carbon chain with no carbon-carbon double bonds. The al
ending shows the presence of the -CHO group. The carbon in that group
counts as one of the chain.

Example 2: Write the structural formula for 2-methylpentanal.


This time there are 5 carbons in the longest chain, including the one in
the -CHO group. There aren't any carbon-carbon double bonds. A methyl
group is attached to the number 2 carbon. Notice that in aldehydes, the
carbon in the -CHO group is always counted as the number 1 carbon.

http://www.chemguide.co.uk/basicorg/conventions/names.html (12 of 14)30/12/2004 11:07:30

Understanding the names of organic compounds

Ketones
Ketones contain a carbon-oxygen double bond just like aldehydes, but
this time it's in the middle of a carbon chain. There isn't a hydrogen atom
attached to the group as there is in aldehydes.
Ketones are shown by the ending one.
Example 1: Write the structural formula for propanone.
This is a 3 carbon chain with no carbon-carbon double bond. The carbonoxygen double bond has to be in the middle of the chain and so must be
on the number 2 carbon.

Ketones are often written in this way to emphasise the carbon-oxygen


double bond.
Example 2: Write the structural formula for pentan-3-one.
This time the position of the carbon-oxygen double bond has to be stated
because there is more than one possibility. It's on the third carbon of a 5
carbon chain with no carbon-carbon double bonds. If it was on the
second carbon, it would be pentan-2-one.

This could equally well be written:

http://www.chemguide.co.uk/basicorg/conventions/names.html (13 of 14)30/12/2004 11:07:30

Understanding the names of organic compounds

Where would you like to go now?


Continue naming more chain compounds. . .
Naming aromatic compounds. . .
To the organic conventions menu. . .
To menu of basic organic chemistry. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/basicorg/conventions/names.html (14 of 14)30/12/2004 11:07:30

Understanding the names of organic compounds

THE NAMES OF MORE ORGANIC COMPOUNDS

This page continues looking at the names of organic compounds


containing chains of carbon atoms. It assumes that you have already
looked at the introductory page covering compounds from alkanes to
ketones.

Note: If you haven't already looked at that page, it would be


a good idea to do so before you go on. The names on this
second page aren't explained in quite as much detail as
those on the introductory page - it assumes that you have
already understood the main principles. If in doubt, follow this
link first.

More types of organic compound


Carboxylic acids
Carboxylic acids contain the -COOH group, which is better written out in
full as:

Carboxylic acids are shown by the ending oic acid. When you count the
carbon chain, you have to remember to include the carbon in the -COOH
group. That carbon is always thought of as number 1 in the chain.
Example 1: Write the structural formula for 3-methylbutanoic acid.
This is a four carbon acid with no carbon-carbon double bonds. There is
a methyl group on the third carbon (counting the -COOH carbon as
number 1).

http://www.chemguide.co.uk/basicorg/conventions/names2.html (1 of 10)30/12/2004 11:07:41

Understanding the names of organic compounds

Example 2: Write the structural formula for 2-hydroxypropanoic acid.


The hydroxy part of the name shows the presence of an -OH group.
Normally, you would show that by the ending ol, but this time you can't
because you've already got another ending. You are forced into this
alternative way of describing it.

The old name for 2-hydroxypropanoic acid is lactic acid. That name
sounds more friendly, but is utterly useless when it comes to writing a
formula for it. In the old days, you would have had to learn the formula
rather than just working it out should you need it.
Example 3: Write the structural formula for 2-chlorobut-3-enoic acid.
This time, not only is there a chlorine attached to the chain, but the chain
also contains a carbon-carbon double bond (en) starting on the number 3
carbon (counting the -COOH carbon as number 1).

Salts of carboxylic acids


Example: Write the structural formula for sodium propanoate.
This is the sodium salt of propanoic acid - so start from that. Propanoic
acid is a three carbon acid with no carbon-carbon double bonds.

http://www.chemguide.co.uk/basicorg/conventions/names2.html (2 of 10)30/12/2004 11:07:41

Understanding the names of organic compounds

When the carboxylic acids form salts, the hydrogen in the -COOH group
is replaced by a metal. Sodium propanoate is therefore:

Notice that there is an ionic bond between the sodium and the
propanoate group. Whatever you do, don't draw a line between the
sodium and the oxygen. That would represent a covalent bond. It's
wrong, and makes you look very incompetent in an exam!
In a shortened version, sodium propanoate would be written
CH3CH2COONa or, if you wanted to emphasise the ionic nature, as
CH3CH2COO- Na+.

Note: The confusing thing about these salts (and even more
so for the esters that are coming up next) is that they are
named the wrong way round. In the formula, the sodium is at
the end, but appears first in the name. Why?
Salts are always named with the metal first - think of sodium
chloride or potassium iodide. So for consistency you would
need to reverse the formula of sodium propanoate NaOOCCH2CH3. But if you reverse the formula, you can't
see immediately that it is related to propanoic acid. So you
learn to live with the inconsistency.

http://www.chemguide.co.uk/basicorg/conventions/names2.html (3 of 10)30/12/2004 11:07:41

Understanding the names of organic compounds

Esters
Esters are one of a number of compounds known collectively as acid
derivatives. In these the acid group is modified in some way. In an ester,
the hydrogen in the -COOH group is replaced by an alkyl group (or
possibly some more complex hydrocarbon group).
Example 1: Write the structural formula for methyl propanoate.
An ester name has two parts - the part that comes from the acid
(propanoate) and the part that shows the alkyl group (methyl).
Start by thinking about propanoic acid - a 3 carbon acid with no carboncarbon double bonds.

The hydrogen in the -COOH group is replaced by an alkyl group - in this


case, a methyl group.

Ester names are confusing because the name is written backwards from
the way the structure is drawn. There's no way round this - you just have
to get used to it!
In the shortened version, this formula would be written CH3CH2COOCH3.
Example 2: Write the structural formula for ethyl ethanoate.
This is probably the most commonly used example of an ester. It is
based on ethanoic acid ( hence, ethanoate) - a 2 carbon acid. The
hydrogen in the -COOH group is replaced by an ethyl group.

http://www.chemguide.co.uk/basicorg/conventions/names2.html (4 of 10)30/12/2004 11:07:41

Understanding the names of organic compounds

Make sure that you draw the ethyl group the right way round. A fairly
common mistake is to try to join the CH3 group to the oxygen. If you
count the bonds if you do that, you will find that both the CH3 carbon and
the CH2 carbon have the wrong number of bonds.

Acyl chlorides (acid chlorides)


An acyl chloride is another acid derivative. In this case, the -OH group of
the acid is replaced by -Cl. All acyl chlorides contain the -COCl group:

Example: Write the structural formula for ethanoyl chloride.


Acyl chlorides are shown by the ending oyl chloride. So ethanoyl
chloride is based on a 2 carbon chain with no carbon-carbon double
bonds and a -COCl group. The carbon in that group counts as part of the
chain. In a longer chain, with side groups attached, the -COCl carbon is
given the number 1 position.

Acid anhydrides
Another acid derivative! An acid anhydride is what you get if you
dehydrate an acid - that is, remove water from it.

http://www.chemguide.co.uk/basicorg/conventions/names2.html (5 of 10)30/12/2004 11:07:41

Understanding the names of organic compounds

Example: Write the structural formula for propanoic anhydride.


These are most easily worked out by writing it down on a scrap of paper
in the following way:

Draw two molecules of acid arranged so that the -OH groups are next to
each other. Tweak out a molecule of water - and then join up what's left.
In this case, because you want propanoic anhydride, you draw two
molecules of propanoic acid.

Amides
Yet another acid derivative! Amides contain the group -CONH2 where the
-OH of an acid is replaced by -NH2.
Example: Write the structural formula for propanamide.
This is based on a 3 carbon chain with no carbon-carbon double bonds.
At the end of the chain is a -CONH2 group. The carbon in that group
counts as part of the chain.

Nitriles
Nitriles contain a -CN group, and used to be called cyanides.
http://www.chemguide.co.uk/basicorg/conventions/names2.html (6 of 10)30/12/2004 11:07:41

Understanding the names of organic compounds

Example 1: Write the structural formula for ethanenitrile.


The name shows a 2 carbon chain with no carbon-carbon double bond.
nitrile shows a -CN group at the end of the chain. As with the previous
examples involving acids and acid derivatives, don't forget that the
carbon in the -CN group counts as part of the chain.

The old name for this would have been methyl cyanide. You might think
that that's easier, but as soon as the chain gets more complicated, it
doesn't work - as the next example shows.
Example 2: Write the structural formula for 2-hydroxypropanenitrile.
Here we've got a 3 carbon chain, no carbon-carbon double bonds, and a CN group on the end of the chain. The carbon in the -CN group counts
as the number 1 carbon. On the number 2 carbon there is an -OH group
(hydroxy). Notice that you can't use the ol ending because you've
already got a nitrile ending.

Primary amines
A primary amine contains the group -NH2 attached to a hydrocarbon
chain or ring. You can think of amines in general as being derived from
ammonia, NH3. In a primary amine, one of the hydrogens has been
replaced by a hydrocarbon group.
Example 1: Write the structural formula for ethylamine.
In this case, an ethyl group is attached to the -NH2 group.

http://www.chemguide.co.uk/basicorg/conventions/names2.html (7 of 10)30/12/2004 11:07:41

Understanding the names of organic compounds

This name (ethylamine) is fine as long as you've only got a short chain
where there isn't any ambiguity about where the -NH2 group is found. But
suppose you had a 3 carbon chain - in this case, the -NH2 group could
be on an end carbon or on the middle carbon. How you get around that
problem is illustrated in the next example.
Example 2: Write the structural formula for 2-aminopropane.
The name shows a 3 carbon chain with an amino group attached to the
second carbon. amino shows the -NH2 group.

Ethylamine (example 1 above) could equally well have been called


aminoethane.

Secondary and tertiary amines


You are only likely to come across simple examples of these. In a
secondary amine, two of the hydrogen atoms in an ammonia molecule
have been replaced by hydrocarbon groups. In a tertiary amine, all three
hydrogens have been replaced.
Example 1: Write the structural formula for dimethylamine.
In this case, two of the hydrogens in ammonia have been replaced by
methyl groups.

Example 2: Write the structural formula for trimethylamine.


Here, all three hydrogens in ammonia have been replaced by methyl
groups.
http://www.chemguide.co.uk/basicorg/conventions/names2.html (8 of 10)30/12/2004 11:07:41

Understanding the names of organic compounds

Amino acids
An amino acid contains both an amino group, -NH2, and a carboxylic
acid group, -COOH, in the same molecule. As with all acids the carbon
chain is numbered so that the carbon in the -COOH group is counted as
number 1.
Example: Write the structural formula for 2-aminopropanoic acid.
This has a 3 carbon chain with no carbon-carbon double bonds. On the
second carbon (counting the -COOH carbon as number 1) there is an
amino group, -NH2.

Where would you like to go now?


Naming aromatic compounds. . .
To the organic conventions menu. . .
To menu of basic organic chemistry. . .
To Main Menu . . .

http://www.chemguide.co.uk/basicorg/conventions/names2.html (9 of 10)30/12/2004 11:07:41

Understanding the names of organic compounds

Jim Clark 2000

http://www.chemguide.co.uk/basicorg/conventions/names2.html (10 of 10)30/12/2004 11:07:41

Naming aromatic compounds

THE NAMES OF AROMATIC COMPOUNDS

This page looks at the names of some simple aromatic compounds. An


aromatic compound is one which contains a benzene ring. It assumes
that you are reasonably confident about naming compounds containing
chains of carbon atoms (aliphatic compounds).

Note: If you aren't sure about naming aliphatic compounds


follow this link before you go on.

Naming aromatic compounds isn't quite so straightforward as naming


chain compounds. Often, more than one name is acceptable and it's not
uncommon to find the old names still in use as well.

Background
The benzene ring
All aromatic compounds are based on benzene, C6H6, which has a ring
of six carbon atoms and has the symbol:

Each corner of the hexagon has a carbon atom with a hydrogen attached.

http://www.chemguide.co.uk/basicorg/conventions/names3.html (1 of 10)30/12/2004 11:07:48

Naming aromatic compounds

Note: If you don't understand this structure, it is explained in


full in two pages on the structure of benzene elsewhere in
this site. Following this link could well take you some time!

The phenyl group


Remember that you get a methyl group, CH3, by removing a hydrogen
from methane, CH4.
You get a phenyl group, C6H5, by removing a hydrogen from a benzene
ring, C6H6. Like a methyl or an ethyl group, a phenyl group is always
attached to something else.

Aromatic compounds with only one group attached to the


benzene ring
Cases where the name is based on benzene
chlorobenzene
This is a simple example of a halogen attached to the benzene ring. The
name is self-obvious.

The simplified formula for this is C6H5Cl. You could therefore (although
you never do!) call it phenyl chloride. Whenever you draw a benzene ring
with one other thing attached to it, you are in fact drawing a phenyl
group. In order to attach something else, you have to remove one of the
existing hydrogen atoms, and so automatically make a phenyl group.
nitrobenzene
http://www.chemguide.co.uk/basicorg/conventions/names3.html (2 of 10)30/12/2004 11:07:48

Naming aromatic compounds

The nitro group, NO2, is attached to a benzene ring.

The simplified formula for this is C6H5NO2.


methylbenzene
Another obvious name - the benzene ring has a methyl group attached.
Other alkyl side-chains would be named similarly - for example,
ethylbenzene. The old name for methylbenzene is toluene, and you may
still meet that.

The simplified formula for this is C6H5CH3.


(chloromethyl)benzene
A variant on this which you may need to know about is where one of the
hydrogens on the CH3 group is replaced by a chlorine atom. Notice the
brackets around the (chloromethyl) in the name. This is so that you are
sure that the chlorine is part of the methyl group and not somewhere else
on the ring.

If more than one of the hydrogens had been replaced by chlorine, the
names would be (dichloromethyl)benzene or (trichloromethyl)benzene.
Again, notice the importance of the brackets in showing that the chlorines
http://www.chemguide.co.uk/basicorg/conventions/names3.html (3 of 10)30/12/2004 11:07:48

Naming aromatic compounds

are part of the side group and not directly attached to the ring.
benzoic acid (benzenecarboxylic acid)
Benzoic acid is the older name, but is still in common use - it's a lot
easier to say and write than the modern alternative! Whatever you call it,
it has a carboxylic acid group, -COOH, attached to the benzene ring.

Cases where the name is based on phenyl


Remember that the phenyl group is a benzene ring minus a hydrogen
atom - C6H5. If you draw a benzene ring with one group attached, you
have drawn a phenyl group.
phenylamine
Phenylamine is a primary amine and contains the -NH2 group attached to
a benzene ring.

The old name for phenylamine is aniline, and you could also reasonably
call it aminobenzene. Phenylamine is what it is most commonly called at
A'level.

http://www.chemguide.co.uk/basicorg/conventions/names3.html (4 of 10)30/12/2004 11:07:48

Naming aromatic compounds

Note: In all cases where there is some possibility of


alternative names, you need to know what your examiners
are likely to call a particular compound. Refer to your syllabus
and recent exam papers. If you haven't got these, follow this
link to find out how to get hold of them.

phenylethene
This is an ethene molecule with a phenyl group attached. Ethene is a two
carbon chain with a carbon-carbon double bond. Phenylethene is
therefore:

The old name for phenylethene is styrene - the monomer from which
polystyrene is made.
phenylethanone
This is a slightly awkward name - take it to pieces. It consists of a two
carbon chain with no carbon-carbon double bond. The one ending shows
that it is a ketone, and so has a C=O group somewhere in the middle.
Attached to the carbon chain is a phenyl group. Putting that together
gives:

phenyl ethanoate
This is an ester based on ethanoic acid. The hydrogen atom in the COOH group has been replaced by a phenyl group.
http://www.chemguide.co.uk/basicorg/conventions/names3.html (5 of 10)30/12/2004 11:07:48

Naming aromatic compounds

Note: If you aren't happy about naming esters, follow this


link before you go on.

phenol
Phenol has an -OH group attached to a benzene ring and so has a
formula C6H5OH.

Aromatic compounds with more than one group attached


to the benzene ring
Numbering the ring
Any group already attached to the ring is given the number 1 position.
Where you draw it on the ring (at the top or in any other position) doesn't
matter - that's just a question of rotating the molecule a bit. It's much
easier, though, to get in the habit of drawing your main group at the top.
The other ring positions are then numbered from 2 to 6. You can number
them either clockwise or anti-clockwise. As with chain compounds, you
number the ring so that the name you end up with has the smallest
possible numbers in it. Examples will make this clear.
http://www.chemguide.co.uk/basicorg/conventions/names3.html (6 of 10)30/12/2004 11:07:48

Naming aromatic compounds

Some simple examples


Substituting chlorine atoms on the ring
Look at these compounds:

All of these are based on methylbenzene and so the methyl group is


given the number 1 position on the ring.
Why is it 2-chloromethylbenzene rather than 6-chloromethylbenzene?
The ring is numbered clockwise in this case because that produces a 2in the name rather than a 6-. 2 is smaller than 6.

Warning! You will find all sorts of variations on this


depending on the age of the book you look it up in, and
where it was published. What I have described above isn't in
strict accordance with the most modern interpretation of the
IUPAC recommendations for naming organic compounds.
The names should actually be 1-chloro-2-methylbenzene, 1chloro-3-methylbenzene, and so on. The substituted groups
are named in alphabetical order, and the "1" position is
assigned to the first of these - rather than to the more logical
methyl group.
This produces some silly inconsistencies. For example, if you
had the exactly equivalent compounds containing nitro
groups in place of the chlorines, the names would change
completely, to 1-methyl-2-nitrobenzene, 1-methyl-3nitrobenzene, etc. In this case, the normal practice of naming
the hydrocarbon first, and then attaching other things to it has
been completely wrecked.
http://www.chemguide.co.uk/basicorg/conventions/names3.html (7 of 10)30/12/2004 11:07:48

Naming aromatic compounds

Do you need to worry about this? NO! It is extremely unlikely


that you would ever be asked to name these in an exam, and
it is always easy to write a structure from one of these names
- however illogical it may be! There is a simple rule for exam
purposes. Unless you are specifically asked for the name of
anything remotely complicated, don't give it. As long as you
have got the structure right, that's all that matters.

2-hydroxybenzoic acid
This might also be called 2-hydroxybenzenecarboxylic acid. There is a COOH group attached to the ring and, because the name is based on
benzoic acid, that group is assigned the number 1 position. Next door to
it in the 2 position is a hydroxy group, -OH.

benzene-1,4-dicarboxylic acid
The di shows that there are two carboxylic acid groups, -COOH, one of
them in the 1 position and the other opposite it in the 4 position.

2,4,6-trichlorophenol
This is based on phenol - with an -OH group attached in the number 1
position on the ring. There are 3 chlorine atoms substituted onto the ring
in the 2, 4 and 6 positions.

http://www.chemguide.co.uk/basicorg/conventions/names3.html (8 of 10)30/12/2004 11:07:48

Naming aromatic compounds

2,4,6-trichlorophenol is the familiar antiseptic TCP.


methyl 3-nitrobenzoate
This is a name you might come across as a part of a practical exercise in
nitrating benzene rings. It's included partly for that reason, and partly
because it is a relatively complicated name to finish with!
The structure of the name shows that it is an ester. You can tell that from
the oate ending, and the methyl group floating separately from the rest of
the name at the beginning.
The ester is based on the acid, 3-nitrobenzoic acid - so start with that.
There will be a benzene ring with a -COOH group in the number 1
position and a nitro group, NO2, in the 3 position. The -COOH group is
modified to make an ester by replacing the hydrogen of the -COOH
group by a methyl group.
Methyl 3-nitrobenzoate is therefore:

Where would you like to go now?


To the organic conventions menu. . .
To menu of basic organic chemistry. . .
http://www.chemguide.co.uk/basicorg/conventions/names3.html (9 of 10)30/12/2004 11:07:48

Naming aromatic compounds

To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/basicorg/conventions/names3.html (10 of 10)30/12/2004 11:07:48

properties of organic compounds menu

Understanding Chemistry

PROPERTIES OF ORGANIC COMPOUNDS MENU

You can find details of the physical and chemical properties for each of
these types of compound:

Aliphatic compounds
These are compounds where the functional group is not attached directly
to a benzene ring.
alkanes . . .
alkenes . . .
halogenoalkanes (haloalkanes or alkyl halides) . . .
alcohols . . .
aldehydes and ketones . . .
carboxylic acids . . .
acyl chlorides (acid chlorides) . . .
acid anhydrides . . .
esters . . .
amides . . .
http://www.chemguide.co.uk/orgpropsmenu.html (1 of 2)30/12/2004 11:07:50

properties of organic compounds menu

nitriles . . .
amines . . .
amino acids and proteins . . .

Aromatic compounds
These are compounds based on benzene rings. This isn't a complete list,
but shows only those compounds where there are important differences
between them and their aliphatic equivalents.
arenes (benzene and methylbenzene) . . .
aryl halides (e.g. chlorobenzene) . . .
phenol . . .
phenylamine (aniline) and diazonium compounds . . .

Go to Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/orgpropsmenu.html (2 of 2)30/12/2004 11:07:50

Alkanes Menu

Understanding Chemistry

ALKANES MENU

Background . . .
An introduction to the alkanes (including cycloalkanes) and their
physical properties.
Burning alkanes . . .
The combustion of alkanes, including incomplete combustion.
Halogenation . . .
The reaction between alkanes and chlorine or bromine.
Cracking . . .
A brief look at cracking alkanes in the oil industry.

Go to menu of other organic compounds . . .


Go to Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/alkanemenu.html (1 of 2)30/12/2004 11:07:51

Alkanes Menu

http://www.chemguide.co.uk/organicprops/alkanemenu.html (2 of 2)30/12/2004 11:07:51

an introduction to alkanes and cycloalkanes

INTRODUCING ALKANES AND CYCLOALKANES

This is an introductory page about alkanes such as methane, ethane,


propane, butane and the rest. It deals with their formulae and isomerism,
their physical properties, and an introduction to their chemical reactivity.

What are alkanes and cycloalkanes?


Alkanes
Formulae
Alkanes are the simplest family of hydrocarbons - compounds containing
carbon and hydrogen only. They only contain carbon-hydrogen bonds
and carbon-carbon single bonds. The first six are:
methane

CH4

ethane

C2H6

propane

C3H8

butane

C4H10

pentane

C5H12

hexane

C6H14

You can work out the formula of any of them using: CnH2n+2
Isomerism
All the alkanes with 4 or more carbon atoms in them show structural
isomerism. This means that there are two or more different structural
formulae that you can draw for each molecular formula.

http://www.chemguide.co.uk/organicprops/alkanes/background.html (1 of 9)30/12/2004 11:07:56

an introduction to alkanes and cycloalkanes

For example, C4H10 could be either of these two different molecules:

These are called respectively butane and 2-methylpropane.

Note: If you aren't confident about naming organic


compounds, the various ways of drawing organic
compounds, or structural isomerism, then you really ought to
follow these links before you go on.
You should read the whole of the page about drawing organic
molecules, but there is no need to read the other two beyond
where they talk about alkanes.
Use the BACK button on your browser to return to this page.

Cycloalkanes
Cycloalkanes again only contain carbon-hydrogen bonds and carboncarbon single bonds, but this time the carbon atoms are joined up in a
ring. The smallest cycloalkane is cyclopropane.

http://www.chemguide.co.uk/organicprops/alkanes/background.html (2 of 9)30/12/2004 11:07:56

an introduction to alkanes and cycloalkanes

If you count the carbons and hydrogens, you will see that they no longer
fit the general formula CnH2n+2. By joining the carbon atoms in a ring,
you have had to lose two hydrogen atoms.
You are unlikely to ever need it, but the general formula for a cycloalkane
is CnH2n.
Don't imagine that these are all flat molecules. All the cycloalkanes from
cyclopentane upwards exist as "puckered rings".
Cyclohexane, for example, has a ring structure which looks like this:

This is known as the "chair" form of cyclohexane - from its shape which
vaguely resembles a chair.

http://www.chemguide.co.uk/organicprops/alkanes/background.html (3 of 9)30/12/2004 11:07:56

an introduction to alkanes and cycloalkanes

Note: This molecule is constantly changing, with the atom


on the left which is currently pointing down flipping up, and
the one on the right flipping down. During the process,
another (slightly less stable) form of cyclohexane is formed
known as the "boat" form. In this arrangement, both of these
atoms are either pointing up or down at the same time.

Physical Properties
Boiling Points
The facts

The boiling points shown are all for the "straight chain" isomers where
there are more than one.
Notice that the first four alkanes are gases at room temperature. Solids
don't start to appear until about C17H36.
You can't be more precise than that because each isomer has a different
melting and boiling point. By the time you get 17 carbons into an alkane,
there are unbelievable numbers of isomers!

http://www.chemguide.co.uk/organicprops/alkanes/background.html (4 of 9)30/12/2004 11:07:56

an introduction to alkanes and cycloalkanes

Cycloalkanes have boiling points which are about 10 - 20 K higher than


the corresponding straight chain alkane.
Explanations
There isn't much electronegativity difference between carbon and
hydrogen, so there is hardly any bond polarity. The molecules
themselves also have very little polarity. A totally symmetrical molecule
like methane is completely non-polar.

Note: If you aren't sure about electronegativity and polarity,


then you really ought to follow this link before you go on.
Use the BACK button on your browser to return to this page.

This means that the only attractions between one molecule and its
neighbours will be Van der Waals dispersion forces. These will be very
small for a molecule like methane, but will increase as the molecules get
bigger. That's why the boiling points of the alkanes increase with
molecular size.

Note: If you aren't sure about Van der Waals forces, then
you should follow this link before you go on.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/alkanes/background.html (5 of 9)30/12/2004 11:07:56

an introduction to alkanes and cycloalkanes

Where you have isomers, the more branched the chain, the lower the
boiling point tends to be. Van der Waals dispersion forces are smaller for
shorter molecules, and only operate over very short distances between
one molecule and its neighbours. It is more difficult for short fat
molecules (with lots of branching) to lie as close together as long thin
ones.
For example, the boiling points of the three isomers of C5H12 are:
boiling point (K)
pentane

309.2

2-methylbutane

301.0

2,2-dimethylpropane

282.6

The slightly higher boiling points for the cycloalkanes are presumably
because the molecules can get closer together because the ring
structure makes them tidier and less "wriggly"!

Solubility
The facts
What follows applies equally to alkanes and cycloalkanes.
Alkanes are virtually insoluble in water, but dissolve in organic solvents.
The liquid alkanes are good solvents for many other covalent compounds.
Explanations
Solubility in water
When a molecular substance dissolves in water, you have to

break the intermolecular forces within the substance. In the case


of the alkanes, these are Van der Waals dispersion forces.

http://www.chemguide.co.uk/organicprops/alkanes/background.html (6 of 9)30/12/2004 11:07:56

an introduction to alkanes and cycloalkanes

break the intermolecular forces in the water so that the substance


can fit between the water molecules. In water the main
intermolecular attractions are hydrogen bonds.

Note: If you aren't sure about hydrogen bonds, then you


should follow this link before you go on.
Use the BACK button on your browser to return to this page.

Breaking either of these attractions costs energy, although the amount of


energy to break the Van der Waals dispersion forces in something like
methane is pretty negligible. That isn't true of the hydrogen bonds in
water, though.
As something of a simplification, a substance will dissolve if there is
enough energy released when new bonds are made between the
substance and the water to make up for what is used in breaking the
original attractions.
The only new attractions between the alkane and water molecules are
Van der Waals. These don't release anything like enough energy to
compensate for what you need to break the hydrogen bonds in water.
The alkane doesn't dissolve.

Note: The reason that this is a simplification is that you also


have to consider entropy changes when things dissolve. If
you don't yet know about entropy, don't worry about it!

http://www.chemguide.co.uk/organicprops/alkanes/background.html (7 of 9)30/12/2004 11:07:56

an introduction to alkanes and cycloalkanes

Solubility in organic solvents


In most organic solvents, the main forces of attraction between the
solvent molecules are Van der Waals - either dispersion forces or dipoledipole attractions.
That means that when an alkane dissolves in an organic solvent, you are
breaking Van der Waals forces and replacing them by new Van der
Waals forces. The two processes more or less cancel each other out
energetically - so there isn't any barrier to solubility.

Chemical Reactivity
Alkanes
Alkanes contain strong carbon-carbon single bonds and strong carbonhydrogen bonds. The carbon-hydrogen bonds are only very slightly polar
and so there aren't any bits of the molecules which carry any significant
amount of positive or negative charge which other things might be
attracted to.
The net effect is that alkanes have a fairly restricted set of reactions.
You can

burn them - destroying the whole molecule;


react them with some of the halogens, breaking carbon-hydrogen
bonds;
crack them, breaking carbon-carbon bonds.

These reactions are all covered on separate pages if you go to the


alkanes menu (see below).
Cycloalkanes
Cycloalkanes are very similar to the alkanes in reactivity, except for the

http://www.chemguide.co.uk/organicprops/alkanes/background.html (8 of 9)30/12/2004 11:07:56

an introduction to alkanes and cycloalkanes

very small ones - especially cyclopropane. Cyclopropane is much more


reactive than you would expect.
The reason has to do with the bond angles in the ring. Normally, when
carbon forms four single bonds, the bond angles are about 109.5. In
cyclopropane, they are 60.

With the electron pairs this close together, there is a lot of repulsion
between the bonding pairs joining the carbon atoms. That makes the
bonds easier to break.
The effect of this is explored on the page about reactions of these
compounds with halogens which you can access from the alkanes menu
below.

Where would you like to go now?


To the alkanes menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/alkanes/background.html (9 of 9)30/12/2004 11:07:56

Alkenes Menu

Understanding Chemistry

ALKENES MENU

Background . . .
An introduction to the alkenes, their reactivity and their physical
properties.
Making alkenes . . .
Making alkenes in the lab from alcohols.
Hydrogenation . . .
The reaction between alkenes and hydrogen.
Reactions with halogens . . .
The reactions with chlorine and bromine, including the test using
bromine water.
Reactions with hydrogen halides . . .
The reactions with compounds such as hydrogen chloride and
hydrogen bromide.
Reaction with sulphuric acid . . .
The reaction between alkenes and concentrated sulphuric acid
(including further reaction to give alcohols).
Reaction with potassium manganate(VII) . . .

http://www.chemguide.co.uk/organicprops/alkenemenu.html (1 of 2)30/12/2004 11:07:59

Alkenes Menu

The reaction between alkenes and potassium manganate(VII)


(potassium permanganate) solution.
Hydration . . .
The industrial hydration of alkenes (their reaction with water) to
give alcohols.
Polymerisation . . .
The polymerisation of alkenes, including some uses of the
polymers formed.
Making epoxyethane . . .
The manufacture of epoxyethane (old name: ethylene oxide) from
ethene, including its uses.

Go to menu of other organic compounds . . .


Go to Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/alkenemenu.html (2 of 2)30/12/2004 11:07:59

an introduction to alkenes

INTRODUCING ALKENES

This is an introductory page about alkenes such as ethene, propene and


the rest. It deals with their formulae and isomerism, their physical
properties, and an introduction to their chemical reactivity.

What are alkenes?


Formulae
Alkenes are a family of hydrocarbons (compounds containing carbon and
hydrogen only) containing a carbon-carbon double bond. The first two
are:
ethene

C2H4

propene

C3H6

You can work out the formula of any of them using: CnH2n
The table is limited to the first two, because after that there are isomers
which affect the names.

Isomerism in the alkenes


Structural isomerism
All the alkenes with 4 or more carbon atoms in them show structural
isomerism. This means that there are two or more different structural
formulae that you can draw for each molecular formula.
For example, with C4H8, it isn't too difficult to come up with these three
structural isomers:
http://www.chemguide.co.uk/organicprops/alkenes/background.html (1 of 8)30/12/2004 11:08:26

an introduction to alkenes

Note: If you aren't confident about naming organic


compounds, the various ways of drawing organic
compounds, or structural isomerism, then you really ought to
follow these links before you go on.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/alkenes/background.html (2 of 8)30/12/2004 11:08:26

an introduction to alkenes

There is, however, another isomer. But-2-ene also exhibits geometric


isomerism.
Geometric (cis-trans) isomerism
The carbon-carbon double bond doesn't allow any rotation about it. That
means that it is possible to have the CH3 groups on either end of the
molecule locked either on one side of the molecule or opposite each
other.
These are called cis-but-2-ene (where the groups are on the same side)
or trans-but-2-ene (where they are on opposite sides).

Note: If you aren't confident about geometric isomerism,


then it is essential that you follow this link before you go on.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/alkenes/background.html (3 of 8)30/12/2004 11:08:26

an introduction to alkenes

Physical properties of the alkenes


Boiling Points
The boiling point of each alkene is very similar to that of the alkane with
the same number of carbon atoms. Ethene, propene and the various
butenes are gases at room temperature. All the rest that you are likely to
come across are liquids.
In each case, the alkene has a boiling point which is a small number of
degrees lower than the corresponding alkane. The only attractions
involved are Van der Waals dispersion forces, and these depend on the
shape of the molecule and the number of electrons it contains. Each
alkene has 2 fewer electrons than the alkane with the same number of
carbons.

Note: If you aren't sure about Van der Waals forces, then
you should follow this link before you go on.
You will find the boiling points of the alkanes explained in
some detail on the introductory alkanes page. Everything
said there applies equally to the alkenes.
You will find the way geometric isomerism affects melting and
boiling points explained towards the bottom of the page you
get to by following this link.
Use the BACK button on your browser to return to this page.

Solubility
Alkenes are virtually insoluble in water, but dissolve in organic solvents.

http://www.chemguide.co.uk/organicprops/alkenes/background.html (4 of 8)30/12/2004 11:08:26

an introduction to alkenes

Note: The reasons for this are exactly the same as for the
alkanes. You will find a detailed explanation on the
introductory alkanes page.
Use the BACK button on your browser to return to this page.

Chemical Reactivity
Bonding in the alkenes
We just need to look at ethene, because what is true of C=C in ethene
will be equally true of C=C in more complicated alkenes.
Ethene is often modelled like this:

The double bond between the carbon atoms is, of course, two pairs of
shared electrons. What the diagram doesn't show is that the two pairs
aren't the same as each other.
One of the pairs of electrons is held on the line between the two carbon
nuclei as you would expect, but the other is held in a molecular orbital
above and below the plane of the molecule. A molecular orbital is a
region of space within the molecule where there is a high probability of
finding a particular pair of electrons.

http://www.chemguide.co.uk/organicprops/alkenes/background.html (5 of 8)30/12/2004 11:08:26

an introduction to alkenes

In this diagram, the line between the two carbon atoms represents a
normal bond - the pair of shared electrons lies in a molecular orbital on
the line between the two nuclei where you would expect them to be. This
sort of bond is called a sigma bond.
The other pair of electrons is found somewhere in the shaded part above
and below the plane of the molecule. This bond is called a pi bond. The
electrons in the pi bond are free to move around anywhere in this shaded
region and can move freely from one half to the other.

Note: This diagram shows a side view of an ethene


molecule. The dotted lines to two of the hydrogens show
bonds going back into the screen or paper away from you.
The wedge shapes show bonds coming out towards you.

The pi electrons are not as fully under the control of the carbon nuclei as
the electrons in the sigma bond and, because they lie exposed above
and below the rest of the molecule, they are relatively open to attack by
other things.

http://www.chemguide.co.uk/organicprops/alkenes/background.html (6 of 8)30/12/2004 11:08:26

an introduction to alkenes

Note: Check your syllabus to see if you need to know how a


pi bond is formed. If you are a UK A level student and haven't
got a syllabus, find out how to get one by following this link.
If you do need to know about the bonding in ethene in detail,
follow this link as well.
Use the BACK button on your browser to return to this page.

The reactions of alkenes


Like any other hydrocarbons, alkenes burn in air or oxygen, but these
reactions are unimportant. Alkenes are too valuable to waste in this way.
The important reactions all centre around the double bond. Typically, the
pi bond breaks and the electrons from it are used to join the two carbon
atoms to other things. Alkenes undergo addition reactions.
For example, using a general molecule X-Y . . .

The rather exposed electrons in the pi bond are particularly open to


attack by things which carry some degree of positive charge. These are
called electrophiles. If you explore the rest of the alkene menu, you will
find lots of examples of this kind.

http://www.chemguide.co.uk/organicprops/alkenes/background.html (7 of 8)30/12/2004 11:08:26

an introduction to alkenes

Note: If you need to know about organic reaction


mechanisms, it would be a good idea to read the page
explaining the background to electrophilic addition before you
start looking at individual cases from the alkenes menu.
In fact, if you are only really interested in mechanisms, then
look at that page and then explore the rest of the electrophilic
addition menu in the mechanisms section of this site.

Where would you like to go now?


To the alkenes menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/alkenes/background.html (8 of 8)30/12/2004 11:08:26

making alkenes in the lab

MAKING ALKENES IN THE LAB

This page looks at ways of preparing alkenes in the lab by the


dehydration of alcohols.

Dehydration of alcohols using aluminium oxide as catalyst


The dehydration of ethanol to give ethene
This is a simple way of making gaseous alkenes like ethene. If ethanol
vapour is passed over heated aluminium oxide powder, the ethanol is
essentially cracked to give ethene and water vapour.

To make a few test tubes of ethene, you can use this apparatus:

It wouldn't be too difficult to imagine scaling this up by boiling some


ethanol in a flask and passing the vapour over aluminium oxide heated in
a long tube.

Dehydration of alcohols using an acid catalyst


http://www.chemguide.co.uk/organicprops/alkenes/making.html (1 of 4)30/12/2004 11:08:37

making alkenes in the lab

The acid catalysts normally used are either concentrated sulphuric acid
or concentrated phosphoric(V) acid, H3PO4.
Concentrated sulphuric acid produces messy results. Not only is it an
acid, but it is also a strong oxidising agent. It oxidises some of the
alcohol to carbon dioxide and at the same time is reduced itself to
sulphur dioxide. Both of these gases have to be removed from the alkene.
It also reacts with the alcohol to produce a mass of carbon. There are
other side reactions as well, but these aren't required by any current UK
A level syllabus.

The dehydration of ethanol to give ethene


Ethanol is heated with an excess of concentrated sulphuric acid at a
temperature of 170C. The gases produced are passed through sodium
hydroxide solution to remove the carbon dioxide and sulphur dioxide
produced from side reactions.
The ethene is collected over water.

WARNING! This is potentially an extremely dangerous


preparation because of the close proximity of the very hot
concentrated sulphuric acid and the sodium hydroxide
solution. I knew of one chemistry teacher who put several
students into hospital by getting it wrong! That was many
years ago before safety was taken quite so seriously as it is
now.

The concentrated sulphuric acid is a catalyst. Write it over the arrow


rather than in the equation.

http://www.chemguide.co.uk/organicprops/alkenes/making.html (2 of 4)30/12/2004 11:08:37

making alkenes in the lab

Note: You will find the mechanism for the dehydration of


alcohols in the mechanism section of this site. You will also
find a discussion of how to cope with questions about the
dehydration of more complicated alcohols if you follow a link
at the bottom of that page.
Use the BACK button (or the HISTORY file or GO menu) on
your browser if you want to return to this page.

The dehydration of cyclohexanol to give cyclohexene


This is a preparation commonly used at this level to illustrate the
formation and purification of a liquid product. The fact that the carbon
atoms happen to be joined in a ring makes no difference whatever to the
chemistry of the reaction.
Cyclohexanol is heated with concentrated phosphoric(V) acid and the
liquid cyclohexene distils off and can be collected and purified.
Phosphoric(V) acid tends to be used in place of sulphuric acid because it
is safer and produces a less messy reaction.

Where would you like to go now?


To the alkenes menu . . .

http://www.chemguide.co.uk/organicprops/alkenes/making.html (3 of 4)30/12/2004 11:08:37

making alkenes in the lab

To the menu of other organic compounds . . .


To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/alkenes/making.html (4 of 4)30/12/2004 11:08:37

dehydration of ethanol to give ethene

THE DEHYDRATION OF ALCOHOLS

This page gives you the facts and a simple, uncluttered mechanism for
the acid catalysed dehydration of a simple alcohol like ethanol to give an
alkene like ethene. If you want the mechanism explained to you in detail,
there is a link at the bottom of the page.
You will also find a link to a page on the dehydration of more complicated
alcohols where more than one product may be formed.

The dehydration of ethanol


The facts
Ethanol can be dehydrated to give ethene by heating it with an excess of
concentrated sulphuric acid at about 170C. Concentrated phosphoric(V)
acid, H3PO4, can be used instead.

The acids aren't written into the equation because they serve as
catalysts. If you like, you could write, for example, "conc H2SO4" over the
top of the arrow.

Note: There are many side reactions which go on at the


same time. These aren't required by any current A' level
syllabus.

http://www.chemguide.co.uk/mechanisms/elim/dhethanol.html (1 of 4)30/12/2004 11:08:42

dehydration of ethanol to give ethene

The mechanism - the full version


We are going to discuss the mechanism using sulphuric acid. Afterwards,
we'll describe how you can use a simplified version which will work for
any acid, including phosphoric(V) acid.
In the first stage, one of the lone pairs of electrons on the oxygen picks
up a hydrogen ion from the sulphuric acid. The alcohol is said to be
protonated.

The protonated ethanol loses a water molecule to give a carbocation (a


carbonium ion).

Finally, a hydrogensulphate ion (from the sulphuric acid) pulls off a


hydrogen ion from the carbocation.

The mechanism - a simplified version

http://www.chemguide.co.uk/mechanisms/elim/dhethanol.html (2 of 4)30/12/2004 11:08:42

dehydration of ethanol to give ethene

People normally quote a simplified version of this mechanism. The only


Exam Board to want the mechanism for the dehydration of alcohols
(AQA) is happy to accept this version.
Instead of showing the full structure of the sulphuric acid, you write it as if
it were simply a hydrogen ion, H+. That leaves the full mechanism:

An advantage of this (apart from the fact that it doesn't require you to
draw the structure of sulphuric acid) is that it can be used for any acid
catalyst without changing it at all. For example, if you use this version,
you wouldn't need to worry about the structure of phosphoric(V) acid.

Note: Although most people probably write the mechanism


in this form, it is actually quite misleading because it suggests
the possibility of a free hydrogen ion in a chemical system. A
free hydrogen ion is a raw proton and this is always attached
to something else during a chemical reaction.
Personally, I find this simplification sloppy - but if your
examiners are happy to accept it, who am I to argue! Go for
the simple life!

http://www.chemguide.co.uk/mechanisms/elim/dhethanol.html (3 of 4)30/12/2004 11:08:42

dehydration of ethanol to give ethene

Where would you like to go now?


Help! Talk me through these mechanisms . . .
How to deal with more complicated cases . . .
To menu of elimination reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/elim/dhethanol.html (4 of 4)30/12/2004 11:08:42

Explaining the dehydration of ethanol to give ethene

EXPLAINING THE DEHYDRATION OF ALCOHOLS

This page guides you through the mechanism for the acid catalysed
dehydration of a simple alcohol like ethanol to give an alkene like ethene.
Dehydration of more complicated alcohols is dealt with on a separate
page. This is an essential part of this topic, and you should follow the link
at the bottom of this page if you haven't already done so.

The dehydration of ethanol


The full version of the mechanism
This full version covers the mechanism using sulphuric acid. Afterwards,
we'll look at the simplified version which will work for any acid, including
phosphoric(V) acid.
The oxygen atom in the ethanol has two active lone pairs of electrons,
and one of these picks up a hydrogen ion from the sulphuric acid. The
alcohol is said to be protonated.

http://www.chemguide.co.uk/mechanisms/elim/dhethanoltt.html (1 of 5)30/12/2004 11:08:43

Explaining the dehydration of ethanol to give ethene

Note: Only one of the lone pairs on the oxygen is being


shown because only one is used in the reaction. Similarly, all
the multitude of lone pairs in the sulphuric acid are omitted
because they aren't relevant to the mechanism.
If you aren't happy about the use of curly arrows in
mechanisms, follow this link before you go on. Use the BACK
button on your browser to return to this page.

The negative ion produced is the hydrogensulphate ion, HSO4-.


Notice that the oxygen atom in the alcohol has gained a positive charge.
That charge has to be there for two reasons:

On the left hand side of the equation you start with two overall
neutral molecules. Assuming you forgot about the positive charge,
you would end up with a neutral species and a negative ion on the
right. Charges must balance in equations, so something is wrong.
The oxygen looks wrong! The oxygen atom is joined to 3 things
rather than its usual 2. Oxygen can only join to 3 things if it carries
a positive charge.

Note: Oxygen with a positive charge has the same


arrangement of electrons as a nitrogen atom - which normally
forms 3 bonds.

http://www.chemguide.co.uk/mechanisms/elim/dhethanoltt.html (2 of 5)30/12/2004 11:08:43

Explaining the dehydration of ethanol to give ethene

In the second stage of the reaction the protonated ethanol loses a water
molecule to leave a carbocation (previously known as a carbonium ion) an ion with a positive charge on a carbon atom. The carbon atom is
positive because it has lost the electron that it originally contributed to the
carbon-oxygen bond. Both of the electrons in that bond have moved onto
the oxygen atom, neutralising the oxygen's charge.

One of the several things that can now happen to this carbocation is for it
to lose a hydrogen ion from the CH3 group. This hydrogen ion is pulled
off by a hydrogensulphate ion to regenerate the sulphuric acid catalyst.

Note: The other things that might happen to the carbocation


lead to products like ethoxyethane and ethyl
hydrogensulphate. The mechanisms for these aren't required
by any current A' level syllabus.

The simplified version of the mechanism


Instead of showing the full structure of the sulphuric acid, it is commonly
written as if it were simply a hydrogen ion, H+.
An advantage of this (apart from the fact that it doesn't require you to
http://www.chemguide.co.uk/mechanisms/elim/dhethanoltt.html (3 of 5)30/12/2004 11:08:43

Explaining the dehydration of ethanol to give ethene

draw the structure of sulphuric acid) is that it can be used for any acid
catalyst without changing it at all. For example, if you use this version,
you wouldn't need to worry about the structure of phosphoric(V) acid.
AQA (the only Exam Board to demand the dehydration mechanism at the
moment) is happy to accept this version.

In the first stage, the ethanol gets protonated exactly as before - the only
difference is that you are writing H+ instead of the full structure of the
sulphuric acid.
The second stage is identical to the one in the full version of the
mechanism.
The final stage shows a hydrogen ion "falling off" the carbocation - rather
than being pulled off. This is seriously misleading, but it's what the
examiners want!

Where would you like to go now?


How to deal with more complicated cases . . .
To menu of elimination reactions. . .

http://www.chemguide.co.uk/mechanisms/elim/dhethanoltt.html (4 of 5)30/12/2004 11:08:43

Explaining the dehydration of ethanol to give ethene

To menu of other types of mechanism. . .


To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/elim/dhethanoltt.html (5 of 5)30/12/2004 11:08:43

dehydration of more complicated alcohols

DEHYDRATION OF MORE COMPLICATED


ALCOHOLS

This page builds on your understanding of the acid catalysed dehydration


of alcohols.
You have to be wary with more complicated alcohols in case there is the
possibility of more than one alkene being formed. Butan-2-ol is a good
example of this, with no less than three different alkenes being formed
when it is dehydrated.
Butan-2-ol is just an example to illustrate the problems. It is important
that you understand it so that you can work out what will happen in
similar cases. It would be quite impossible for you to learn what happens
with every single alcohol you might be presented with.
The basic facts and mechanisms for these reactions are exactly the
same as with a simple alcohol like ethanol. This page only deals with the
extra problems created by the possibility of more than one dehydration
product.

Important! What follow assumes that you are familiar with


the mechanism for the dehydration of ethanol. If you aren't, it
is essential that you follow this link before you go on.

http://www.chemguide.co.uk/mechanisms/elim/dhcomplex.html (1 of 5)30/12/2004 11:08:46

dehydration of more complicated alcohols

Background
To make the diagrams less cluttered, we'll use the simplified version of
the mechanism showing gain and loss of H+.
Remember that the mechanism takes place in three stages:

The alcohol is protonated by the acid catalyst.


The protonated alcohol loses a water molecule to give a
carbocation (carbonium ion).
The carbocation formed loses a hydrogen ion and forms a double
bond.

So, in the case of the dehydration of a simple alcohol like ethanol:

The dehydration of butan-2-ol


The first two stages
There is nothing new at all in these stages.

http://www.chemguide.co.uk/mechanisms/elim/dhcomplex.html (2 of 5)30/12/2004 11:08:46

dehydration of more complicated alcohols

In the first stage, the alcohol is protonated by picking up a hydrogen ion


from the sulphuric acid.

In the second stage, the positive ion then sheds a water molecule and
produces a carbocation.

The complication arises in the next step. When the carbocation loses a
hydrogen ion, where is it going to come from?
Where does the hydrogen get removed from?
So that a double bond can form, it will have to come from one of the
carbons next door to the one with the positive charge.
If a hydrogen ion is lost from the CH3 group

But-1-ene is formed.
If a hydrogen ion is lost from the CH2 group

http://www.chemguide.co.uk/mechanisms/elim/dhcomplex.html (3 of 5)30/12/2004 11:08:46

dehydration of more complicated alcohols

This time the product is but-2-ene, CH3CH=CHCH3.


In fact the situation is even more complicated than it looks, because but2-ene exhibits geometric isomerism. You get a mixture of two isomers
formed - cis-but-2-ene and trans-but-2-ene.

Which isomer gets formed is just a matter of chance.

Geometric isomerism: Isomerism is where you can draw


more than one arrangement of the atoms for a given
molecular formula. Geometric isomerism is a special case of
this involving molecules which have restricted rotation around
one of the bonds - in this case, a carbon-carbon double bond.
The C=C bond could only rotate if enough energy is put in to
break the pi bond. Effectively, except at high temperatures,
the C=C bond is "locked".
In the case of but-2-ene, the two CH3 groups will either both
be locked on one side of the C=C (to give the cis isomer), or
on opposite sides (to give the trans one).
For a full discussion of geometric isomerism follow this link.
Use the BACK button on your browser to return to this page.
Beware! It is easy to miss geometric isomers in an exam.
Always draw alkenes with the correct 120 bond angles
around the C=C bond as shown in the diagrams for the cis
and trans isomers above. If you take a short cut and write but2-ene as CH3CH=CHCH3, you will almost certainly miss the
fact that cis and trans forms are possible.
This is a rich source of questions in an exam. You could
easily throw away marks if you miss these possibilities.

http://www.chemguide.co.uk/mechanisms/elim/dhcomplex.html (4 of 5)30/12/2004 11:08:46

dehydration of more complicated alcohols

The overall result


Dehydration of butan-2-ol leads to a mixture containing:

but-1-ene

cis-but-2-ene

trans-but-2-ene

Where would you like to go now?


To menu of elimination reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/elim/dhcomplex.html (5 of 5)30/12/2004 11:08:46

Elimination Mechanisms Menu

Understanding Chemistry

ELIMINATION MECHANISMS MENU

Elimination reactions involving halogenoalkanes


Elimination from 2-bromopropane . . .
The formation of an alkene (propene) from 2-bromopropane.
Elimination from unsymmetrical halogenoalkanes . . .
Explains how to cope with cases where more than one elimination
product can be formed from a single halogenoalkane.
Elimination v. substitution . . .
Discusses how the reaction between a halogenoalkane and
hydroxide ions can lead to either an elimination reaction or
nucleophilic substitution.

The dehydration of alcohols


The dehydration of ethanol . . .
Facts and mechanism for the dehydration of ethanol using an acid
catalyst to give ethene.
The dehydration of more complicated alcohols . . .
Explains how to cope with cases where more than dehydration
http://www.chemguide.co.uk/mechanisms/elimmenu.html (1 of 2)30/12/2004 11:08:48

Elimination Mechanisms Menu

product is possible.

Go to menu of other types of mechanism. . .


Go to Main Menu . . .

You might also be interested in:


properties and reactions of halogenoalkanes . . .
Covers all the physical and chemical properties of halogenoalkanes
(haloalkanes or alkyl halides) required by UK A level syllabuses.
properties and reactions of alcohols . . .
Covers all the physical and chemical properties of alcohols required
by UK A level syllabuses.

Jim Clark 2000 (modified 2004)

http://www.chemguide.co.uk/mechanisms/elimmenu.html (2 of 2)30/12/2004 11:08:48

elimination from 2-bromopropane to give propene

THE ELIMINATION REACTIONS PRODUCING


ALKENES FROM SIMPLE HALOGENOALKANES

This page gives you the facts and a simple, uncluttered mechanism for
the elimination reaction between a simple halogenoalkane like 2bromopropane and hydroxide ions (from, for example, sodium hydroxide)
to give an alkene like propene. If you want the mechanism explained to
you in detail, there is a link at the bottom of the page.
You will also find a link to a page on elimination from more complicated
halogenoalkanes where more than one product may be formed.
Exam questions on this topic frequently ask about another possibility in
the reactions between halogenoalkanes and hydroxide ions - nucleophilic
substitution. There is also a link to a page discussing this.

Note: The competition between substitution and elimination


(including the conditions needed and the mechanisms for
both) is a rich source of exam questions if your syllabus
includes it. You will probably find that the questions centre
around secondary halogenoalkanes like 2-bromopropane,
because these can easily be persuaded to do either reaction.
That's why this page deals exclusively with 2-bromopropane.

The elimination reaction involving 2-bromopropane and


hydroxide ions
The facts
2-bromopropane is heated under reflux with a concentrated solution of
sodium or potassium hydroxide in ethanol. Heating under reflux involves
heating with a condenser placed vertically in the flask to avoid loss of
volatile liquids. Propene is formed and, because this is a gas, it passes
through the condenser and can be collected.

http://www.chemguide.co.uk/mechanisms/elim/elim.html (1 of 3)30/12/2004 11:08:51

elimination from 2-bromopropane to give propene

Everything else present (including anything formed in the alternative


substitution reaction) will be trapped in the flask.
The mechanism
In elimination reactions, the hydroxide ion acts as a base - removing a
hydrogen as a hydrogen ion from the carbon atom next door to the one
holding the bromine.
The resulting re-arrangement of the electrons expels the bromine as a
bromide ion and produces propene.

Where would you like to go now?


Help! Talk me through this mechanism . . .
How to deal with more complicated cases . . .
The competition between elimination and substitution . . .
To menu of elimination reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

http://www.chemguide.co.uk/mechanisms/elim/elim.html (2 of 3)30/12/2004 11:08:51

elimination from 2-bromopropane to give propene

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/elim/elim.html (3 of 3)30/12/2004 11:08:51

Explaining the elimination reaction producing propene from 2-bromopropane

EXPLAINING THE ELIMINATION REACTIONS


PRODUCING ALKENES FROM SIMPLE
HALOGENOALKANES

This page guides you through the elimination mechanism for the reaction
between simple halogenoalkanes like 2-bromopropane and hydroxide
ions from, for example, sodium hydroxide.
Elimination involving more complicated halogenoalkanes and the
competition between elimination and substitution in these reactions are
dealt with on separate pages. These are essential parts of this topic, and
you should follow the links at the bottom of this page if you haven't
already done so.

The elimination reaction involving 2-bromopropane and


hydroxide ions
The role of the OH- ion in an elimination reaction
Hydroxide ions have a very strong tendency to
combine with hydrogen ions to make water - in
other words, the OH- ion is a very strong base.
In an elimination reaction, the hydroxide ion hits
one of the hydrogen atoms in the CH3 group and
pulls it off. This leads to a cascade of electron pair movements resulting
in the formation of a carbon-carbon double bond, and the loss of the
bromine as Br-.

Note: If you aren't happy about the use of curly arrows in


mechanisms, follow this link before you go on. Use the BACK
button on your browser to return to this page.

http://www.chemguide.co.uk/mechanisms/elim/elimtt.html (1 of 3)30/12/2004 11:08:54

Explaining the elimination reaction producing propene from 2-bromopropane

The complete elimination mechanism

The OH- ion takes one of the hydrogens from the CH3 group, but it
only needs the hydrogen nucleus (a hydrogen ion). That means
that the two electrons which originally joined the hydrogen to the
carbon aren't being used any more.
Those two electrons move to form a double bond between the two
carbon atoms.
The approach of those electrons repels the electrons in the carbonbromine bond right out onto the bromine, throwing the bromine off
as a negative ion.

The attack could equally well have been on any of the other hydrogens
on the left-hand carbon, or on any on the right-hand one - it simply
depends on what the OH- ion hit.

Beyond A' level: The mechanism being described here is


known as an E2 mechanism. The E stands for elimination,
and the 2 is because the initial slow part of the reaction
involves 2 species. This mechanism is used by primary and
secondary halogenoalkanes.
There is also an E1 mechanism which would apply to tertiary
halogenoalkanes. (Secondary ones will also do it to some
extent.) No current A' level syllabus is likely to ask you about
that.

http://www.chemguide.co.uk/mechanisms/elim/elimtt.html (2 of 3)30/12/2004 11:08:54

Explaining the elimination reaction producing propene from 2-bromopropane

Where would you like to go now?


How to deal with more complicated cases . . .
The competition between elimination and substitution . . .
To menu of elimination reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/elim/elimtt.html (3 of 3)30/12/2004 11:08:54

elimination from unsymmetric halogenoalkanes

ELIMINATION FROM UNSYMMETRIC


HALOGENOALKANES

This page looks at elimination from


unsymmetric halogenoalkanes such as 2bromobutane.

2-bromobutane is an unsymmetric halogenoalkane in the sense that it


has a CH3 group one side of the C-Br bond and a CH2CH3 group the
other.
You have to be careful with compounds like this because of the
possibility of more than one elimination product depending on where the
hydrogen is removed from.
The basic facts and mechanisms for these reactions are exactly the
same as with simple halogenoalkanes like 2-bromopropane. This page
only deals with the extra problems created by the possibility of more than
one elimination product.

Important! What follow assumes that you are familiar with


the mechanism for elimination from 2-bromopropane. If you
aren't, it is essential that you follow this link before you go on.

http://www.chemguide.co.uk/mechanisms/elim/elimunsym.html (1 of 5)30/12/2004 11:08:56

elimination from unsymmetric halogenoalkanes

Background to the mechanism


You will remember that elimination happens when a hydroxide ion (from,
for example, sodium hydroxide) acts as a base and removes a hydrogen
as a hydrogen ion from the halogenoalkane.
For example, in the simple case of elimination from 2-bromopropane:

The hydroxide ion removes a hydrogen from one of the carbon atoms
next door to the carbon-bromine bond, and the various electron shifts
then lead to the formation of the alkene - in this case, propene.
With an unsymmetric halogenoalkane like 2-bromobutane, there are
several hydrogens which might possibly get removed. You need to think
about each of these possibilities.
Where does the hydrogen get removed from?
The hydrogen has to be removed from a carbon atom adjacent to the
carbon-bromine bond. If an OH- ion hit one of the hydrogens on the righthand CH3 group in the 2-bromobutane (as we've drawn it), there's
nowhere for the reaction to go.

http://www.chemguide.co.uk/mechanisms/elim/elimunsym.html (2 of 5)30/12/2004 11:08:57

elimination from unsymmetric halogenoalkanes

To make room for the electron pair to form a double bond between the
carbons, you would have to expel a hydrogen from the CH2 group as a
hydride ion, H-. That is energetically much too difficult, and so this
reaction doesn't happen.
That still leaves the possibility of removing a hydrogen either from the lefthand CH3 or from the CH2 group.
If it was removed from the CH3 group:

The product is but-1-ene, CH2=CHCH2CH3.


If it was removed from the CH2 group:

This time the product is but-2-ene, CH3CH=CHCH3.


In fact the situation is even more complicated than it looks, because but2-ene exhibits geometric isomerism. You get a mixture of two isomers
formed - cis-but-2-ene and trans-but-2-ene.

http://www.chemguide.co.uk/mechanisms/elim/elimunsym.html (3 of 5)30/12/2004 11:08:57

elimination from unsymmetric halogenoalkanes

Which isomer gets formed is just a matter of chance.

Geometric isomerism: Isomerism is where you can draw


more than one arrangement of the atoms for a given
molecular formula. Geometric isomerism is a special case of
this involving molecules which have restricted rotation around
one of the bonds - in this case, a carbon-carbon double bond.
The C=C bond could only rotate if enough energy is put in to
break the pi bond. Effectively, except at high temperatures,
the C=C bond is "locked".
In the case of but-2-ene, the two CH3 groups will either both
be locked on one side of the C=C (to give the cis isomer), or
on opposite sides (to give the trans one).
For a full discussion of geometric isomerism follow this link.
Use the BACK button on your browser to return to this page.
Beware! It is easy to miss geometric isomers in an exam.
Always draw alkenes with the correct 120 bond angles
around the C=C bond as shown in the diagrams for the cis
and trans isomers above. If you take a short cut and write but2-ene as CH3CH=CHCH3, you will almost certainly miss the
fact that cis and trans forms are possible.
This is a rich source of questions in an exam. You could
easily throw away marks if you miss these possibilities.

http://www.chemguide.co.uk/mechanisms/elim/elimunsym.html (4 of 5)30/12/2004 11:08:57

elimination from unsymmetric halogenoalkanes

The overall result


Elimination from 2-bromobutane leads to a mixture containing:

but-1-ene

cis-but-2-ene

trans-but-2-ene

Where would you like to go now?


To menu of elimination reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/elim/elimunsym.html (5 of 5)30/12/2004 11:08:57

elimination v nucleophilic substitution in halogenoalkanes

ELIMINATION VERSUS SUBSTITUTION IN


HALOGENOALKANES

This page discusses the factors that decide whether halogenoalkanes


undergo elimination reactions or nucleophilic substitution when they react
with hydroxide ions from, say, sodium hydroxide or potassium hydroxide.
Details for each of these types of reaction are given elsewhere, and you
will find links to them from this page.

The reactions
Both reactions involve heating the halogenoalkane under reflux with
sodium or potassium hydroxide solution.
Nucleophilic substitution
The hydroxide ions present are good nucleophiles, and one possibility is
a replacement of the halogen atom by an -OH group to give an alcohol
via a nucleophilic substitution reaction.

In the example, 2-bromopropane is converted into propan-2-ol.

Note: If you want to read about nucleophilic substitution in


this reaction in detail, follow this link.

http://www.chemguide.co.uk/mechanisms/elim/elimvsubst.html (1 of 5)30/12/2004 11:09:00

elimination v nucleophilic substitution in halogenoalkanes

Elimination
Halogenoalkanes also undergo elimination reactions in the presence of
sodium or potassium hydroxide.

The 2-bromopropane has reacted to give an alkene - propene.


Notice that a hydrogen atom has been removed from one of the end
carbon atoms together with the bromine from the centre one. In all simple
elimination reactions the things being removed are on adjacent carbon
atoms, and a double bond is set up between those carbons.

Note: If you want to read about elimination in this reaction in


detail, follow this link.

What decides whether you get substitution or elimination?


The reagents you are using are the same for both substitution or
elimination - the halogenoalkane and either sodium or potassium
hydroxide solution. In all cases, you will get a mixture of both reactions
happening - some substitution and some elimination. What you get most
of depends on a number of factors.
The type of halogenoalkane
This is the most important factor.
type of halogenoalkane

substitution or elimination?

primary

mainly substitution

http://www.chemguide.co.uk/mechanisms/elim/elimvsubst.html (2 of 5)30/12/2004 11:09:00

elimination v nucleophilic substitution in halogenoalkanes

secondary

both substitution and elimination

tertiary

mainly elimination

Important! If you aren't clear about the various types of


halogenoalkanes, it is essential that you follow this link
before you read on.
Use the BACK button on your browser to return to this page.

For example, whatever you do with tertiary halogenoalkanes, you will


tend to get mainly the elimination reaction, whereas with primary ones
you will tend to get mainly substitution. However, you can influence
things to some extent by changing the conditions.
The solvent
The proportion of water to ethanol in the solvent matters.

Water encourages substitution.

Ethanol encourages elimination.

The temperature
Higher temperatures encourage elimination.
Concentration of the sodium or potassium hydroxide solution
Higher concentrations favour elimination.
In summary
For a given halogenoalkane, to favour elimination rather than
substitution, use:

heat

http://www.chemguide.co.uk/mechanisms/elim/elimvsubst.html (3 of 5)30/12/2004 11:09:00

elimination v nucleophilic substitution in halogenoalkanes

a concentrated solution of sodium or potassium hydroxide

pure ethanol as the solvent

Note: The explanations for these effects are well beyond the
demands of UK A level syllabuses. Some things you just
have to know!

The role of the hydroxide ions


The role of the hydroxide ion in a substitution reaction
In the substitution reaction between a
halogenoalkane and OH- ions, the hydroxide ions
are acting as nucleophiles. For example, one of the
lone pairs on the oxygen can attack the slightly
positive carbon. This leads on to the loss of the
bromine as a bromide ion, and the -OH group
becoming attached in its place.

The role of the hydroxide ion in an elimination reaction


Hydroxide ions have a very strong tendency to
combine with hydrogen ions to make water - in
other words, the OH- ion is a very strong base. In
an elimination reaction, the hydroxide ion hits one
of the hydrogen atoms in the CH3 group and pulls
it off. This leads to a cascade of electron pair
movements resulting in the formation of a carboncarbon double bond, and the loss of the bromine as Br-.

http://www.chemguide.co.uk/mechanisms/elim/elimvsubst.html (4 of 5)30/12/2004 11:09:00

elimination v nucleophilic substitution in halogenoalkanes

Note: The competition between substitution and elimination


(including the conditions needed and the mechanisms for
both) is a rich source of exam questions if your syllabus
includes it. You will probably find that the questions centre
around secondary halogenoalkanes like 2-bromopropane,
because these can easily be persuaded to do either reaction.
Important: If the elimination mechanism is on your syllabus,
you are quite likely to be asked questions in which the
substitution reaction crops up as well. If you need to revise
this in detail, now is a good time to do it.

Where would you like to go now?


To menu of elimination reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/elim/elimvsubst.html (5 of 5)30/12/2004 11:09:00

nucleophilic substitution - halogenoalkanes and hydroxide ions

THE NUCLEOPHILIC SUBSTITUTION


REACTIONS BETWEEN HALOGENOALKANES
AND HYDROXIDE IONS

This page gives you the facts and simple, uncluttered mechanisms for
the nucleophilic substitution reactions between halogenoalkanes and
hydroxide ions (from, for example, sodium hydroxide). If you want the
mechanisms explained to you in detail, there is a link at the bottom of the
page.

The reaction of primary halogenoalkanes with hydroxide


ions
Important! If you aren't sure about the difference between
primary, secondary and tertiary halogenoalkanes, it is
essential that you follow this link before you go on.
Use the BACK button on your browser to return to this page.

The facts
If a halogenoalkane is heated under reflux with a solution of sodium or
potassium hydroxide, the halogen is replaced by -OH and an alcohol is
produced. Heating under reflux means heating with a condenser placed
vertically in the flask to prevent loss of volatile substances from the
mixture.
The solvent is usually a 50/50 mixture of ethanol and water, because
everything will dissolve in that. The halogenoalkane is insoluble in water.
If you used water alone as the solvent, the halogenoalkane and the
sodium hydroxide solution wouldn't mix and the reaction could only
happen where the two layers met.

http://www.chemguide.co.uk/mechanisms/nucsub/hydroxide.html (1 of 6)30/12/2004 11:09:07

nucleophilic substitution - halogenoalkanes and hydroxide ions

For example, using 1-bromopropane as a typical primary halogenoalkane:

You could write the full equation rather than the ionic one, but it slightly
obscures what's going on:

The bromine (or other halogen) in the halogenoalkane is simply replaced


by an -OH group - hence a substitution reaction. In this example, propan1-ol is formed.
The mechanism
Here is the mechanism for the reaction involving bromoethane:

This is an example ofnucleophilic substitution.


Because the mechanism involves collision between two species in the
slow step (in this case, the only step) of the reaction, it is known as an
SN2 reaction.

Note: Unless your syllabus specifically mentions SN2 by


name, you can just call it nucleophilic substitution.

http://www.chemguide.co.uk/mechanisms/nucsub/hydroxide.html (2 of 6)30/12/2004 11:09:07

nucleophilic substitution - halogenoalkanes and hydroxide ions

If your examiners want you to show the transition state, draw the
mechanism like this:

The reaction of tertiary halogenoalkanes with hydroxide


ions
The facts
The facts of the reaction are exactly the same as with primary
halogenoalkanes. If the halogenoalkane is heated under reflux with a
solution of sodium or potassium hydroxide in a mixture of ethanol and
water, the halogen is replaced by -OH, and an alcohol is produced.
For example:

Or if you want the full equation rather than the ionic one:

The mechanism
http://www.chemguide.co.uk/mechanisms/nucsub/hydroxide.html (3 of 6)30/12/2004 11:09:07

nucleophilic substitution - halogenoalkanes and hydroxide ions

This mechanism involves an initial ionisation of the halogenoalkane:

followed by a very rapid attack by the hydroxide ion on the carbocation


(carbonium ion) formed:

This is again an example ofnucleophilic substitution.


This time the slow step of the reaction only involves one species - the
halogenoalkane. It is known as an SN1 reaction.

The reaction of secondary halogenoalkanes with


hydroxide ions
The facts
The facts of the reaction are exactly the same as with primary or tertiary
halogenoalkanes. The halogenoalkane is heated under reflux with a
solution of sodium or potassium hydroxide in a mixture of ethanol and
water.
For example:

The mechanism

http://www.chemguide.co.uk/mechanisms/nucsub/hydroxide.html (4 of 6)30/12/2004 11:09:07

nucleophilic substitution - halogenoalkanes and hydroxide ions

Secondary halogenoalkanes use both SN2 and SN1 mechanisms. For


example, the SN2 mechanism is:

Should you need it, the two stages of the SN1 mechanism are:

Note: There is another reaction between halogenoalkanes


and hydroxide ions involving an elimination reaction. If you
want to explore that as an alternative, follow this link.

Where would you like to go now?


Help! Talk me through these mechanisms . . .
To menu of nucleophilic substitution reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

http://www.chemguide.co.uk/mechanisms/nucsub/hydroxide.html (5 of 6)30/12/2004 11:09:07

nucleophilic substitution - halogenoalkanes and hydroxide ions

Jim Clark 2000 (modified 2004)

http://www.chemguide.co.uk/mechanisms/nucsub/hydroxide.html (6 of 6)30/12/2004 11:09:07

Types of halogenoalkanes

HALOGENOALKANES
Halogenoalkanes are also called haloalkanes or alkyl halides.
All halogenoalkanes contain a halogen atom - fluorine, chlorine, bromine
or iodine - attached to an alkyl group.

Note: An alkyl group is a group such as methyl, CH3, or


ethyl, CH3CH2. These are groups containing chains of
carbon atoms which may be branched. Alkyl groups are
given the general symbol R.

The different kinds of halogenoalkanes


Primary halogenoalkanes
In a primary (1) halogenoalkane, the carbon which carries the halogen
atom is only attached to one other alkyl group.
Some examples of primary halogenoalkanes include:

Notice that it doesn't matter how complicated the attached alkyl group is.
In each case there is only one linkage from the CH2 group holding the
halogen to an alkyl group.
There is an exception to this. CH3Br and the other methyl halides are
often counted as primary halogenoalkanes even though there are no
alkyl groups attached to the carbon with the halogen on it.
Secondary halogenoalkanes
In a secondary (2) halogenoalkane, the carbon with the halogen
attached is joined directly to two other alkyl groups, which may be the
http://www.chemguide.co.uk/mechanisms/nucsub/types.html (1 of 2)30/12/2004 11:09:10

Types of halogenoalkanes

same or different.
Examples:

Tertiary halogenoalkanes
In a tertiary (3) halogenoalkane, the carbon atom holding the halogen is
attached directly to three alkyl groups, which may be any combination of
same or different.
Examples:

Where would you like to go now?


To menu of nucleophilic substitution reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/nucsub/types.html (2 of 2)30/12/2004 11:09:10

Nucleophilic Substitution Mechanisms Menu

Understanding Chemistry

NUCLEOPHILIC SUBSTITUTION MECHANISMS


MENU
Types of halogenoalkanes . . .
Desribes what is meant by primary, secondary and tertiary
halogenoalkanes. Essential knowledge if you are to make sense of
everything else in this section.
What is nucleophilic substitution? . . .
Includes background material on the bonding in halogenoalkanes,
and general mechanisms for their nucleophilic substitution
reactions.
Substitution reactions involving hydroxide ions . . .
The mechanisms for the formation of alcohols from
halogenoalkanes by reaction with hydroxide ions.
Substitution reactions involving water . . .
The mechanisms for the formation of alcohols from
halogenoalkanes by reaction with water.
Substitution reactions involving cyanide ions . . .
The mechanisms for the formation of nitriles from halogenoalkanes
by reaction with cyanide ions.
Substitution reactions involving ammonia . . .
The mechanisms for the formation of various sorts of amines from

http://www.chemguide.co.uk/mechanisms/nucsubmenu.html (1 of 2)30/12/2004 11:09:13

Nucleophilic Substitution Mechanisms Menu

halogenoalkanes by reaction with ammonia. Also includes the


reactions between halogenoalkanes and amines.

Go to menu of other types of mechanism. . .


Go to Main Menu . . .

You might also be interested in:


properties and reactions of halogenoalkanes . . .
Covers all the physical and chemical properties of halogenoalkanes
(haloalkanes or alkyl halides) required by UK A level syllabuses.

Jim Clark 2000 (modified 2004)

http://www.chemguide.co.uk/mechanisms/nucsubmenu.html (2 of 2)30/12/2004 11:09:13

What is nucleophilic substitution?

NUCLEOPHILIC SUBSTITUTION

Background
Bonding in the halogenoalkanes
Halogenoalkanes (also known as haloalkanes or alkyl halides) are
compounds containing a halogen atom (fluorine, chlorine, bromine or
iodine) joined to one or more carbon atoms in a chain.
The interesting thing about these compounds is the carbon-halogen
bond, and all the nucleophilic substitution reactions of the
halogenoalkanes involve breaking that bond.
The polarity of the carbon-halogen bonds
With the exception of iodine, all of the halogens are more electronegative
than carbon.
Electronegativity values (Pauling scale)
C

2.5

F
Cl
Br
I

4.0
3.0
2.8
2.5

Note: If you aren't sure about electronegativity and bond


polarity follow this link before you read on.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/mechanisms/nucsub/whatis.html (1 of 12)30/12/2004 11:09:21

What is nucleophilic substitution?

That means that the electron pair in the carbon-halogen bond will be
dragged towards the halogen end, leaving the halogen slightly negative
( -) and the carbon slightly positive ( +) - except in the carbon-iodine
case.
Although the carbon-iodine bond doesn't have a permanent dipole, the
bond is very easily polarised by anything approaching it. Imagine a
negative ion approaching the bond from the far side of the carbon atom:

The fairly small polarity of the carbon-bromine bond will be increased by


the same effect.
The strengths of the carbon-halogen bonds

Note: If you haven't done any work on bond strengths, or are


a bit rusty, it doesn't matter. Just realise that the bigger the
number, the stronger the bond. And don't worry if you have
found slightly different numbers in a different data source there is a lot of variability in the quoted values, but the overall
pattern is still the same.

http://www.chemguide.co.uk/mechanisms/nucsub/whatis.html (2 of 12)30/12/2004 11:09:21

What is nucleophilic substitution?

Look at the strengths of various bonds (all values in kJ mol-1).


C-H

413

C-F
C-Cl
C-Br
C-I

467
346
290
228

In all of these nucleophilic substitution reactions, the carbon-halogen


bond has to be broken at some point during the reaction. The harder it is
to break, the slower the reaction will be.
The carbon-fluorine bond is very strong (stronger than C-H) and isn't
easily broken. It doesn't matter that the carbon-fluorine bond has the
greatest polarity - the strength of the bond is much more important in
determining its reactivity. You might therefore expect fluoroalkanes to be
very unreactive - and they are! We shall simply ignore them from now on.
In the other halogenoalkanes, the bonds get weaker as you go from
chlorine to bromine to iodine.
That means that chloroalkanes react most slowly, bromoalkanes react
faster, and iodoalkanes react faster still.
Rates of reaction: RCl < RBr < RI
Where "<" is read as "is less than" - or, in this instance, "is slower than",
and R represents any alkyl group.

http://www.chemguide.co.uk/mechanisms/nucsub/whatis.html (3 of 12)30/12/2004 11:09:21

What is nucleophilic substitution?

Warning! Before you read on it is essential that you know


exactly what your syllabus says about these reactions so that
you can extract the right amount of detail from what follows.
The problem is that there are two different mechanisms
depending on the type of halogenoalkane you are using
(whether primary, secondary or tertiary). Some syllabuses try
to make things simpler by restricting you to just one of the
two mechanisms. Where the syllabus is vague, look at recent
exam papers and mark schemes, or any support material
published by the Exam Board.
Haven't got a syllabus or recent exam papers? Follow this
link to find out how to get them. You must know what your
examiners expect in this topic.

Nucleophilic substitution in primary halogenoalkanes


You will need to know about this if your syllabus talks about "primary
halogenoalkanes" or about SN2 reactions. If the syllabus is vague, check
recent exam papers and mark schemes, and compare them against what
follows.
Nucleophiles
A nucleophile is a species (an ion or a molecule) which is strongly
attracted to a region of positive charge in something else.
Nucleophiles are either fully negative ions, or else have a strongly charge somewhere on a molecule. The common nucleophiles for A' level
purposes are hydroxide ions, cyanide ions, water and ammonia.

Notice that each of these contains at least one lone pair of electrons,
http://www.chemguide.co.uk/mechanisms/nucsub/whatis.html (4 of 12)30/12/2004 11:09:21

What is nucleophilic substitution?

either on an atom carrying a full negative charge, or on a very


electronegative atom carrying a substantial - charge.
The nucleophilic substitution reaction - an SN2 reaction
We'll talk this mechanism through using an ion as a nucleophile, because
it's slightly easier. The water and ammonia mechanisms involve an extra
step which you can read about on the pages describing those particular
mechanisms.
We'll take bromoethane as a typical primary
halogenoalkane. The bromoethane has a polar bond
between the carbon and the bromine.
We'll look at its reaction with a general purpose
nucleophilic ion which we'll call Nu-. This will have at least one lone pair
of electrons. Nu- could, for example, be OH- or CN-.

The lone pair on the Nu- ion will be strongly attracted to the + carbon,
and will move towards it, beginning to make a co-ordinate (dative
covalent) bond. In the process the electrons in the C-Br bond will be
pushed even closer towards the bromine, making it increasingly negative.

Note: A co-ordinate bond is a covalent bond in which both


electrons come from one of the atoms.

The movement goes on until the -Nu is firmly attached to the carbon, and
the bromine has been expelled as a Br- ion.

http://www.chemguide.co.uk/mechanisms/nucsub/whatis.html (5 of 12)30/12/2004 11:09:21

What is nucleophilic substitution?

Note: We haven't shown all the lone pairs on the bromine.


These other lone pairs aren't involved in the reaction, and
including them simply clutters the diagram to no purpose.

Things to notice
The Nu- ion approaches the + carbon from the side away from the
bromine atom. The large bromine atom hinders attack from its side and,
being -, would repel the incoming Nu- anyway. This attack from the
back is important if you need to understand why tertiary halogenoalkanes
have a different mechanism. We'll discuss this later on this page.
There is obviously a point in which the Nu- is half attached to the carbon,
and the C-Br bond is half way to being broken. This is called a transition
state. It isn't an intermediate. You can't isolate it - even for a very short
time. It's just the mid-point of a smooth attack by one group and the
departure of another.
How to write the mechanism
The simplest way is:

http://www.chemguide.co.uk/mechanisms/nucsub/whatis.html (6 of 12)30/12/2004 11:09:21

What is nucleophilic substitution?

Note: In exam you must show the lone pair of electrons on


the nucleophile (in this case, the Nu- ion). It probably doesn't
matter whether you show them on the departing Br- ion or not.
If you aren't happy about the use of curly arrows in
mechanisms, follow this link before you go on. Use the BACK
button on your browser to return to this page.

Technically, this is known as an SN2 reaction. S stands for substitution, N


for nucleophilic, and the 2 is because the initial stage of the reaction
involves two species - the bromoethane and the Nu- ion. If your syllabus
doesn't refer to SN2 reactions by name, you can just call it nucleophilic
substitution.

Some examiners like you to show the transition state in the mechanism,
in which case you need to write it in a bit more detail - showing how
everything is arranged in space.

Be very careful when you draw the transition state to make a clear

http://www.chemguide.co.uk/mechanisms/nucsub/whatis.html (7 of 12)30/12/2004 11:09:21

What is nucleophilic substitution?

difference between the dotted lines showing the half-made and halfbroken bonds, and those showing the bonds going back into the paper.
Notice that the molecule has been inverted during the reaction - rather
like an umbrella being blown inside-out.

Note: If you aren't happy about the various ways of drawing


bonds, it is important to follow this link to find out exactly what
the various symbols mean.
It is also important to know which of these ways of drawing
the mechanism your particular examiners want you to use. If
you haven't already checked your syllabus, recent exam
papers and mark schemes, you must do so! At the time of
writing, Edexcel, for example, wanted the transition state
included, and that isn't obvious from their syllabus. You have
to check mark schemes and examiners reports.
Use the BACK button on your browser to return to this page.

Nucleophilic substitution in tertiary halogenoalkanes

Warning! Check your syllabus, past papers and any support


material published by your Exam Board to find out whether
you need this. If there's no mention of tertiary
halogenoalkanes or SN1 reactions, then you probably don't
need it.

http://www.chemguide.co.uk/mechanisms/nucsub/whatis.html (8 of 12)30/12/2004 11:09:21

What is nucleophilic substitution?

Remember that a tertiary halogenoalkane has three


alkyl groups attached to the carbon with the halogen
on it. These alkyl groups can be the same or
different, but in this section, we shall just consider a simple one, (CH3)
3CBr - 2-bromo-2-methylpropane.
The nucleophilic substitution reaction - an SN1 reaction
Once again, we'll talk this mechanism through using an ion as a
nucleophile, because it's slightly easier, and again we'll look at the
reaction of a general purpose nucleophilic ion which we'll call Nu-. This
will have at least one lone pair of electrons.
Why is a different mechanism necessary?
You will remember that when a nucleophile attacks a primary
halogenoalkane, it approaches the + carbon atom from the side away
from the halogen atom.
With a tertiary halogenoalkane, this is impossible. The back of the
molecule is completely cluttered with CH3 groups.

Since any other approach is prevented by the bromine atom, the reaction
has to go by an alternative mechanism.
The alternative mechanism

http://www.chemguide.co.uk/mechanisms/nucsub/whatis.html (9 of 12)30/12/2004 11:09:21

What is nucleophilic substitution?

Important! To understand this section, you need to know


what a carbocation (carbonium ion) is, and about the relative
stabilities of primary, secondary and tertiary carbocations.
If you follow this link, use the BACK button on your browser
to return to this page.

The reaction happens in two stages. In the first, a small proportion of the
halogenoalkane ionises to give a carbocation and a bromide ion.

This reaction is possible because tertiary carbocations are relatively


stable compared with secondary or primary ones. Even so, the reaction
is slow.
Once the carbocation is formed, however, it would react immediately it
came into contact with a nucleophile like Nu-. The lone pair on the
nucleophile is strongly attracted towards the positive carbon, and moves
towards it to create a new bond.

How fast the reaction happens is going to be governed by how fast the
halogenoalkane ionises. Because this initial slow step only involves one
species, the mechanism is described as SN1 - substitution, nucleophilic,
one species taking part in the initial slow step.
Why don't primary halogenoalkanes use the SN1 mechanism?
If a primary halogenoalkane did use this mechanism, the first step would
http://www.chemguide.co.uk/mechanisms/nucsub/whatis.html (10 of 12)30/12/2004 11:09:21

What is nucleophilic substitution?

be, for example:

A primary carbocation would be formed, and this is much more


energetically unstable than the tertiary one formed from tertiary
halogenoalkanes - and therefore much more difficult to produce.
This instability means that there will be a very high activation energy for
the reaction involving a primary halogenoalkane. The activation energy is
much less if it undergoes an SN2 reaction - and so that's what it does
instead.

Nucleophilic substitution in secondary halogenoalkanes


There isn't anything new in this. Secondary
halogenoalkanes will use both mechanisms - some
molecules will react using the SN2 mechanism and
others the SN1.

The SN2 mechanism is possible because the back of the molecule isn't
completely cluttered by alkyl groups and so the approaching nucleophile
can still get at the + carbon atom.
The SN1 mechanism is possible because the secondary carbocation
formed in the slow step is more stable than a primary one. It isn't as
stable as a tertiary one though, and so the SN1 route isn't as effective as
it is with tertiary halogenoalkanes.

Where would you like to go now?


To menu of nucleophilic substitution reactions. . .

http://www.chemguide.co.uk/mechanisms/nucsub/whatis.html (11 of 12)30/12/2004 11:09:21

What is nucleophilic substitution?

To menu of other types of mechanism. . .


To Main Menu . . .

Jim Clark 2000 (modified 2004)

http://www.chemguide.co.uk/mechanisms/nucsub/whatis.html (12 of 12)30/12/2004 11:09:21

Explaining nucleophilic substitution between halogenoalkanes and hydroxide ions

EXPLAINING THE NUCLEOPHILIC


SUBSTITUTION REACTIONS BETWEEN
HALOGENOALKANES AND HYDROXIDE IONS

This page guides you through the nucleophilic substitution mechanisms


for the reactions between halogenoalkanes and hydroxide ions from, for
example, sodium hydroxide.

Important! It would help if you first read the page What is


nucleophilic substitution? before you go on.
You must also be clear about the differences between
primary, secondary and tertiary halogenoalkanes.

The reactions between primary or secondary


halogenoalkanes and hydroxide ions - the SN2 mechanism
Hydroxide ions as nucleophiles
A nucleophile is a species (an ion or a molecule) which is strongly
attracted to a region of positive charge in something else.
Nucleophiles are either fully negative ions, or else have a strongly charge somewhere on a molecule.
In the case of the hydroxide ion, there is a full negative
charge on the oxygen, as well as three lone pairs of
electrons.
The nucleophilic substitution reaction - an SN2 reaction

http://www.chemguide.co.uk/mechanisms/nucsub/hydroxidett.html (1 of 7)30/12/2004 11:09:26

Explaining nucleophilic substitution between halogenoalkanes and hydroxide ions

We'll talk this reaction through with a primary


halogenoalkane to start with, taking bromoethane as
typical. The bromoethane has a polar bond between
the carbon and the bromine.

Note: In an exam you must show the lone pair of electrons


on the nucleophile (in this case, the OH- ion). It probably
doesn't matter whether you show them on the departing Brion or not.

One of the lone pairs on the OH- ion will be strongly attracted to the +
carbon, and will move towards it, beginning to form a bond with it. The
approaching negative ion will repel the electrons in the carbon-bromine
bond closer and closer to the bromine.
At some point during this, the -OH group and the bromine will both be
half-attached to the carbon. This is called the transition state for the
reaction. It isn't an intermediate - you can't isolate it and it doesn't have
any independent existence. It's just the half-way stage of a smooth
movement of atoms and electrons.
The movement goes on until the -OH is firmly attached to the carbon,
and the bromine has been expelled as a Br- ion.
You may need to show the formation of the intermediate in the
mechanism (depending on what your examiners want). It simply needs
you to draw the mechanism showing some more detail about how the
various groups are arranged in space.

http://www.chemguide.co.uk/mechanisms/nucsub/hydroxidett.html (2 of 7)30/12/2004 11:09:26

Explaining nucleophilic substitution between halogenoalkanes and hydroxide ions

Be very careful when you draw the transition state to make a clear
difference between the dotted lines showing the half-made and halfbroken bonds, and those showing the bonds going back into the paper.
Notice that the molecule has been inverted during the reaction - rather
like an umbrella being blown inside-out.

Note: If you aren't happy about the various ways of drawing


bonds, it is important to follow this link to find out exactly what
the various symbols mean.
It is also important to know which of these ways of drawing
the mechanism your particular examiners want you to use. If
you haven't already checked your syllabus, recent exam
papers and mark schemes, you must do so! At the time of
writing, Edexcel, for example, wanted the transition state
included, and that isn't obvious from their syllabus. You have
to check mark schemes and examiners reports.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/mechanisms/nucsub/hydroxidett.html (3 of 7)30/12/2004 11:09:26

Explaining nucleophilic substitution between halogenoalkanes and hydroxide ions

Technically, this is known as an SN2 reaction. S stands for substitution, N


for nucleophilic, and the 2 is because the initial stage of the reaction
involves two species - the bromoethane and the OH- ion. If your syllabus
doesn't refer to SN2 reactions by name, you can just call it nucleophilic
substitution.
The SN2 reaction in secondary halogenoalkanes
The reaction can happen in exactly the same way with a secondary
halogenoalkane, although they also have the potential for reacting via a
different mechanism (which we'll deal with shortly).

Again, a lone pair on the approaching hydroxide ion forms a bond with
the + carbon and, in the process, the electrons in the carbon-bromine
bond are forced entirely onto the bromine to create a bromide ion.

The reactions between secondary or tertiary


halogenoalkanes and hydroxide ions - the SN1 mechanism
Note: Are you sure your syllabus wants this? Several
syllabuses restrict you to primary halogenoalkanes.
Haven't got a syllabus? Follow this link to find out how to get
one.

http://www.chemguide.co.uk/mechanisms/nucsub/hydroxidett.html (4 of 7)30/12/2004 11:09:26

Explaining nucleophilic substitution between halogenoalkanes and hydroxide ions

To start with, we'll talk this mechanism through with a


simple tertiary halogenoalkane like the one on the
right (2-bromo-2-methylpropane).
Why do tertiary halogenoalkanes need a different mechanism?
When a nucleophile attacks a primary halogenoalkane, it approaches the
+ carbon atom from the side away from the halogen atom. Any other
approach is prevented by the halogen atom, which is both bulky and
slightly negatively charged. The charge repels the incoming nucleophile.
With a tertiary halogenoalkane, this approach from the back is
impossible. The back of the molecule is completely cluttered with CH3
groups.
The SN1 mechanism
The reaction happens in two stages. In the first, a small proportion of the
halogenoalkane ionises to give a carbocation (carbonium ion) and a
bromide ion.

This reaction is possible because tertiary carbocations are relatively


stable compared with secondary or primary ones. Even so, the reaction
is slow.

Note: Not sure about the stability of carbocations


(carbonium ions)? Follow this link if you need to find out.

http://www.chemguide.co.uk/mechanisms/nucsub/hydroxidett.html (5 of 7)30/12/2004 11:09:26

Explaining nucleophilic substitution between halogenoalkanes and hydroxide ions

Once the carbocation is formed, however, it would react immediately it


came into contact with an OH- ion. The lone pair on the nucleophile is
strongly attracted towards the positive carbon, and moves towards it to
create a new bond.

How fast the reaction happens is going to be governed by how fast the
halogenoalkane ionises - because that's a slow process. Because this
initial slow step only involves one species, the mechanism is described
as SN1 - substitution, nucleophilic, one species taking part in the initial
slow step.
The SN1 mechanism in secondary halogenoalkanes
Secondary halogenoalkanes (like 2-bromopropane) can use either the
SN1 or the SN2 mechanism. The back of the molecule is rather more
cluttered than in a primary halogenoalkane, but there is still room for the
lone pair on the nucleophile to approach and form a bond. We've already
dealt with that reaction.
It is also possible to get some slight ionisation of the halogenoalkane to
give an SN1 mechanism, but this reaction is much less successful than
with tertiary halogenoalkanes, because the secondary carbocation
formed isn't as stable as a tertiary one.

Once the carbocation has been formed, it will react immediately with a
hydroxide ion. A lone pair on the hydroxide ion is strongly attracted to the
positive carbon, moves towards it, and forms a bond.

http://www.chemguide.co.uk/mechanisms/nucsub/hydroxidett.html (6 of 7)30/12/2004 11:09:26

Explaining nucleophilic substitution between halogenoalkanes and hydroxide ions

Where would you like to go now?


To menu of nucleophilic substitution reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000 (modified 2004)

http://www.chemguide.co.uk/mechanisms/nucsub/hydroxidett.html (7 of 7)30/12/2004 11:09:26

nucleophilic substitution - halogenoalkanes and water

THE NUCLEOPHILIC SUBSTITUTION


REACTIONS BETWEEN HALOGENOALKANES
AND WATER

This page gives you the facts and simple, uncluttered mechanisms for
the nucleophilic substitution reactions between halogenoalkanes and
water. If you want the mechanisms explained to you in detail, there is a
link at the bottom of the page.

The reaction of primary halogenoalkanes with water


Important! If you aren't sure about the difference between
primary, secondary and tertiary halogenoalkanes, it is
essential that you follow this link before you go on.
Use the BACK button on your browser to return to this page.

The facts
There is only a slow reaction between a primary halogenoalkane and
water even if they are heated. The halogen atom is replaced by -OH.
For example, using 1-bromoethane as a typical primary halogenoalkane:

An alcohol is produced together with hydrobromic acid. Be careful not to


call this hydrogen bromide. Hydrogen bromide is a gas. When it is
dissolved it in water (as it will be here), it's called hydrobromic acid.
The mechanism

http://www.chemguide.co.uk/mechanisms/nucsub/water.html (1 of 5)30/12/2004 11:09:34

nucleophilic substitution - halogenoalkanes and water

The mechanism involves two steps. The first is a simple nucleophilic


substitution reaction:

Because the mechanism involves collision between two species in this


slow step of the reaction, it is known as an SN2 reaction.

Note: Unless your syllabus specifically mentions SN2 by


name, you can just call it nucleophilic substitution.

The nucleophilic substitution is very slow because water isn't a very good
nucleophile. It lacks the full negative charge of, say, a hydroxide ion.
The second step of the reaction simply tidies up the product. A water
molecule removes one of the hydrogens attached to the oxygen to give
an alcohol and a hydroxonium ion (also known as a hydronium ion or an
oxonium ion).

The hydroxonium ion and the bromide ion (from the nucleophilic
substitution stage of the reaction) make up the hydrobromic acid which is
formed as well as the alcohol.

The reaction of tertiary halogenoalkanes with water


The facts
If the halogenoalkane is heated under reflux with water, the halogen is
http://www.chemguide.co.uk/mechanisms/nucsub/water.html (2 of 5)30/12/2004 11:09:34

nucleophilic substitution - halogenoalkanes and water

replaced by -OH to give an alcohol. Heating under reflux means heating


with a condenser placed vertically in the flask to prevent loss of volatile
substances from the mixture. The reaction happens much faster than the
corresponding one involving a primary halogenoalkane.
For example:

The mechanism
This mechanism involves an initial ionisation of the halogenoalkane:

followed by a very rapid attack by the water on the carbocation


(carbonium ion) formed:

This is again an example ofnucleophilic substitution.


This time the slow step of the reaction only involves one species - the
halogenoalkane. It is known as an SN1 reaction.
Now there is a final stage in which the product is tidied up. A water
molecule removes one of the hydrogens attached to the oxygen to give
an alcohol and a hydroxonium ion - exactly as happens with primary
halogenoalkanes.

http://www.chemguide.co.uk/mechanisms/nucsub/water.html (3 of 5)30/12/2004 11:09:34

nucleophilic substitution - halogenoalkanes and water

The rate of the overall reaction is governed entirely by how fast the
halogenoalkane ionises. The fact that water isn't as good a nucleophile
as, say, OH- doesn't make any difference. The water isn't involved in the
slow step of the reaction.

The reaction of secondary halogenoalkanes with water


No current A' level syllabus is likely to ask you about this. In the
extremely unlikely event that you will ever need it, secondary
halogenoalkanes use both an SN2 mechanism and an SN1.
Make sure you understand what happens with primary and tertiary
halogenoalkanes, and then adapt it for secondary ones should ever need
to.

Where would you like to go now?


Help! Talk me through these mechanisms . . .
To menu of nucleophilic substitution reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

http://www.chemguide.co.uk/mechanisms/nucsub/water.html (4 of 5)30/12/2004 11:09:34

nucleophilic substitution - halogenoalkanes and water

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/nucsub/water.html (5 of 5)30/12/2004 11:09:34

Explaining nucleophilic substitution between halogenoalkanes and water

EXPLAINING THE NUCLEOPHILIC


SUBSTITUTION REACTIONS BETWEEN
HALOGENOALKANES AND WATER

This page guides you through the nucleophilic substitution mechanisms


for the reactions between halogenoalkanes and water. It deals only with
primary and tertiary halogenoalkanes. No current A' level syllabus is
likely to ask about the reaction between secondary halogenoalkanes and
water. (It's not difficult - it's just not there!)

Important! It would help if you first read the page What is


nucleophilic substitution? before you go on.
You must also be clear about the differences between
primary, secondary and tertiary halogenoalkanes.

The reaction between primary halogenoalkanes and water


- the SN2 mechanism
Water as a nucleophile
A nucleophile is a species (an ion or a molecule) which is strongly
attracted to a region of positive charge in something else.
Nucleophiles are either fully negative ions, or else have a strongly charge somewhere on a molecule.
Water obviously doesn't carry a negative charge.
However, oxygen is much more electronegative than
hydrogen, and so the oxygen atom has a fairly
substantial - charge to back up its two lone pairs.

http://www.chemguide.co.uk/mechanisms/nucsub/watertt.html (1 of 6)30/12/2004 11:09:36

Explaining nucleophilic substitution between halogenoalkanes and water

Note: If you aren't sure about electronegativity follow this link


before you read on.
Use the BACK button on your browser to return to this page.

The attack on the halogenoalkane is therefore by one of the lone pairs on


the oxygen. Because there isn't a full negative charge, water isn't going
to be as good a nucleophile as a negative ion like OH-, and so the
reaction is slower.
The nucleophilic substitution reaction - an SN2 reaction
We'll talk this reaction through with bromoethane as a
typical primary halogenoalkane. The bromoethane has
a polar bond between the carbon and the bromine.

One of the lone pairs on the water will be strongly attracted to the +
carbon, and will move towards it, beginning to form a bond with it. The
approaching lone pair will repel the electrons in the carbon-bromine bond
closer and closer to the bromine.
The movement goes on until the water is firmly attached to the carbon,
and the bromine has been expelled as a Br- ion.
Notice that the oxygen in the product ion carries a positive charge
(highlighted in red to draw attention to it). That charge has to be there for
two reasons:

On the left hand side of the equation you start with two overall
neutral molecules. Assuming you forgot about the positive charge,

http://www.chemguide.co.uk/mechanisms/nucsub/watertt.html (2 of 6)30/12/2004 11:09:36

Explaining nucleophilic substitution between halogenoalkanes and water

you would end up with a neutral species and a negative ion on the
right. Charges must balance in equations, so something is wrong.

The oxygen looks wrong! The oxygen atom is joined to 3 things


rather than its usual 2. Oxygen can only join to 3 things if it carries
a positive charge.

Note: Oxygen with a positive charge has the same


arrangement of electrons as a nitrogen atom - which normally
forms 3 bonds.

Technically, this is known as an SN2 reaction. S stands for substitution, N


for nucleophilic, and the 2 is because the initial stage of the reaction
involves two species - the bromoethane and the water. If your syllabus
doesn't refer to SN2 reactions by name, you can just call it nucleophilic
substitution.
Finally a hydrogen ion is pulled off the product ion by another water
molecule from the solution. A lone pair on the new water molecule forms
a bond with a hydrogen atom, forcing the bonding pair of electrons back
on to the positive oxygen. That cancels the positive charge.

The organic product is ethanol. The formula is distorted in this diagram


so that you can clearly see the relationship between the atoms on both
sides of the equation.
The other product is a hydroxonium ion (also known as a hydronium ion
or an oxonium ion). This is just a hydrogen ion attached to a water
molecule - what is often written as H+(aq).

The reaction between tertiary halogenoalkanes and water http://www.chemguide.co.uk/mechanisms/nucsub/watertt.html (3 of 6)30/12/2004 11:09:36

Explaining nucleophilic substitution between halogenoalkanes and water

the SN1 mechanism


We'll talk this mechanism through with a simple
tertiary halogenoalkane like the one on the right (2bromo-2-methylpropane).
Why do tertiary halogenoalkanes need a different mechanism?
When a nucleophile attacks a primary halogenoalkane, it approaches the
+ carbon atom from the side away from the halogen atom. Any other
approach is prevented by the halogen atom, which is both bulky and
slightly negatively charged. The charge repels the incoming nucleophile.
With a tertiary halogenoalkane, this approach from the back is
impossible. The back of the molecule is completely cluttered with CH3
groups.
The SN1 mechanism
In the first stage, a small proportion of the halogenoalkane ionises to give
a carbocation (carbonium ion) and a bromide ion.

This reaction is possible because tertiary carbocations are relatively


stable compared with secondary or primary ones. Even so, the reaction
is slow.

Note: Not sure about the stability of carbocations


(carbonium ions)? Follow this link if you need to find out.

http://www.chemguide.co.uk/mechanisms/nucsub/watertt.html (4 of 6)30/12/2004 11:09:36

Explaining nucleophilic substitution between halogenoalkanes and water

Once the carbocation is formed, however, it would react immediately it


came into contact with a water molecule. One of the lone pairs on the
water is strongly attracted towards the positive carbon, and moves
towards it to create a new bond.

How fast the reaction happens overall is going to be governed by how


fast the halogenoalkane ionises - because that's a slow process.
Because this initial slow step only involves one species, the mechanism
is described as SN1 - substitution, nucleophilic, one species taking part in
the initial slow step.
The water takes part in the fast step of the reaction, and so the fact that a
weakish nucleophile like water is involved doesn't significantly slow the
overall reaction down. The rate is determined by the slow ionisation of
the halogenoalkane.
As with primary halogenalkanes, there is a final stage to this reaction in
which a hydrogen ion is transferred from the organic ion to a water
molecule in the solution. What happens is exactly the same as with the
primary halogenoalkanes described above.

Where would you like to go now?


To menu of nucleophilic substitution reactions. . .
http://www.chemguide.co.uk/mechanisms/nucsub/watertt.html (5 of 6)30/12/2004 11:09:36

Explaining nucleophilic substitution between halogenoalkanes and water

To menu of other types of mechanism. . .


To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/nucsub/watertt.html (6 of 6)30/12/2004 11:09:36

nucleophilic substitution - halogenoalkanes and cyanide ions

THE NUCLEOPHILIC SUBSTITUTION


REACTIONS BETWEEN HALOGENOALKANES
AND CYANIDE IONS

This page gives you the facts and simple, uncluttered mechanisms for
the nucleophilic substitution reactions between halogenoalkanes and
cyanide ions (from, for example, potassium cyanide). If you want the
mechanisms explained to you in detail, there is a link at the bottom of the
page.

The reaction of primary halogenoalkanes with cyanide ions


Important! If you aren't sure about the difference between
primary, secondary and tertiary halogenoalkanes, it is
essential that you follow this link before you go on.
Use the BACK button on your browser to return to this page.

The facts
If a halogenoalkane is heated under reflux with a solution of sodium or
potassium cyanide in ethanol, the halogen is replaced by a -CN group
and a nitrile is produced. Heating under reflux means heating with a
condenser placed vertically in the flask to prevent loss of volatile
substances from the mixture.
The solvent is important. If water is present you tend to get substitution
by -OH instead of -CN.

http://www.chemguide.co.uk/mechanisms/nucsub/cyanide.html (1 of 6)30/12/2004 11:09:44

nucleophilic substitution - halogenoalkanes and cyanide ions

Note: A solution of potassium cyanide in water is quite


alkaline, and contains significant amounts of hydroxide ions.
These react with the halogenoalkane.

For example, using 1-bromopropane as a typical primary halogenoalkane:

You could write the full equation rather than the ionic one, but it slightly
obscures what's going on:

The bromine (or other halogen) in the halogenoalkane is simply replaced


by a -CN group - hence a substitution reaction. In this example,
butanenitrile is formed.

Note: When you are naming nitriles, you have to remember


to include the carbon in the -CN group when you count the
longest chain. In this example, there are 4 carbons in the
longest chain - hence butanenitrile.

The mechanism
Here is the mechanism for the reaction involving bromoethane:

This is an example ofnucleophilic substitution.


Because the mechanism involves collision between two species in the

http://www.chemguide.co.uk/mechanisms/nucsub/cyanide.html (2 of 6)30/12/2004 11:09:44

nucleophilic substitution - halogenoalkanes and cyanide ions

slow step (in this case, the only step) of the reaction, it is known as an
SN2 reaction.

Note: Unless your syllabus specifically mentions SN2 by


name, you can just call it nucleophilic substitution.

If your examiners want you to show the transition state, draw the
mechanism like this:

The reaction of tertiary halogenoalkanes with cyanide ions


The facts
The facts of the reaction are exactly the same as with primary
halogenoalkanes. If the halogenoalkane is heated under reflux with a
solution of sodium or potassium cyanide in ethanol, the halogen is
replaced by -CN, and a nitrile is produced.
For example:

http://www.chemguide.co.uk/mechanisms/nucsub/cyanide.html (3 of 6)30/12/2004 11:09:44

nucleophilic substitution - halogenoalkanes and cyanide ions

Or if you want the full equation rather than the ionic one:

The mechanism
This mechanism involves an initial ionisation of the halogenoalkane:

followed by a very rapid attack by the cyanide ion on the carbocation


(carbonium ion) formed:

This is again an example ofnucleophilic substitution.


This time the slow step of the reaction only involves one species - the
halogenoalkane. It is known as an SN1 reaction.

The reaction of secondary halogenoalkanes with cyanide


ions
The facts
The facts of the reaction are exactly the same as with primary or tertiary
halogenoalkanes. The halogenoalkane is heated under reflux with a
solution of sodium or potassium cyanide in ethanol.
For example:
http://www.chemguide.co.uk/mechanisms/nucsub/cyanide.html (4 of 6)30/12/2004 11:09:44

nucleophilic substitution - halogenoalkanes and cyanide ions

The mechanism
Secondary halogenoalkanes use both SN2 and SN1 mechanisms. For
example, the SN2 mechanism is:

Should you need it, the two stages of the SN1 mechanism are:

Where would you like to go now?


Help! Talk me through these mechanisms . . .
To menu of nucleophilic substitution reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

http://www.chemguide.co.uk/mechanisms/nucsub/cyanide.html (5 of 6)30/12/2004 11:09:44

nucleophilic substitution - halogenoalkanes and cyanide ions

Jim Clark 2000 (modified 2004)

http://www.chemguide.co.uk/mechanisms/nucsub/cyanide.html (6 of 6)30/12/2004 11:09:44

Explaining nucleophilic substitution between halogenoalkanes and cyanide ions

EXPLAINING THE NUCLEOPHILIC


SUBSTITUTION REACTIONS BETWEEN
HALOGENOALKANES AND CYANIDE IONS

This page guides you through the nucleophilic substitution mechanisms


for the reactions between halogenoalkanes and cyanide ions from, for
example, potassium cyanide.

Important! It would help if you first read the page What is


nucleophilic substitution? before you go on.
You must also be clear about the differences between
primary, secondary and tertiary halogenoalkanes.

The reactions between primary or secondary


halogenoalkanes and cyanide ions - the SN2 mechanism
Cyanide ions as nucleophiles
A nucleophile is a species (an ion or a molecule) which is strongly
attracted to a region of positive charge in something else.
Nucleophiles are either fully negative ions, or else have a strongly charge somewhere on a molecule.
In the case of the cyanide ion, there is a full negative
charge on the carbon, as well as a lone pair of electrons.

http://www.chemguide.co.uk/mechanisms/nucsub/cyanidett.html (1 of 7)30/12/2004 11:09:47

Explaining nucleophilic substitution between halogenoalkanes and cyanide ions

Note: There is a lone pair on the nitrogen atom as well, but


this isn't shown to avoid confusion. The one on the carbon is
more important in this instance because that's where the
negative charge is. The combination of the lone pair and the
negative charge makes the carbon end of the ion the
nucleophile.

The nucleophilic substitution reaction - an SN2 reaction


We'll talk this reaction through with a primary
halogenoalkane to start with, taking bromoethane as
typical. The bromoethane has a polar bond between
the carbon and the bromine.

Note: In an exam you must show the lone pair of electrons


on the nucleophile (in this case, the CN- ion). It probably
doesn't matter whether you show them on the departing Brion or not.

The lone pair on the carbon of the cyanide ion will be strongly attracted
to the + carbon, and will move towards it, beginning to form a bond with
it. The approaching negative ion will repel the electrons in the carbonbromine bond closer and closer to the bromine.
At some point during this, the -CN group and the bromine will both be
half-attached to the carbon. This is called the transition state for the
reaction. It isn't an intermediate - you can't isolate it and it doesn't have
any independent existence. It's just the half-way stage of a smooth
movement of atoms and electrons.
The movement goes on until the -CN is firmly attached to the carbon,
http://www.chemguide.co.uk/mechanisms/nucsub/cyanidett.html (2 of 7)30/12/2004 11:09:47

Explaining nucleophilic substitution between halogenoalkanes and cyanide ions

and the bromine has been expelled as a Br- ion.


You may need to show the formation of the intermediate in the
mechanism (depending on what your examiners want). It simply needs
you to draw the mechanism showing some more detail about how the
various groups are arranged in space.

Be very careful when you draw the transition state to make a clear
difference between the dotted lines showing the half-made and halfbroken bonds, and those showing the bonds going back into the paper.
Notice that the molecule has been inverted during the reaction - rather
like an umbrella being blown inside-out.

http://www.chemguide.co.uk/mechanisms/nucsub/cyanidett.html (3 of 7)30/12/2004 11:09:47

Explaining nucleophilic substitution between halogenoalkanes and cyanide ions

Note: If you aren't happy about the various ways of drawing


bonds, it is important to follow this link to find out exactly what
the various symbols mean.
It is also important to know which of these ways of drawing
the mechanism your particular examiners want you to use. If
you haven't already checked your syllabus, recent exam
papers and mark schemes, you must do so! At the time of
writing, Edexcel, for example, wanted the transition state
included, and that isn't obvious from their syllabus. You have
to check mark schemes and examiners reports.
Use the BACK button on your browser to return to this page.

Technically, this is known as an SN2 reaction. S stands for substitution, N


for nucleophilic, and the 2 is because the initial stage of the reaction
involves two species - the bromoethane and the CN- ion. If your syllabus
doesn't refer to SN2 reactions by name, you can just call it nucleophilic
substitution.
The SN2 reaction in secondary halogenoalkanes
The reaction can happen in exactly the same way with a secondary
halogenoalkane, although they also have the potential for reacting via a
different mechanism (which we'll deal with shortly).

Again, the lone pair on the approaching cyanide ion forms a bond with
the + carbon and, in the process, the electrons in the carbon-bromine
bond are forced entirely onto the bromine to create a bromide ion.

http://www.chemguide.co.uk/mechanisms/nucsub/cyanidett.html (4 of 7)30/12/2004 11:09:47

Explaining nucleophilic substitution between halogenoalkanes and cyanide ions

The reactions between secondary or tertiary


halogenoalkanes and cyanide ions - the SN1 mechanism
Note: Are you sure your syllabus wants this? Several
syllabuses restrict you to primary halogenoalkanes.
Haven't got a syllabus? Follow this link to find out how to get
one.

To start with, we'll talk this mechanism through with a


simple tertiary halogenoalkane like the one on the
right (2-bromo-2-methylpropane).
Why do tertiary halogenoalkanes need a different mechanism?
When a nucleophile attacks a primary halogenoalkane, it approaches the
+ carbon atom from the side away from the halogen atom. Any other
approach is prevented by the halogen atom, which is both bulky and
slightly negatively charged. The charge repels the incoming nucleophile.
With a tertiary halogenoalkane, this approach from the back is
impossible. The back of the molecule is completely cluttered with CH3
groups.
The SN1 mechanism
The reaction happens in two stages. In the first, a small proportion of the
halogenoalkane ionises to give a carbocation (carbonium ion) and a
bromide ion.

http://www.chemguide.co.uk/mechanisms/nucsub/cyanidett.html (5 of 7)30/12/2004 11:09:47

Explaining nucleophilic substitution between halogenoalkanes and cyanide ions

This reaction is possible because tertiary carbocations are relatively


stable compared with secondary or primary ones. Even so, the reaction
is slow.

Note: Not sure about the stability of carbocations


(carbonium ions)? Follow this link if you need to find out.

Once the carbocation is formed, however, it would react immediately it


came into contact with a CN- ion. The lone pair on the nucleophile is
strongly attracted towards the positive carbon, and moves towards it to
create a new bond.

How fast the reaction happens is going to be governed by how fast the
halogenoalkane ionises - because that's a slow process. Because this
initial slow step only involves one species, the mechanism is described
as SN1 - substitution, nucleophilic, one species taking part in the initial
slow step.
The SN1 mechanism in secondary halogenoalkanes
Secondary halogenoalkanes (like 2-bromopropane) can use either the
SN1 or the SN2 mechanism. The back of the molecule is rather more
cluttered than in a primary halogenoalkane, but there is still room for the
lone pair on the nucleophile to approach and form a bond. We've already
dealt with that reaction.
It is also possible to get some slight ionisation of the halogenoalkane to
give an SN1 mechanism, but this reaction is much less successful than
with tertiary halogenoalkanes, because the secondary carbocation
formed isn't as stable as a tertiary one.

http://www.chemguide.co.uk/mechanisms/nucsub/cyanidett.html (6 of 7)30/12/2004 11:09:47

Explaining nucleophilic substitution between halogenoalkanes and cyanide ions

Once the carbocation has been formed, it will react immediately with a
cyanide ion. The lone pair on the cyanide ion is strongly attracted to the
positive carbon, moves towards it, and forms a bond.

Where would you like to go now?


To menu of nucleophilic substitution reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000 (modified 2004)

http://www.chemguide.co.uk/mechanisms/nucsub/cyanidett.html (7 of 7)30/12/2004 11:09:47

nucleophilic substitution - halogenoalkanes and ammonia

THE NUCLEOPHILIC SUBSTITUTION REACTIONS


BETWEEN HALOGENOALKANES AND AMMONIA

This page gives you the facts and simple, uncluttered mechanisms for the
nucleophilic substitution reactions between halogenoalkanes and
ammonia to produce primary amines. If you want the mechanisms
explained to you in detail, there is a link at the bottom of the page. If you
are interested in further substitution reactions, you will also find a link to a
separate page dealing with these.

The reaction of primary halogenoalkanes with ammonia


Important! If you aren't sure about the difference between
primary, secondary and tertiary halogenoalkanes, it is
essential that you follow this link before you go on.
Use the BACK button on your browser to return to this page.

The facts
The halogenoalkane is heated with a concentrated solution of ammonia in
ethanol. The reaction is carried out in a sealed tube. You couldn't heat this
mixture under reflux, because the ammonia would simply escape up the
condenser as a gas.
We'll talk about the reaction using 1-bromoethane as a typical primary
halogenoalkane.
The reaction happens in two stages. In the first stage, a salt is formed - in
this case, ethylammonium bromide. This is just like ammonium bromide,
except that one of the hydrogens in the ammonium ion is replaced by an
ethyl group.

http://www.chemguide.co.uk/mechanisms/nucsub/ammonia.html (1 of 5)30/12/2004 11:10:08

nucleophilic substitution - halogenoalkanes and ammonia

There is then the possibility of a reversible reaction between this salt and
excess ammonia in the mixture.

The ammonia removes a hydrogen ion from the ethylammonium ion to


leave a primary amine - ethylamine.
The more ammonia there is in the mixture, the more the forward reaction
is favoured.

Note: You will find considerable disagreement in textbooks


and other sources about the exact nature of the products in
this reaction. Some of the information you'll come across is
simply wrong!
You can read the arguments about the products of this
reaction by following this link.

The mechanism
The mechanism involves two steps. The first is a simple nucleophilic
substitution reaction:

Because the mechanism involves collision between two species in this


slow step of the reaction, it is known as an SN2 reaction.

http://www.chemguide.co.uk/mechanisms/nucsub/ammonia.html (2 of 5)30/12/2004 11:10:08

nucleophilic substitution - halogenoalkanes and ammonia

Note: Unless your syllabus specifically mentions SN2 by


name, you can just call it nucleophilic substitution.

In the second step of the reaction an ammonia molecule may remove one
of the hydrogens on the -NH3+. An ammonium ion is formed, together with
a primary amine - in this case, ethylamine.

This reaction is, however, reversible. Your product will therefore contain a
mixture of ethylammonium ions, ammonia, ethylamine and ammonium
ions. Your major product will only be ethylamine if the ammonia is present
in very large excess.
Unfortunately the reaction doesn't stop here. Ethylamine is a good
nucleophile, and goes on to attack unused bromoethane. This gets so
complicated that it is dealt with on a separate page. You will find a link at
the bottom of this page.

The reaction of tertiary halogenoalkanes with ammonia


The facts
The facts of the reactions are exactly the same as with primary
halogenoalkanes. The halogenoalkane is heated in a sealed tube with a
solution of ammonia in ethanol.
For example:

Followed by:
http://www.chemguide.co.uk/mechanisms/nucsub/ammonia.html (3 of 5)30/12/2004 11:10:08

nucleophilic substitution - halogenoalkanes and ammonia

The mechanism
This mechanism involves an initial ionisation of the halogenoalkane:

followed by a very rapid attack by the ammonia on the carbocation


(carbonium ion) formed:

This is again an example ofnucleophilic substitution.


This time the slow step of the reaction only involves one species - the
halogenoalkane. It is known as an SN1 reaction.
There is a second stage exactly as with primary halogenoalkanes. An
ammonia molecule removes a hydrogen ion from the -NH3+ group in a
reversible reaction. An ammonium ion is formed, together with an amine.

The reaction of secondary halogenoalkanes with ammonia


No current A' level syllabus is likely to ask you about this. In the extremely
unlikely event that you will ever need it, secondary halogenoalkanes use

http://www.chemguide.co.uk/mechanisms/nucsub/ammonia.html (4 of 5)30/12/2004 11:10:08

nucleophilic substitution - halogenoalkanes and ammonia

both an SN2 mechanism and an SN1.


Make sure you understand what happens with primary and tertiary
halogenoalkanes, and then adapt it for secondary ones should ever need
to.

Where would you like to go now?


Help! Talk me through these mechanisms . . .
To look at further substitution in these reactions . . .
To menu of nucleophilic substitution reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/nucsub/ammonia.html (5 of 5)30/12/2004 11:10:08

products of the ammonia - bromoethane reaction

THE PRODUCTS OF THE REACTION BETWEEN


AMMONIA AND BROMOETHANE

This page looks at the problems in writing an equation for the reaction
between ammonia and a halogenoalkane like bromoethane. Although the
discussion involves a primary halogenoalkane, it would apply just as well
to any other kind - secondary or tertiary.

Various incorrect solutions


The problem comes in trying to write a single overall equation for the
reaction. Unfortunately it can't be done without throwing up all sorts of
inconsistencies. If you refer to textbooks or to A' level mark schemes, you
will find three different attempts, all of them slightly unsatisfactory - or
worse!
The first possibility
This comes directly out of the mechanism. The problem is that it produces
the salt of the amine, and not the primary amine itself. Some primary
amine has to be formed to account for further substitution reactions
leading to secondary (etc) amines.
For example, with bromoethane and ammonia, ethylammonium bromide is
formed.

Ethylammonium bromide is a salt of a primary amine and the acid, HBr. A


primary amine has the formula R-NH2. It is primary in the sense that there
is only one alkyl group attached to the nitrogen atom. Primary amines are
weak bases very similar to ammonia and so form salts with acids. For
example, ammonia reacts with HBr to give ammonium bromide, NH4+ Br-.
The salt in the equation above is the one formed from ethylamine,
CH3CH2NH2, and HBr.
http://www.chemguide.co.uk/mechanisms/nucsub/ammoniaeq.html (1 of 4)30/12/2004 11:10:14

products of the ammonia - bromoethane reaction

The second possibility


If a very large excess of ammonia is used, you could get these reactions:

A second ammonia molecule removes a hydrogen ion from the


ethylammonium bromide which is formed to start with, so that you end up
with ethylamine and ammonium bromide. The overall equation, which is
what is normally quoted, is:

The problem with this is that the ethylamine is a stronger base than
ammonia, and so will tend to hang on to the hydrogen ion. It won't easily
give it up to the weaker base, ammonia.
The best you could hope for is an equilibrium mixture containing some of
everything - ammonia, ethylamine and the two salts. You would only get
the above equation as the major reaction if you had a huge excess of
ammonia.
The third possibility

Unless the reaction is done at extremely high temperatures, and never


allowed to cool, this is quite simply wrong! You cannot produce a mixture
containing a base and an acid. They will react together to form a salt.
Surprisingly, some otherwise very reliable sources quote this piece of
chemical nonsense!

Help! What do I need to learn?

http://www.chemguide.co.uk/mechanisms/nucsub/ammoniaeq.html (2 of 4)30/12/2004 11:10:14

products of the ammonia - bromoethane reaction

There is no simple answer to this - it depends entirely on what line your


examiners are currently taking. You need to look at recent mark schemes,
so that you can see what they are accepting. You could also look at any
support material your Exam Board publishes. Once you've found out what
they want, use that equation and forget everything else.

Important! If you haven't got a syllabus and recent exam


papers and mark schemes, it is essential that you get hold of
them. Follow this link to find out how to do that.
Use the BACK button on your browser to return to this page.

The solution used on this site


The best solution - and this is the line you'll find taken in the other pages
on this topic - is a slight modification of the second possibility above.
You simply have to abandon any attempt to simplify this into one equation,
and instead write two equations:

The first equation shows the formation of the salt exactly as before. The
second equation emphasises that the next step is reversible. These
equations are entirely consistent with the mechanism. It is only when you
try to combine them into one equation that the problems start.
If this isn't in line with what your examiners want, learn their version - even
if it's wrong! This only affects the overall equation(s) for the reaction. It
doesn't affect in any way what you write down for the mechanisms.

Where would you like to go now?

http://www.chemguide.co.uk/mechanisms/nucsub/ammoniaeq.html (3 of 4)30/12/2004 11:10:14

products of the ammonia - bromoethane reaction

Return to main page on ammonia and halogenoalkanes . . .


To menu of nucleophilic substitution reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/nucsub/ammoniaeq.html (4 of 4)30/12/2004 11:10:14

Explaining nucleophilic substitution between halogenoalkanes and ammonia

EXPLAINING THE NUCLEOPHILIC


SUBSTITUTION REACTIONS BETWEEN
HALOGENOALKANES AND AMMONIA

This page guides you through the nucleophilic substitution mechanisms


for the reactions between halogenoalkanes and ammonia to produce
primary amines. It deals only with primary and tertiary halogenoalkanes.
No current A' level syllabus is likely to ask about the reaction between
secondary halogenoalkanes and ammonia. (It's not difficult - it's just not
there!)
If you are interested in further substitution reactions, at the bottom of this
page you will find a link to a separate page dealing with these.

Important! It would help if you first read the page What is


nucleophilic substitution? before you go on.
You must also be clear about the differences between
primary, secondary and tertiary halogenoalkanes.

The reaction between primary halogenoalkanes and


ammonia - the SN2 mechanism
Ammonia as a nucleophile
A nucleophile is a species (an ion or a molecule) which is strongly
attracted to a region of positive charge in something else.
Nucleophiles are either fully negative ions, or else have a strongly charge somewhere on a molecule.

http://www.chemguide.co.uk/mechanisms/nucsub/ammoniatt.html (1 of 7)30/12/2004 11:10:18

Explaining nucleophilic substitution between halogenoalkanes and ammonia

Ammonia obviously doesn't carry a negative


charge. However, nitrogen is more electronegative
than hydrogen, and so the nitrogen atom carries
some degree of negative charge. It also has an
active lone pair of electrons. The attack on the +
carbon atom in the halogenoalkane is by this lone pair on the nitrogen
atom.

Note: If you aren't sure about electronegativity follow this link


before you read on.
Use the BACK button on your browser to return to this page.

The nucleophilic substitution reaction - an SN2 reaction


We'll talk this reaction through with bromoethane as a
typical primary halogenoalkane. The bromoethane has
a polar bond between the carbon and the bromine.
The lone pair on the nitrogen will be strongly attracted
to the + carbon, and will move towards it, beginning to form a bond with
it. The approaching lone pair will repel the electrons in the carbonbromine bond closer and closer to the bromine.
The movement goes on until the ammonia is firmly attached to the
carbon, and the bromine has been expelled as a Br- ion.

Notice that the nitrogen in the product ion carries a positive charge
(highlighted in red to draw attention to it). That charge has to be there for
http://www.chemguide.co.uk/mechanisms/nucsub/ammoniatt.html (2 of 7)30/12/2004 11:10:18

Explaining nucleophilic substitution between halogenoalkanes and ammonia

two reasons:

On the left hand side of the equation you start with two overall
neutral molecules. Assuming you forgot about the positive charge,
you would end up with a neutral species and a negative ion on the
right. Charges must balance in equations, so something is wrong.
The nitrogen looks wrong! The nitrogen atom is joined to 4 things
rather than its usual 3. Nitrogen can only join to 4 things if it carries
a positive charge.

Note: Nitrogen with a positive charge has the same


arrangement of electrons as a carbon atom - which normally
forms 4 bonds.

Technically, this is known as an SN2 reaction. S stands for substitution, N


for nucleophilic, and the 2 is because the initial stage of the reaction
involves two species - the bromoethane and the ammonia. If your
syllabus doesn't refer to SN2 reactions by name, you can just call it
nucleophilic substitution.
The product of this reaction is a salt called ethylammonium bromide.
Ethylammonium bromide is a salt of a primary amine and the acid, HBr.
A primary amine has the formula R-NH2. It is primary in the sense that
there is only one alkyl group attached to the nitrogen atom. Primary
amines are weak bases very similar to ammonia and so form salts with
acids. For example, ammonia reacts with HBr to give ammonium
bromide, NH4+ Br-. The salt in the equation above is the one formed from
ethylamine, CH3CH2NH2, and HBr.
At this point you need to know whether your examiners are happy for you
to stop there, or want you to go a stage further to form the free primary
amine.

http://www.chemguide.co.uk/mechanisms/nucsub/ammoniatt.html (3 of 7)30/12/2004 11:10:18

Explaining nucleophilic substitution between halogenoalkanes and ammonia

Note: You need to check the mark schemes for recent exam
papers, or any support material published by your Exam
Board. You won't be able to tell this by looking at your
syllabus, or at the exam papers themselves.
You can find out how to get hold of this material by visiting
your Exam Board's web site. You can find a link to this on the
syllabuses page.

If they want the second step, an ammonia molecule now removes a


hydrogen ion from the nitrogen.

The organic product is ethylamine, CH3CH2NH2.


Notice that this change is reversible. The curly arrows for the reverse
change haven't been shown to avoid confusion, but you can easily
picture the lone pair on the nitrogen reclaiming its hydrogen from the NH4
+

ion. What you will end up with is a mixture of all four of the species in
this last equation - together, of course, with the bromide ions formed in
the first stage. The higher the proportion of ammonia in the original
reaction mixture, the greater the chance of the free ethylamine being
formed.

The reaction between tertiary halogenoalkanes and


ammonia - the SN1 mechanism
We'll talk this mechanism through with a simple
tertiary halogenoalkane like the one on the right (2bromo-2-methylpropane).

http://www.chemguide.co.uk/mechanisms/nucsub/ammoniatt.html (4 of 7)30/12/2004 11:10:18

Explaining nucleophilic substitution between halogenoalkanes and ammonia

Why do tertiary halogenoalkanes need a different mechanism?


When a nucleophile attacks a primary halogenoalkane, it approaches the
+ carbon atom from the side away from the halogen atom. Any other
approach is prevented by the halogen atom, which is both bulky and
slightly negatively charged. The charge repels the incoming nucleophile.
With a tertiary halogenoalkane, this approach from the back is
impossible. The back of the molecule is completely cluttered with CH3
groups.
The SN1 mechanism
In the first stage, a small proportion of the halogenoalkane ionises to give
a carbocation (carbonium ion) and a bromide ion.

This reaction is possible because tertiary carbocations are relatively


stable compared with secondary or primary ones. Even so, the reaction
is slow.

Note: Not sure about the stability of carbocations (carbonium


ions)? Follow this link if you need to find out.

http://www.chemguide.co.uk/mechanisms/nucsub/ammoniatt.html (5 of 7)30/12/2004 11:10:18

Explaining nucleophilic substitution between halogenoalkanes and ammonia

Once the carbocation is formed, however, it would react immediately it


came into contact with an ammonia molecule. The lone pair on the
nitrogen is strongly attracted towards the positive carbon, and moves
towards it to create a new bond.

How fast the reaction happens overall is going to be governed by how


fast the halogenoalkane ionises - because that's a slow process.
Because this initial slow step only involves one species, the mechanism
is described as SN1 - substitution, nucleophilic, one species taking part in
the initial slow step.
A salt is formed, and again you need to know whether your examiners
want you to go on beyond this to the formation of the free amine. If they
do, one of the hydrogens attached to the nitrogen is removed by another
ammonia molecule.

Where would you like to go now?


To look at further substitution in these reactions . . .
To menu of nucleophilic substitution reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

http://www.chemguide.co.uk/mechanisms/nucsub/ammoniatt.html (6 of 7)30/12/2004 11:10:18

Explaining nucleophilic substitution between halogenoalkanes and ammonia

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/nucsub/ammoniatt.html (7 of 7)30/12/2004 11:10:18

multiple nucleophilic substitution - halogenoalkanes and ammonia

MULTIPLE NUCLEOPHILIC SUBSTITUTION IN


THE REACTION BETWEEN
HALOGENOALKANES AND AMMONIA

This page looks at further substitution in the nucleophilic substitution


reaction between halogenoalkanes and ammonia following the formation
of the primary amine. You could also think of this as a description of how
primary, secondary and tertiary amines act as nucleophiles in a
sequence of reactions with a halogenoalkane.
Only primary halogenoalkanes are considered on this page. One current
A' level syllabus (AQA) is likely to ask about these reactions, and that
only asks about primary halogenoalkanes in this context.

Warning! Don't even think about reading this page unless


you are confident about the nucleophilic substitution reaction
involving a primary halogenoalkane and ammonia giving a
primary amine. You need only worry about the SN2 reaction.
The reactions on the current page follow on from those
described on that page. They are no harder, but they do look
more difficult. They are certainly much more tedious! It is
important that you understand what is explained on this page.
Life is too valuable to waste it learning each of the reactions
described here. If you understand what's happening, you can
work out the equations if you need to.

http://www.chemguide.co.uk/mechanisms/nucsub/amines.html (1 of 7)30/12/2004 11:10:22

multiple nucleophilic substitution - halogenoalkanes and ammonia

Why do you get multiple substitution?


The initial substitution - a reminder
The lone pair on the nitrogen atom in an ammonia molecule is attracted
towards the + carbon in the halogenoalkane - in this example,
bromoethane. It forms a bond with it - in the process expelling the
bromine as a bromide ion.

A salt is formed - ethylammonium bromide.


There is then the possibility of another ammonia molecule removing a
hydrogen from the positive ion to give a primary amine - ethylamine. A
primary amine has the general formula RNH2. This reaction is reversible,
and you will only get significant amounts of the free amine if you use a
large excess of ammonia.

The product as a nucleophile


At the end of the initial substitution there will be a certain amount of free
primary amine formed - the CH3CH2NH2 in the example above. There
may not be very much of it, but there will be some. If that small amount
reacts with something else, the reverse reaction can't happen any longer.
The free amine will continue to be produced, and what was originally a
reversible reaction becomes a one-way reaction - provided the free
amine is removed by some other reaction as soon as it is formed.
Now compare the structures of ammonia and ethylamine:

http://www.chemguide.co.uk/mechanisms/nucsub/amines.html (2 of 7)30/12/2004 11:10:22

multiple nucleophilic substitution - halogenoalkanes and ammonia

You can think of the primary amine as a slightly modified ammonia


molecule. It has a lone pair on the nitrogen atom and an even bigger charge than in ammonia.
That means that the primary amine is going to be a better nucleophile
than ammonia is. You can therefore get a reaction between it and a
molecule of the halogenoalkane.

Note: If you are interested in why the nitrogen in ethylamine


has a greater degree of negative charge than the one in
ammonia, you could follow this link and read about organic
bases.

Making a secondary amine from a primary amine


What is a secondary amine?
A secondary amine has the general formula R2NH. It is like an ammonia
molecule (NH3) in which two of the hydrogens have been replaced by
alkyl groups.
The mechanism

The lone pair on the nitrogen in the primary amine attacks the + carbon
exactly the same as the ammonia did. Bromine is lost as a bromide ion,
http://www.chemguide.co.uk/mechanisms/nucsub/amines.html (3 of 7)30/12/2004 11:10:22

multiple nucleophilic substitution - halogenoalkanes and ammonia

and the immediate product is a salt called diethylammonium bromide (CH3CH2)2NH2+ Br-. This is essentially ammonium bromide in which two
of the hydrogens attached to the nitrogen have been replaced by ethyl
groups.
This then reacts with ammonia in a reversible reaction, exactly as we've
seen before:

The organic product is a secondary amine - diethylamine. It is secondary


because there are two alkyl groups attached to the nitrogen atom.

Making a tertiary amine from the secondary amine


What is a tertiary amine?
A tertiary amine has the general formula R3N. It is like an ammonia
molecule in which all three of the hydrogens have been replaced by alkyl
groups.
The mechanism
The secondary amine still has an active lone pair of electrons on the
nitrogen atom. That, in turn, can attack bromoethane if it happens to
collide with it. The two step sequence is exactly as before:

http://www.chemguide.co.uk/mechanisms/nucsub/amines.html (4 of 7)30/12/2004 11:10:22

multiple nucleophilic substitution - halogenoalkanes and ammonia

In this first step, the bromine is again displaced as a bromide ion and you
get a salt formed called triethylammonium bromide.

Note: It has been necessary to rearrange the formula of the


secondary amine in order to keep the appearance of the
mechanism the same as the previous ones. Look carefully at
it to be sure there isn't anything nasty going on!

An ammonia molecule can then remove the hydrogen from the nitrogen
in the reversible reaction:

The organic product of this reaction is the tertiary amine, triethylamine. It


is tertiary because of the three alkyl groups attached to the nitrogen.

The final stage - making a quaternary ammonium salt


What is a quaternary ammonium salt?
A quaternary ammonium salt is an ammonium salt (for example, NH4+
Br-) in which all the hydrogens have been replaced by an alkyl group - for
example, (CH3CH2)4N+ Br-.
http://www.chemguide.co.uk/mechanisms/nucsub/amines.html (5 of 7)30/12/2004 11:10:22

multiple nucleophilic substitution - halogenoalkanes and ammonia

The mechanism
The tertiary amine still has an active lone pair on the nitrogen and, once
again, that can attack the + carbon in the bromoethane.

But this time there is nowhere else for the reaction to go. There is no
longer a hydrogen atom on the nitrogen that an ammonia molecule could
remove, and so the reaction finally comes to an end.
The product is a salt called tetraethylammonium bromide, (CH3CH2)4N+
Br-.

If you react bromoethane and ammonia, what do you


actually get?
You will always get a mixture of all the above products unless you use a
very large excess of bromoethane so that there is enough of it to carry
the sequence all the way through to the quaternary ammonium salt.
You could favour formation of the primary amine (or its salt) by using a
very large excess of ammonia. That way there is always a greater
chance of a bromoethane molecule being hit by an ammonia molecule
than any other possible nucleophile. Even so, there will always, by
chance, be some collisions leading to the follow-on products.

Where would you like to go now?


To menu of nucleophilic substitution reactions. . .

http://www.chemguide.co.uk/mechanisms/nucsub/amines.html (6 of 7)30/12/2004 11:10:22

multiple nucleophilic substitution - halogenoalkanes and ammonia

To menu of other types of mechanism. . .


To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/nucsub/amines.html (7 of 7)30/12/2004 11:10:22

Explaining the strength of organic bases

ORGANIC BASES

This page explains why simple organic bases are basic and looks at the
factors which affect their relative strengths. For A'level purposes, all the
bases we are concerned with are primary amines - compounds in which
one of the hydrogens in an ammonia molecule, NH3, is replaced either
by an alkyl group or a benzene ring.

Why are primary amines basic?


Ammonia as a weak base
All of the compounds we are concerned with are derived from ammonia
and so we'll start by looking at the reason for its basic properties.
For the purposes of this topic, we are going to take the definition of a
base as "a substance which combines with hydrogen ions (protons)". We
are going to get a measure of this by looking at how easily the bases
take hydrogen ions from water molecules when they are in solution in
water.
Ammonia in solution sets up this equilibrium:

An ammonium ion is formed together with hydroxide ions. Because the


ammonia is only a weak base, it doesn't hang on to the extra hydrogen
ion very effectively and so the reaction is reversible. At any one time,
about 99% of the ammonia is present as unreacted molecules. The
position of equilibrium lies well to the left.
The ammonia reacts as a base because of the active lone pair on the
nitrogen. Nitrogen is more electronegative than hydrogen and so attracts
the bonding electrons in the ammonia molecule towards itself. That
means that in addition to the lone pair, there is a build-up of negative
charge around the nitrogen atom. That combination of extra negativity
http://www.chemguide.co.uk/basicorg/acidbase/bases.html (1 of 7)30/12/2004 11:10:28

Explaining the strength of organic bases

and active lone pair attracts the new hydrogen from the water.

Comparing the strengths of weak bases


The strengths of weak bases are measured on the pKb scale. The
smaller the number on this scale, the stronger the base is.
Three of the compounds we shall be looking at, together with their pKb
values are:

Remember - the smaller the number the stronger the base. Comparing
the other two to ammonia, you will see that methylamine is a stronger
base, whereas phenylamine is very much weaker.
Methylamine is typical of aliphatic primary amines - where the -NH2
http://www.chemguide.co.uk/basicorg/acidbase/bases.html (2 of 7)30/12/2004 11:10:28

Explaining the strength of organic bases

group is attached to a carbon chain. All aliphatic primary amines are


stronger bases than ammonia.
Phenylamine is typical of aromatic primary amines - where the -NH2
group is attached directly to a benzene ring. These are very much
weaker bases than ammonia.

Explaining the differences in base strengths


The factors to consider
Two of the factors which influence the strength of a base are:

the ease with which the lone pair picks up a hydrogen ion,

the stability of the ions being formed.

Why are aliphatic primary amines stronger bases than ammonia?


Methylamine
Methylamine has the structure:

The only difference between this and ammonia is the presence of the
CH3 group in the methylamine. But that's important! Alkyl groups have a
tendency to "push" electrons away from themselves. That means that
there will be a small amount of extra negative charge built up on the
nitrogen atom. That extra negativity around the nitrogen makes the lone
pair even more attractive towards hydrogen ions.

http://www.chemguide.co.uk/basicorg/acidbase/bases.html (3 of 7)30/12/2004 11:10:28

Explaining the strength of organic bases

Making the nitrogen more negative helps the lone pair to pick up a
hydrogen ion.
What about the effect on the positive methylammonium ion formed? Is
this more stable than a simple ammonium ion?
Compare the methylammonium ion with an ammonium ion:

In the methylammonium ion, the positive charge is spread around the ion
by the "electron-pushing" effect of the methyl group. The more you can
spread charge around, the more stable an ion becomes. In the
ammonium ion there isn't any way of spreading the charge.
To summarise:

The nitrogen is more negative in methylamine than in ammonia,


and so it picks up a hydrogen ion more readily.
The ion formed from methylamine is more stable than the one
formed from ammonia, and so is less likely to shed the hydrogen
ion again.

http://www.chemguide.co.uk/basicorg/acidbase/bases.html (4 of 7)30/12/2004 11:10:28

Explaining the strength of organic bases

Taken together, these mean that methylamine is a stronger base than


ammonia.

Note: This is a bit of a simplification for A' level purposes. As


bases get more complex, another factor concerning the
stability of the ions formed becomes important. That
concerns the way they interact with water molecules in the
solution. You don't need to worry about that at this level.

The other aliphatic primary amines


The other alkyl groups have "electron-pushing" effects very similar to the
methyl group, and so the strengths of the other aliphatic primary amines
are very similar to methylamine.

Note: If you want more information about the inductive effect


of alkyl groups, you could read about carbocations
(carbonium ions) in the mechanism section of this site.

For example:
pKb
CH3NH2

3.36

CH3CH2NH2

3.27

CH3CH2CH2NH2

3.16

CH3CH2CH2CH2NH2

3.39

Why are aromatic primary amines much weaker bases than


ammonia?
An aromatic primary amine is one in which the -NH2 group is attached
http://www.chemguide.co.uk/basicorg/acidbase/bases.html (5 of 7)30/12/2004 11:10:28

Explaining the strength of organic bases

directly to a benzene ring. The only one you are likely to come across is
phenylamine.
Phenylamine has the structure:

The lone pair on the nitrogen touches the delocalised ring electrons . . .

. . . and becomes delocalised with them:

That means that the lone pair is no longer fully available to combine with
hydrogen ions. The nitrogen is still the most electronegative atom in the
molecule, and so the delocalised electrons will be attracted towards it,
but the intensity of charge around the nitrogen is nothing like what it is in,
say, an ammonia molecule.

http://www.chemguide.co.uk/basicorg/acidbase/bases.html (6 of 7)30/12/2004 11:10:28

Explaining the strength of organic bases

The other problem is that if the lone pair is used to join to a hydrogen ion,
it is no longer available to contribute to the delocalisation. That means
that the delocalisation would have to be disrupted if the phenylamine acts
as a base. Delocalisation makes molecules more stable, and so
disrupting the delocalisation costs energy and won't happen easily.
Taken together - the lack of intense charge around the nitrogen, and the
need to break some delocalisation - this means that phenylamine is a
very weak base indeed.

Where would you like to go now?


To the acids and bases menu. . .
To menu of basic organic chemistry. . .
To Main Menu . . .

You might also be interested in:


The properties and reactions of:
aliphatic amines . . .
phenylamine (aniline) . . .

Jim Clark 2000

http://www.chemguide.co.uk/basicorg/acidbase/bases.html (7 of 7)30/12/2004 11:10:28

Organic acids and bases menu

Understanding Chemistry

ORGANIC ACIDS AND BASES MENU

Organic acids . . .
Explains why organic acids are acidic, and what affects their
strengths.
Organic bases . . .
Explains why primary amines are basic, and what affects their
strengths.

Go to menu of basic organic chemistry. . .


Go to Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/basicorg/acidmenu.html30/12/2004 11:10:29

Amines Menu

Understanding Chemistry

ALIPHATIC AMINES MENU

This only covers amines where the functional group is not attached
directly to a benzene ring. There is a link towards the bottom of the page
to a separate section about phenylamine (aniline) if you are interested.

Background . . .
An introduction to amines including the various types of amine
(primary, secondary and tertiary) and their physical properties.
Preparation of amines . . .
Their preparation from halogenoalkanes (haloalkanes or alkyl
halides) and from nitriles.
Amines as bases . . .
The importance of the lone pair on the nitrogen in the reactions of
amines as bases. Their reactions with acids, water and copper(II)
ions.
Amines as nucleophiles . . .
The importance of the lone pair on the nitrogen in the reactions of
amines as nucleophiles. Their reactions with halogenoalkanes
(haloalkanes or alkyl halides), with acyl chlorides (acid chlorides)
and with acid anhydrides.
The reaction of amines with nitrous acid (nitric(III) acid) . . .
http://www.chemguide.co.uk/organicprops/aminemenu.html (1 of 2)30/12/2004 11:10:30

Amines Menu

The use of nitrous acid in testing for amines.

Information about phenylamine (aniline) . . .


The preparation and physical and chemical properties of the
aromatic amine, phenylamine (aniline). This section also includes
the preparation and properties of diazonium salts.

Go to menu of other organic compounds . . .


Go to Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/aminemenu.html (2 of 2)30/12/2004 11:10:30

an introduction to amines

INTRODUCING AMINES

This page explains what amines are, and what the difference is between
primary, secondary and tertiary amines. It looks in some detail at their
simple physical properties such as solubility and boiling points. Details of
the chemical reactions of amines are described on separate pages.

Note: This page only deals with amines where the functional
group is not attached directly to a benzene ring. Aromatic
amines such as phenylamine (aniline) are sufficiently
different that they are covered in a separate section. Follow
this link if you are mainly interested in phenylamine.

What are amines?


The easiest way to think of amines is as near relatives of ammonia, NH3.
In amines, the hydrogen atoms in the ammonia have been replaced one
at a time by hydrocarbon groups. On this page, we are only looking at
cases where the hydrocarbon groups are simple alkyl groups.

The different kinds of amines


Amines fall into different classes depending on how many of the
hydrogen atoms are replaced.
Primary amines
In primary amines, only one of the hydrogen atoms in the ammonia
molecule has been replaced. That means that the formula of the primary
amine will be RNH2 where "R" is an alkyl group.
Examples include:
http://www.chemguide.co.uk/organicprops/amines/background.html (1 of 7)30/12/2004 11:10:36

an introduction to amines

Naming amines can be quite confusing because there are so many


variations on the names. For example, the simplest amine, CH3NH2, can
be called methylamine, methanamine or aminomethane.

The commonest name at this level is methylamine and, similarly, the


second compound drawn above is usually called ethylamine.
Where there might be confusion about where the -NH2 group is attached
to a chain, the simplest way of naming the compound is to use the
"amino" form.
For example:

Secondary amines
In a secondary amine, two of the hydrogens in an ammonia molecule
have been replaced by hydrocarbon groups. At this level, you are only
likely to come across simple ones where both of the hydrocarbon groups
are alkyl groups and both are the same.
For example:

http://www.chemguide.co.uk/organicprops/amines/background.html (2 of 7)30/12/2004 11:10:36

an introduction to amines

There are other variants on the names, but this is the commonest and
simplest way of naming these small secondary amines.

Tertiary amines
In a tertiary amine, all of the hydrogens in an ammonia molecule have
been replaced by hydrocarbon groups. Again, you are only likely to come
across simple ones where all three of the hydrocarbon groups are alkyl
groups and all three are the same.
The naming is similar to secondary amines. For example:

Physical properties of amines


Boiling points
The table shows the boiling points of some simple amines.

type

formula

boiling point (C)

primary

CH3NH2

-6.3

primary

CH3CH2NH2

16.6

http://www.chemguide.co.uk/organicprops/amines/background.html (3 of 7)30/12/2004 11:10:36

an introduction to amines

primary

CH3CH2CH2NH2

48.6

secondary

(CH3)2NH

7.4

tertiary

(CH3)3N

3.5

We will need to look at this with some care to sort out the patterns and
reasons. Concentrate first on the primary amines.
Primary amines
It is useful to compare the boiling point of methylamine, CH3NH2, with
that of ethane, CH3CH3.
Both molecules contain the same number of electrons and have, as near
as makes no difference, the same shape. However, the boiling point of
methylamine is -6.3C, whereas ethane's boiling point is much lower at 88.6C.
The reason for the higher boiling points of the primary amines is that they
can form hydrogen bonds with each other as well as van der Waals
dispersion forces and dipole-dipole interactions.

Note: If you aren't happy about intermolecular forces


(including van der Waals dispersion forces and hydrogen
bonds) then you really ought to follow this link before you go
on. The next bit won't make much sense to you if you aren't
familiar with the various sorts of intermolecular forces.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/amines/background.html (4 of 7)30/12/2004 11:10:36

an introduction to amines

Hydrogen bonds can form between the lone pair on the very
electronegative nitrogen atom and the slightly positive hydrogen atom in
another molecule.

The hydrogen bonding isn't as efficient as it is in, say, water, because


there is a shortage of lone pairs. Some slightly positive hydrogen atoms
won't be able to find a lone pair to hydrogen bond with. There are twice
as many suitable hydrogens are there are lone pairs.
The boiling points of the primary amines increase as you increase chain
length because of the greater amount of van der Waals dispersion forces
between the bigger molecules.

Secondary amines
For a fair comparison you would have to compare the boiling point of
dimethylamine with that of ethylamine. They are isomers of each other each contains exactly the same number of the same atoms.
The boiling point of the secondary amine is a little lower than the
corresponding primary amine with the same number of carbon atoms.
Secondary amines still form hydrogen bonds, but having the nitrogen
atom in the middle of the chain rather than at the end makes the
permanent dipole on the molecule slightly less.
The lower boiling point is due to the lower dipole-dipole attractions in the
dimethylamine compared with ethylamine.

http://www.chemguide.co.uk/organicprops/amines/background.html (5 of 7)30/12/2004 11:10:36

an introduction to amines

Tertiary amines
This time to make a fair comparison you would have to compare
trimethylamine with its isomer 1-aminopropane.
If you look back at the table further up the page, you will see that the
trimethylamine has a much lower boiling point (3.5C) than 1aminopropane (48.6C).
In a tertiary amine there aren't any hydrogen atoms attached directly to
the nitrogen. That means that hydrogen bonding between tertiary amine
molecules is impossible. That's why the boiling point is much lower.

Solubility in water
The small amines of all types are very soluble in water. In fact, the ones
that would normally be found as gases at room temperature are normally
sold as solutions in water - in much the same way that ammonia is
usually supplied as ammonia solution.
All of the amines can form hydrogen bonds with water - even the tertiary
ones.
Although the tertiary amines don't have a hydrogen atom attached to the
nitrogen and so can't form hydrogen bonds with themselves, they can
form hydrogen bonds with water molecules just using the lone pair on the
nitrogen.

Solubility falls off as the hydrocarbon chains get longer - noticeably so


after about 6 carbons. The hydrocarbon chains have to force their way

http://www.chemguide.co.uk/organicprops/amines/background.html (6 of 7)30/12/2004 11:10:36

an introduction to amines

between water molecules, breaking hydrogen bonds between water


molecules.
However, they don't replace them by anything as strong, and so the
process of forming a solution becomes less and less energetically
feasible as chain length grows.

Smell
The very small amines like methylamine and ethylamine smell very
similar to ammonia - although if you compared them side by side, the
amine smells are slightly more complex.
As the amines get bigger, they tend to smell more "fishy", or they smell of
decay.
If you are familiar with the smell of hawthorn blossom (and similarly
smelling things like cotoneaster blossom), this is the smell of
trimethylamine - a sweet and rather sickly smell like the early stages of
decaying flesh.

Where would you like to go now?


To the amines menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/amines/background.html (7 of 7)30/12/2004 11:10:36

Phenylamine (aniline) Menu

Understanding Chemistry

PHENYLAMINE (ANILINE) MENU

Background . . .
An introduction to phenylamine (aniline) and its physical properties.
Preparation . . .
Making phenylamine from benzene via nitrobenzene.
Reactions of phenylamine as a primary amine . . .
This covers the reactions that phenylamine has in common with
other primary amines - its reactions as a base, its acylation with
acyl chlorides and acid anhydrides, and its reaction with
halogenoalkanes.
Making diazonium salts . . .
The reaction of phenylamine with nitrous acid to produce
diazonium ions.
Reactions of diazonium salts . . .
Using them to make phenol and iodobenzene, and the coupling
reactions with phenol, naphthalen-2-ol (2-naphthol) and
phenylamine.

Go to menu of other organic compounds . . .


http://www.chemguide.co.uk/organicprops/anilinemenu.html (1 of 2)30/12/2004 11:10:38

Phenylamine (aniline) Menu

Go to Main Menu . . .

You might also be interested in:


properties and reactions of aliphatic amines. . .
Covers the physical and chemical properties of amines where the
amine group is attached to a carbon chain (or just a methyl group)
rather than a benzene ring.

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/anilinemenu.html (2 of 2)30/12/2004 11:10:38

an introduction to phenylamine (aniline)

INTRODUCING PHENYLAMINE

This page looks at the structure and physical properties of phenylamine also known as aniline or aminobenzene. Phenylamine has an -NH2
group attached directly to a benzene ring.

The structure of phenylamine


Phenylamine is a primary amine - a compound in which one of the
hydrogen atoms in an ammonia molecule has been replaced by a
hydrocarbon group.
However, in comparison with simple primary amines like methylamine,
the properties of phenylamine are slightly different. This is because the
lone pair on the nitrogen atom interacts with the delocalised electrons in
the benzene ring.
The simplest way to draw the structure of phenylamine is:

http://www.chemguide.co.uk/organicprops/aniline/background.html (1 of 5)30/12/2004 11:10:45

an introduction to phenylamine (aniline)

Warning! You need to understand about the bonding in


benzene in order to make sense of this next bit.
If you follow this link, you may have to explore several other
pages before you are ready to come back here again. Use
the BACK button (or HISTORY file or GO menu) on your
browser to return to this page.

There is an interaction between the delocalised electrons in the benzene


ring and the lone pair on the nitrogen atom. The lone pair overlaps with
the delocalised ring electron system . . .

. . . giving a structure rather like this:

The donation of the nitrogen's lone pair into the ring system increases
the electron density around the ring. That makes the ring much more
reactive than it is in benzene itself.
http://www.chemguide.co.uk/organicprops/aniline/background.html (2 of 5)30/12/2004 11:10:45

an introduction to phenylamine (aniline)

Note: The ring reactions of phenylamine aren't on any of the


current UK A level chemistry syllabuses, so I haven't followed
this up anywhere on this site.

It also reduces the availability of the lone pair on the nitrogen to take part
in other reactions. In particular, it makes phenylamine much more weakly
basic than primary amines where the -NH2 group isn't attached to a
benzene ring. That will be explored elsewhere in this section. (See the
phenylamine menu - link at the bottom of this page.)

Physical properties
Pure phenylamine is a colourless liquid, but it darkens rapidly on
exposure to light and air. It is normally a brown oily liquid.
Melting and boiling points
It is useful to compare phenylamine's melting and boiling points with
those of methylbenzene (toluene). Both molecules contain a similar
number of electrons and are a very similar shape. That means that the
intermolecular attractions due to van der Waals dispersion forces are
going to be very similar.

Note: If you aren't happy about intermolecular forces


(including van der Waals dispersion forces and hydrogen
bonds) then you really ought to follow this link before you go
on. This section won't make much sense to you if you aren't
familiar with the various sorts of intermolecular forces.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/aniline/background.html (3 of 5)30/12/2004 11:10:45

an introduction to phenylamine (aniline)

melting point (C) boiling point (C)


C6H5NH2

-6.2

184

C6H5CH3

-95.0

111

The reason for the higher values for phenylamine is in part due to
permanent dipole-dipole attractions due to the electronegativity of the
nitrogen - but is mainly due to hydrogen bonding.
Hydrogen bonds can form between a lone pair on a nitrogen on one
molecule and the hydrogen on the -NH2 group of one of its neighbours.

Note: If you are amazingly wide-awake, you might wonder


how the lone pair can form hydrogen bonds if it is delocalised
into the ring electrons. The truth is that the delocalisation isn't
complete. You can think of the lone pair as still being there,
but not as effective as it would otherwise be.

Solubility in water
Phenylamine is slightly soluble in water - about 3.6 g (depending on
where you get the data from!) of phenylamine will dissolve in 100 g of
water at 20C. Mixtures containing more phenylamine than this separate
into two layers, with the phenylamine forming the bottom one.
Phenylamine is somewhat soluble in water because of its ability to form
hydrogen bonds with the water.
However, the benzene rings in the phenylamine break more hydrogen
bonds between water molecules than are reformed between water and
the -NH2 groups. The water molecules also disrupt fairly strong van der
Waals attractions between the phenylamine molecules.
Both of these effects mean that dissolving phenylamine in water isn't very
http://www.chemguide.co.uk/organicprops/aniline/background.html (4 of 5)30/12/2004 11:10:45

an introduction to phenylamine (aniline)

energetically profitable, and so stop the phenylamine from being very


soluble.

Where would you like to go now?


To the phenylamine menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

You might also be interested in:


properties and reactions of aliphatic amines. . .
Covers the physical and chemical properties of amines where the
amine group is attached to a carbon chain (or just a methyl group)
rather than a benzene ring.

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/aniline/background.html (5 of 5)30/12/2004 11:10:45

preparation of amines

MAKING AMINES

This page looks at the preparation of amines from halogenoalkanes (also


known as haloalkanes or alkyl halides) and from nitriles.
It only deals with amines where the functional group is not attached
directly to a benzene ring. Aromatic amines such as phenylamine (aniline)
are usually made differently and are discussed on a separate page.

Note: Follow this link if you are mainly interested in the


preparation of phenylamine.

Making amines from halogenoalkanes


The halogenoalkane is heated with a concentrated solution of ammonia in
ethanol. The reaction is carried out in a sealed tube. You couldn't heat this
mixture under reflux, because the ammonia would simply escape up the
condenser as a gas.
We'll talk about the reaction using 1-bromoethane as a typical
halogenoalkane.
You get a mixture of amines formed together with their salts. The reactions
happen one after another.

Making a primary amine


The reaction happens in two stages. In the first stage, a salt is formed - in
this case, ethylammonium bromide. This is just like ammonium bromide,
except that one of the hydrogens in the ammonium ion is replaced by an
ethyl group.

http://www.chemguide.co.uk/organicprops/amines/preparation.html (1 of 7)30/12/2004 11:10:52

preparation of amines

There is then the possibility of a reversible reaction between this salt and
excess ammonia in the mixture.

The ammonia removes a hydrogen ion from the ethylammonium ion to


leave a primary amine - ethylamine.
The more ammonia there is in the mixture, the more the forward reaction is
favoured.

Note: You will find considerable disagreement in textbooks


and other sources about the exact nature of the products in this
reaction. Some of the information you'll come across is simply
wrong!
You can read the arguments about the products of this reaction
by following this link.
Warning! That page is in the mechanism section of the site.
Return to the current page using the BACK button on your
browser. If you use the links at the bottom of that page, you
could get seriously lost!

http://www.chemguide.co.uk/organicprops/amines/preparation.html (2 of 7)30/12/2004 11:10:52

preparation of amines

Making a secondary amine


The reaction doesn't stop at a primary amine. The ethylamine also reacts
with bromoethane - in the same two stages as before.
In the first stage, you get a salt formed - this time, diethylammonium
bromide. Think of this as ammonium bromide with two hydrogens replaced
by ethyl groups.

There is again the possibility of a reversible reaction between this salt and
excess ammonia in the mixture.

The ammonia removes a hydrogen ion from the diethylammonium ion to


leave a secondary amine - diethylamine. A secondary amine is one which
has two alkyl groups attached to the nitrogen.

Making a tertiary amine


And still it doesn't stop! The diethylamine also reacts with bromoethane - in
the same two stages as before.
In the first stage, you get triethylammonium bromide.

http://www.chemguide.co.uk/organicprops/amines/preparation.html (3 of 7)30/12/2004 11:10:52

preparation of amines

There is again the possibility of a reversible reaction between this salt and
excess ammonia in the mixture.

The ammonia removes a hydrogen ion from the triethylammonium ion to


leave a tertiary amine - triethylamine. A tertiary amine is one which has
three alkyl groups attached to the nitrogen.

Making a quaternary ammonium salt


The final stage! The triethylamine reacts with bromoethane to give
tetraethylammonium bromide - a quaternary ammonium salt (one in which
all four hydrogens have been replaced by alkyl groups).

This time there isn't any hydrogen left on the nitrogen to be removed. The
reaction stops here.

Note: This whole reaction sequence is a complete pain if you


are going to have to learn it. It is much, much easier to work it
out if you need to, provided you understand the mechanisms
for the reactions.
You can explore the mechanisms for the various stages of the
reaction by following this link. This will lead you to several
pages in the mechanism section of this site. If all you want to
do is make some sense of the above reactions, it would
probably pay you to just read the parts of those pages
concerned with primary halogenoalkanes like bromoethane.

http://www.chemguide.co.uk/organicprops/amines/preparation.html (4 of 7)30/12/2004 11:10:52

preparation of amines

What do you actually get if you react bromoethane with ammonia?


Whatever you do, you get a mixture of all of the products (including the
various amines and their salts) shown on this page.
To get mainly the quaternary ammonium salt, you can use a large excess
of bromoethane. If you look at the reactions going on, each one needs
additional bromoethane. If you provide enough, then the chances are that
the reaction will go to completion, given enough time.
On the other hand, if you use a very large excess of ammonia, the
chances are always greatest that a bromoethane molecule will hit an
ammonia molecule rather than one of the amines being formed. That will
help to prevent the formation of secondary (etc) amines - although it won't
stop it entirely.

Making primary amines from nitriles


Nitriles are compounds containing the -CN group, and can be reduced in
various ways. Two possible methods are described here.

Reducing nitriles using LiAlH4


One possible reducing agent is lithium tetrahydridoaluminate(III) - often
just called lithium tetrahydridoaluminate or lithium aluminium hydride.
The nitrile reacts with the lithium tetrahydridoaluminate in solution in
ethoxyethane (diethyl ether, or just "ether") followed by treatment of the
product of that reaction with a dilute acid.
Overall, the carbon-nitrogen triple bond is reduced to give a primary amine.
For example, with ethanenitrile you get ethylamine:

http://www.chemguide.co.uk/organicprops/amines/preparation.html (5 of 7)30/12/2004 11:10:52

preparation of amines

Notice that this is a simplified equation - perfectly acceptable to UK A level


examiners. [H] means "hydrogen from a reducing agent".

Note: If you know about the reduction of aldehydes and


ketones, you may know that they are also reduced by the
similar compound NaBH4.
However, NaBH4 isn't a strong enough reducing agent to
reduce nitriles.

The reduction of nitriles using hydrogen and a metal catalyst


The carbon-nitrogen triple bond in a nitrile can also be reduced by reaction
with hydrogen gas in the presence of a variety of metal catalysts.
Commonly quoted catalysts are palladium, platinum or nickel.
The reaction will take place at a raised temperature and pressure. It is
impossible to give exact details because it will vary from catalyst to
catalyst.
For example, ethanenitrile can be reduced to ethylamine by reaction with
hydrogen in the presence of a palladium catalyst.

Note: Notice that this time the hydrogen is written normally as


H2. This is a proper equation involving hydrogen gas - not a
simplification.

http://www.chemguide.co.uk/organicprops/amines/preparation.html (6 of 7)30/12/2004 11:10:52

preparation of amines

Where would you like to go now?


To the amines menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/amines/preparation.html (7 of 7)30/12/2004 11:10:52

the preparation of phenylamine (aniline)

MAKING PHENYLAMINE

This page looks in outline at the preparation of phenylamine (also known


as aniline or aminobenzene) starting from benzene. The benzene is first
converted to nitrobenzene which is in turn reduced to phenylamine.

Benzene to nitrobenzene
Benzene is nitrated by replacing one of the hydrogen atoms on the
benzene ring by a nitro group, NO2.
The benzene is treated with a mixture of concentrated nitric acid and
concentrated sulphuric acid at a temperature not exceeding 50C. The
mixture is held at this temperature for about half an hour. Yellow oily
nitrobenzene is formed.

You could write this in a more condensed form as:

The concentrated sulphuric acid is acting as a catalyst and so isn't


written into the equations.
The temperature is kept relatively low to prevent more than one nitro
group being substituted onto the ring.

http://www.chemguide.co.uk/organicprops/aniline/preparation.html (1 of 4)30/12/2004 11:10:58

the preparation of phenylamine (aniline)

Note: Follow this link if you want the mechanism for the
nitration of benzene.
Use the BACK button (or the HISTORY file or GO menu) on
your browser to return to this page later.

Nitrobenzene to phenylamine
The conversion is done in two main stages:
Stage 1: conversion of nitrobenzene into phenylammonium ions
Nitrobenzene is reduced to phenylammonium ions using a mixture of tin
and concentrated hydrochloric acid. The mixture is heated under reflux in
a boiling water bath for about half an hour.
Under the acidic conditions, rather than getting phenylamine directly, you
instead get phenylammonium ions formed. The lone pair on the nitrogen
in the phenylamine picks up a hydrogen ion from the acid.
The electron-half-equation for this reaction is:

The nitrobenzene has been reduced by gaining electrons in the presence


of the acid.
The electrons come from the tin, which forms both tin(II) and tin(IV) ions.

http://www.chemguide.co.uk/organicprops/aniline/preparation.html (2 of 4)30/12/2004 11:10:58

the preparation of phenylamine (aniline)

Note: I have given these as electron-half-equations rather


than attempting full equations in order to try to show what is
happening. Combining them into full equations leads you to
some really scary equations where it is difficult to see what is
going on. The problem is made much worse because the tin
ions formed go on to react with chloride ions from the
hydrochloric acid to form complex ions such as [SnCl6]2-.
If you aren't sure about electron-half-equations you could
follow this link - but it really isn't important for UK A level
purposes to worry too much about this in the present context.
You are unlikely to need much more than the conditions for
the reaction.
Use the BACK button on your browser to return to this page if
you should decide to follow this link.

Stage 2: conversion of the phenylammonium ions into phenylamine


All you need to do is to remove the hydrogen ion from the -NH3+ group.
Sodium hydroxide solution is added to the product of the first stage of the
reaction.

The phenylamine is formed together with a complicated mixture of tin


compounds from reactions between the sodium hydroxide solution and
the complex tin ions formed during the first stage.
The phenylamine is finally separated from this mixture. The separation is
long, tedious and potentially dangerous - involving steam distillation,
solvent extraction and a final distillation.
http://www.chemguide.co.uk/organicprops/aniline/preparation.html (3 of 4)30/12/2004 11:10:58

the preparation of phenylamine (aniline)

Note: The conversion of nitrobenzene into phenylamine is


so time-consuming, complicated, and hazardous at this level
that I'm not going to make any attempt to describe this in
detail. If you want details, refer to any good practical organic
chemistry textbook.

Summary
What you are likely to need for UK A level chemistry purposes can be
summed up by:

You are almost bound to need the mechanism for the nitration reaction
as well.

Where would you like to go now?


To the phenylamine menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/aniline/preparation.html (4 of 4)30/12/2004 11:10:58

Writing ionic equations for redox reactions

WRITING IONIC EQUATIONS FOR REDOX


REACTIONS

This page explains how to work out electron-half-reactions for oxidation


and reduction processes, and then how to combine them to give the
overall ionic equation for a redox reaction. This is an important skill in
inorganic chemistry.
Don't worry if it seems to take you a long time in the early stages. It is a
fairly slow process even with experience. Take your time and practice as
much as you can.

Electron-half-equations
What is an electron-half-equation?
When magnesium reduces hot copper(II) oxide to copper, the ionic
equation for the reaction is:

Note: I am going to leave out state symbols in all the


equations on this page. This topic is awkward enough anyway
without having to worry about state symbols as well as
everything else. Practice getting the equations right, and then
add the state symbols in afterwards if your examiners are likely
to want them.
How do you know whether your examiners will want you to
include them? The best way is to look at their mark schemes.
You should be able to get these from your Exam Board's
website. There are links on the syllabuses page.

http://www.chemguide.co.uk/inorganic/redox/equations.html (1 of 11)30/12/2004 11:11:16

Writing ionic equations for redox reactions

You can split the ionic equation into two parts, and look at it from the point
of view of the magnesium and of the copper(II) ions separately. This
shows clearly that the magnesium has lost two electrons, and the copper
(II) ions have gained them.

These two equations are described as "electron-half-equations" or "halfequations" or "ionic-half-equations" or "half-reactions" - lots of variations all
meaning exactly the same thing!
Any redox reaction is made up of two half-reactions: in one of them
electrons are being lost (an oxidation process) and in the other one those
electrons are being gained (a reduction process).

Note: If you aren't happy about redox reactions in terms of


electron transfer, you MUST read the introductory page on
redox reactions before you go on.

Working out electron-half-equations and using them to build ionic


equations
In the example above, we've got at the electron-half-equations by starting
from the ionic equation and extracting the individual half-reactions from it.
That's doing everything entirely the wrong way round!
In reality, you almost always start from the electron-half-equations and use
them to build the ionic equation.
Example 1: The reaction between chlorine and iron(II) ions
Chlorine gas oxidises iron(II) ions to iron(III) ions. In the process, the
chlorine is reduced to chloride ions.
http://www.chemguide.co.uk/inorganic/redox/equations.html (2 of 11)30/12/2004 11:11:16

Writing ionic equations for redox reactions

You would have to know this, or be told it by an examiner. In building


equations, there is quite a lot that you can work out as you go along, but
you have to have somewhere to start from!
You start by writing down what you know for each of the half-reactions. In
the chlorine case, you know that chlorine (as molecules) turns into chloride
ions:

The first thing to do is to balance the atoms that you have got as far as you
possibly can:

ALWAYS check that you have the existing atoms balanced before you do
anything else. If you forget to do this, everything else that you do
afterwards is a complete waste of time!
Now you have to add things to the half-equation in order to make it
balance completely.
All you are allowed to add are:

electrons

water

hydrogen ions (unless the reaction is being done under alkaline


conditions - in which case, you can add hydroxide ions instead)

In the chlorine case, all that is wrong with the existing equation that we've
produced so far is that the charges don't balance. The left-hand side of the
equation has no charge, but the right-hand side carries 2 negative charges.
That's easily put right by adding two electrons to the left-hand side. The
final version of the half-reaction is:

http://www.chemguide.co.uk/inorganic/redox/equations.html (3 of 11)30/12/2004 11:11:16

Writing ionic equations for redox reactions

Now you repeat this for the iron(II) ions. You know (or are told) that they
are oxidised to iron(III) ions. Write this down:

The atoms balance, but the charges don't. There are 3 positive charges on
the right-hand side, but only 2 on the left.
You need to reduce the number of positive charges on the right-hand side.
That's easily done by adding an electron to that side:

Combining the half-reactions to make the ionic equation for the reaction
What we've got at the moment is this:

It is obvious that the iron reaction will have to happen twice for every
chlorine molecule that reacts. Allow for that, and then add the two halfequations together.

But don't stop there!! Check that everything balances - atoms and charges.
It is very easy to make small mistakes, especially if you are trying to
multiply and add up more complicated equations.
http://www.chemguide.co.uk/inorganic/redox/equations.html (4 of 11)30/12/2004 11:11:16

Writing ionic equations for redox reactions

You will notice that I haven't bothered to include the electrons in the addedup version. If you think about it, there are bound to be the same number on
each side of the final equation, and so they will cancel out. If you aren't
happy with this, write them down and then cross them out afterwards!

Example 2: The reaction between hydrogen peroxide and manganate


(VII) ions
The first example was a simple bit of chemistry which you may well have
come across. The technique works just as well for more complicated (and
perhaps unfamiliar) chemistry.
Manganate(VII) ions, MnO4-, oxidise hydrogen peroxide, H2O2, to oxygen
gas. The reaction is done with potassium manganate(VII) solution and
hydrogen peroxide solution acidified with dilute sulphuric acid.
During the reaction, the manganate(VII) ions are reduced to manganese(II)
ions.
Let's start with the hydrogen peroxide half-equation. What we know is:

The oxygen is already balanced. What about the hydrogen?


All you are allowed to add to this equation are water, hydrogen ions and
electrons. If you add water to supply the extra hydrogen atoms needed on
the right-hand side, you will mess up the oxygens again - that's obviously
wrong!
Add two hydrogen ions to the right-hand side.

Now all you need to do is balance the charges. You would have to add 2
electrons to the right-hand side to make the overall charge on both sides
zero.

http://www.chemguide.co.uk/inorganic/redox/equations.html (5 of 11)30/12/2004 11:11:16

Writing ionic equations for redox reactions

Now for the manganate(VII) half-equation:


You know (or are told) that the manganate(VII) ions turn into manganese
(II) ions. Write that down.

The manganese balances, but you need four oxygens on the right-hand
side. These can only come from water - that's the only oxygen-containing
thing you are allowed to write into one of these equations in acid
conditions.

By doing this, we've introduced some hydrogens. To balance these, you


will need 8 hydrogen ions on the left-hand side.

Now that all the atoms are balanced, all you need to do is balance the
charges. At the moment there are a net 7+ charges on the left-hand side
(1- and 8+), but only 2+ on the right. Add 5 electrons to the left-hand side
to reduce the 7+ to 2+.

This is the typical sort of half-equation which you will have to be able to
work out. The sequence is usually:

Balance the atoms apart from oxygen and hydrogen.

Balance the oxygens by adding water molecules.

Balance the hydrogens by adding hydrogen ions.

Balance the charges by adding electrons.

http://www.chemguide.co.uk/inorganic/redox/equations.html (6 of 11)30/12/2004 11:11:16

Writing ionic equations for redox reactions

Combining the half-reactions to make the ionic equation for the reaction
The two half-equations we've produced are:

You have to multiply the equations so that the same number of electrons
are involved in both. In this case, everything would work out well if you
transferred 10 electrons.

But this time, you haven't quite finished. During the checking of the
balancing, you should notice that there are hydrogen ions on both sides of
the equation:

You can simplify this down by subtracting 10 hydrogen ions from both
sides to leave the final version of the ionic equation - but don't forget to
check the balancing of the atoms and charges!

You will often find that hydrogen ions or water molecules appear on both
sides of the ionic equation in complicated cases built up in this way.
Always check, and then simplify where possible.
http://www.chemguide.co.uk/inorganic/redox/equations.html (7 of 11)30/12/2004 11:11:16

Writing ionic equations for redox reactions

Example 3: The oxidation of ethanol by acidified potassium


dichromate(VI)
This technique can be used just as well in examples involving organic
chemicals. Potassium dichromate(VI) solution acidified with dilute
sulphuric acid is used to oxidise ethanol, CH3CH2OH, to ethanoic acid,
CH3COOH.
The oxidising agent is the dichromate(VI) ion, Cr2O72-. This is reduced to
chromium(III) ions, Cr3+.
We'll do the ethanol to ethanoic acid half-equation first. Using the same
stages as before, start by writing down what you know:

Balance the oxygens by adding a water molecule to the left-hand side:

Add hydrogen ions to the right-hand side to balance the hydrogens:

And finally balance the charges by adding 4 electrons to the right-hand


side to give an overall zero charge on each side:

The dichromate(VI) half-equation contains a trap which lots of people fall


into!
Start by writing down what you know:

What people often forget to do at this stage is to balance the chromiums. If

http://www.chemguide.co.uk/inorganic/redox/equations.html (8 of 11)30/12/2004 11:11:16

Writing ionic equations for redox reactions

you don't do that, you are doomed to getting the wrong answer at the end
of the process! When you come to balance the charges you will have to
write in the wrong number of electrons - which means that your multiplying
factors will be wrong when you come to add the half-equations . . . A
complete waste of time!

Now balance the oxygens by adding water molecules . . .

. . . and the hydrogens by adding hydrogen ions:

Now all that needs balancing is the charges. Add 6 electrons to the lefthand side to give a net 6+ on each side.

Combining the half-reactions to make the ionic equation for the reaction
What we have so far is:

What are the multiplying factors for the equations this time? The simplest
way of working this out is to find the smallest number of electrons which
both 4 and 6 will divide into - in this case, 12. That means that you can
multiply one equation by 3 and the other by 2.

http://www.chemguide.co.uk/inorganic/redox/equations.html (9 of 11)30/12/2004 11:11:16

Writing ionic equations for redox reactions

Note: Don't worry too much if you get this wrong and choose
to transfer 24 electrons instead. All that will happen is that your
final equation will end up with everything multiplied by 2. Your
examiners might well allow that.

The multiplication and addition looks like this:

Now you will find that there are water molecules and hydrogen ions
occurring on both sides of the ionic equation. You can simplify this to give
the final equation:

Note: You have now seen a cross-section of the sort of


equations which you could be asked to work out. Now you
need to practice so that you can do this reasonably quickly and
very accurately! Aim to get an averagely complicated example
done in about 3 minutes.
If you want a few more examples, and the opportunity to
practice with answers available, you might be interested in
looking in chapter 1 of my book on Chemistry Calculations.

http://www.chemguide.co.uk/inorganic/redox/equations.html (10 of 11)30/12/2004 11:11:16

Writing ionic equations for redox reactions

Reactions done under alkaline conditions


Working out half-equations for reactions in alkaline solution is decidedly
more tricky than those above. You are fairly unlikely to be asked to do this
at A' level, and for that reason I've covered these on a separate page (link
below). It would be worthwhile checking your syllabus and past papers
before you start worrying about these!

Where would you like to go now?


How to work out half-equations for reactions under alkaline
conditions . . .
To the Redox menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/inorganic/redox/equations.html (11 of 11)30/12/2004 11:11:16

Definitions of oxidation and reduction (redox)

DEFINITIONS OF OXIDATION AND REDUCTION


(REDOX)

This page looks at the various definitions of oxidation and reduction


(redox) in terms of the transfer of oxygen, hydrogen and electrons. It also
explains the terms oxidising agent and reducing agent.

Oxidation and reduction in terms of oxygen transfer


Definitions

Oxidation is gain of oxygen.

Reduction is loss of oxygen.

For example, in the extraction of iron from its ore:

Because both reduction and oxidation are going on side-by-side, this is


known as a redox reaction.
Oxidising and reducing agents
An oxidising agent is substance which oxidises something else. In the
above example, the iron(III) oxide is the oxidising agent.
A reducing agent reduces something else. In the equation, the carbon
monoxide is the reducing agent.

http://www.chemguide.co.uk/inorganic/redox/definitions.html (1 of 5)30/12/2004 11:11:23

Definitions of oxidation and reduction (redox)

Oxidising agents give oxygen to another substance.

Reducing agents remove oxygen from another substance.

Oxidation and reduction in terms of hydrogen transfer


These are old definitions which aren't used very much nowadays. The
most likely place you will come across them is in organic chemistry.
Definitions

Oxidation is loss of hydrogen.

Reduction is gain of hydrogen.

Notice that these are exactly the opposite of the oxygen definitions.
For example, ethanol can be oxidised to ethanal:

You would need to use an oxidising agent to remove the hydrogen from
the ethanol. A commonly used oxidising agent is potassium dichromate
(VI) solution acidified with dilute sulphuric acid.

Note: The equation for this is rather complicated for this


introductory page. If you are interested, you will find a similar
example (ethanol to ethanoic acid) on the page dealing with
writing equations for redox reactions.

http://www.chemguide.co.uk/inorganic/redox/definitions.html (2 of 5)30/12/2004 11:11:23

Definitions of oxidation and reduction (redox)

Ethanal can also be reduced back to ethanol again by adding hydrogen


to it. A possible reducing agent is sodium tetrahydridoborate, NaBH4.
Again the equation is too complicated to be worth bothering about at this
point.

An update on oxidising and reducing agents

Oxidising agents give oxygen to another substance or remove


hydrogen from it.
Reducing agents remove oxygen from another substance or give
hydrogen to it.

Oxidation and reduction in terms of electron transfer


This is easily the most important use of the terms oxidation and reduction
at A' level.
Definitions

Oxidation is loss of electrons.

Reduction is gain of electrons.

It is essential that you remember these definitions. There is a very easy


way to do this. As long as you remember that you are talking about
electron transfer:

http://www.chemguide.co.uk/inorganic/redox/definitions.html (3 of 5)30/12/2004 11:11:23

Definitions of oxidation and reduction (redox)

A simple example
The equation shows a simple redox reaction which can obviously be
described in terms of oxygen transfer.

Copper(II) oxide and magnesium oxide are both ionic. The metals
obviously aren't. If you rewrite this as an ionic equation, it turns out that
the oxide ions are spectator ions and you are left with:

A last comment on oxidising and reducing agents


If you look at the equation above, the magnesium is reducing the copper
(II) ions by giving them electrons to neutralise the charge. Magnesium is
a reducing agent.
Looking at it the other way round, the copper(II) ions are removing
electrons from the magnesium to create the magnesium ions. The copper
(II) ions are acting as an oxidising agent.
Warning!
This is potentially very confusing if you try to learn both what oxidation
and reduction mean in terms of electron transfer, and also learn
definitions of oxidising and reducing agents in the same terms.
Personally, I would recommend that you work it out if you need it. The
argument (going on inside your head) would go like this if you wanted to
know, for example, what an oxidising agent did in terms of electrons:

http://www.chemguide.co.uk/inorganic/redox/definitions.html (4 of 5)30/12/2004 11:11:23

Definitions of oxidation and reduction (redox)

An oxidising agent oxidises something else.

Oxidation is loss of electrons (OIL RIG).

That means that an oxidising agent takes electrons from that other
substance.
So an oxidising agent must gain electrons.

Or you could think it out like this:

An oxidising agent oxidises something else.

That means that the oxidising agent must be being reduced.

Reduction is gain of electrons (OIL RIG).

So an oxidising agent must gain electrons.

Understanding is a lot safer than thoughtless learning!

Where would you like to go now?


To the Redox menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/inorganic/redox/definitions.html (5 of 5)30/12/2004 11:11:23

Redox menu

Understanding Chemistry

REDOX MENU

Definitions of oxidation and reduction . . .


Covers definitions of oxidation and reduction in terms of transfer of
oxygen, hydrogen and electrons.
Writing equations for redox reactions . . .
How to construct ionic equations for redox reactions by working
out electron half equations and then combining them.
Oxidation states (oxidation numbers) . . .
How to work out oxidation states (oxidation numbers) and use
them to decide simply what is being oxidised and what is being
reduced in a reaction.

Go to inorganic chemistry menu . . .


Go to Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/inorganic/redoxmenu.html30/12/2004 11:11:24

Oxidation states (oxidation numbers)

OXIDATION STATES (OXIDATION NUMBERS)

This page explains what oxidation states (oxidation numbers) are and
how to calculate them and make use of them.
Oxidation states are straightforward to work out and to use, but it is quite
difficult to define what they are in any quick way.

Explaining what oxidation states (oxidation numbers) are


Oxidation states simplify the whole process of working out what is being
oxidised and what is being reduced in redox reactions. However, for the
purposes of this introduction, it would be helpful if you knew about:

oxidation and reduction in terms of electron transfer

electron-half-equations

Note: If you aren't sure about either of these things, you


might want to look at the pages on redox definitions and
electron-half-equations. It would probably be best to read on
and come back to these links if you feel you need to.

http://www.chemguide.co.uk/inorganic/redox/oxidnstates.html (1 of 12)30/12/2004 11:11:32

Oxidation states (oxidation numbers)

We are going to look at some examples from vanadium chemistry. If you


don't know anything about vanadium, it doesn't matter in the slightest.
Vanadium forms a number of different ions - for example, V2+ and V3+. If
you think about how these might be produced from vanadium metal, the 2
+ ion will be formed by oxidising the metal by removing two electrons:

The vanadium is now said to be in an oxidation state of +2.


Removal of another electron gives the V3+ ion:

The vanadium now has an oxidation state of +3.


Removal of another electron gives a more unusual looking ion, VO2+.

The vanadium is now in an oxidation state of +4. Notice that the oxidation
state isn't simply counting the charge on the ion (that was true for the first
two cases but not for this one).
The positive oxidation state is counting the total number of electrons
which have had to be removed - starting from the element.
It is also possible to remove a fifth electron to give another ion (easily
confused with the one before!). The oxidation state of the vanadium is
now +5.

Every time you oxidise the vanadium by removing another electron from
it, its oxidation state increases by 1.
Fairly obviously, if you start adding electrons again the oxidation state
will fall. You could eventually get back to the element vanadium which
would have an oxidation state of zero.

http://www.chemguide.co.uk/inorganic/redox/oxidnstates.html (2 of 12)30/12/2004 11:11:32

Oxidation states (oxidation numbers)

What if you kept on adding electrons to the element? You can't actually
do that with vanadium, but you can with an element like sulphur.

The sulphur has an oxidation state of -2.


Summary
Oxidation state shows the total number of electrons which have been
removed from an element (a positive oxidation state) or added to an
element (a negative oxidation state) to get to its present state.
Oxidation involves an increase in oxidation state
Reduction involves a decrease in oxidation state
Recognising this simple pattern is the single most important thing about
the concept of oxidation states. If you know how the oxidation state of an
element changes during a reaction, you can instantly tell whether it is
being oxidised or reduced without having to work in terms of electron-halfequations and electron transfers.

Working out oxidation states


You don't work out oxidation states by counting the numbers of electrons
transferred. It would take far too long. Instead you learn some simple
rules, and do some very simple sums!

The oxidation state of an uncombined element is zero. That's


obviously so, because it hasn't been either oxidised or reduced
yet! This applies whatever the structure of the element - whether it
is, for example, Xe or Cl2 or S8, or whether it has a giant structure
like carbon or silicon.
The sum of the oxidation states of all the atoms or ions in a neutral
compound is zero.
The sum of the oxidation states of all the atoms in an ion is equal

http://www.chemguide.co.uk/inorganic/redox/oxidnstates.html (3 of 12)30/12/2004 11:11:32

Oxidation states (oxidation numbers)

to the charge on the ion.

The more electronegative element in a substance is given a


negative oxidation state. The less electronegative one is given a
positive oxidation state. Remember that fluorine is the most
electronegative element with oxygen second.
Some elements almost always have the same oxidation states in
their compounds:

element

usual oxidation
state

exceptions

Group 1 metals

always +1

Group 2 metals

always +2

Oxygen

usually -2

except in peroxides
and F2O (see below)

Hydrogen

usually +1

except in metal
hydrides where it is 1 (see below)

Fluorine

always -1

Chlorine

usually -1

except in
compounds with O
or F (see below)

The reasons for the exceptions


Hydrogen in the metal hydrides
Metal hydrides include compounds like sodium hydride, NaH. In this, the
hydrogen is present as a hydride ion, H-. The oxidation state of a simple
ion like hydride is equal to the charge on the ion - in this case, -1.

http://www.chemguide.co.uk/inorganic/redox/oxidnstates.html (4 of 12)30/12/2004 11:11:32

Oxidation states (oxidation numbers)

Alternatively, you can think of it that the sum of the oxidation states in a
neutral compound is zero. Since Group 1 metals always have an
oxidation state of +1 in their compounds, it follows that the hydrogen
must have an oxidation state of -1 (+1 -1 = 0).
Oxygen in peroxides
Peroxides include hydrogen peroxide, H2O2. This is an electrically
neutral compound and so the sum of the oxidation states of the hydrogen
and oxygen must be zero.
Since each hydrogen has an oxidation state of +1, each oxygen must
have an oxidation state of -1 to balance it.
Oxygen in F2O
The problem here is that oxygen isn't the most electronegative element.
The fluorine is more electronegative and has an oxidation state of -1. In
this case, the oxygen has an oxidation state of +2.
Chlorine in compounds with fluorine or oxygen
There are so many different oxidation states that chlorine can have in
these, that it is safer to simply remember that the chlorine doesn't have
an oxidation state of -1 in them, and work out its actual oxidation state
when you need it. You will find an example of this below.

Warning!
Don't get too bogged down in these exceptions. In most of the cases you
will come across, they don't apply!

Examples of working out oxidation states


What is the oxidation state of chromium in Cr2+?
That's easy! For a simple ion like this, the oxidation state is the charge on
http://www.chemguide.co.uk/inorganic/redox/oxidnstates.html (5 of 12)30/12/2004 11:11:32

Oxidation states (oxidation numbers)

the ion - in other words: +2 (Don't forget the + sign.)


What is the oxidation state of chromium in CrCl3?
This is a neutral compound so the sum of the oxidation states is zero.
Chlorine has an oxidation state of -1. If the oxidation state of chromium is
n:
n + 3(-1) = 0
n = +3 (Again, don't forget the + sign!)
What is the oxidation state of chromium in Cr(H2O)63+?
This is an ion and so the sum of the oxidation states is equal to the
charge on the ion. There is a short-cut for working out oxidation states in
complex ions like this where the metal atom is surrounded by electrically
neutral molecules like water or ammonia.
The sum of the oxidation states in the attached neutral molecule must be
zero. That means that you can ignore them when you do the sum. This
would be essentially the same as an unattached chromium ion, Cr3+. The
oxidation state is +3.
What is the oxidation state of chromium in the dichromate ion, Cr2O72-?
The oxidation state of the oxygen is -2, and the sum of the oxidation
states is equal to the charge on the ion. Don't forget that there are 2
chromium atoms present.
2n + 7(-2) = -2
n = +6

http://www.chemguide.co.uk/inorganic/redox/oxidnstates.html (6 of 12)30/12/2004 11:11:32

Oxidation states (oxidation numbers)

Warning: Because these are simple sums it is tempting to


try to do them in your head. If it matters (like in an exam)
write them down using as many steps as you need so that
there is no chance of making careless mistakes. Your
examiners aren't going to be impressed by your mental
arithmetic - all they want is the right answer!
If you want some more examples to practice on, you will find
them in most text books, including my chemistry calculations
book.

What is the oxidation state of copper in CuSO4?


Unfortunately, it isn't always possible to work out oxidation states by a
simple use of the rules above. The problem in this case is that the
compound contains two elements (the copper and the sulphur) whose
oxidation states can both change.
The only way around this is to know some simple chemistry! There are
two ways you might approach it. (There might be others as well, but I
can't think of them at the moment!)

You might recognise this as an ionic compound containing copper


ions and sulphate ions, SO42-. To make an electrically neutral
compound, the copper must be present as a 2+ ion. The oxidation
state is therefore +2.
You might recognise the formula as being copper(II) sulphate. The
"(II)" in the name tells you that the oxidation state is 2 (see below).
You will know that it is +2 because you know that metals form
positive ions, and the oxidation state will simply be the charge on
the ion.

Using oxidation states

http://www.chemguide.co.uk/inorganic/redox/oxidnstates.html (7 of 12)30/12/2004 11:11:32

Oxidation states (oxidation numbers)

In naming compounds
You will have come across names like iron(II) sulphate and iron(III)
chloride. The (II) and (III) are the oxidation states of the iron in the two
compounds: +2 and +3 respectively. That tells you that they contain Fe2+
and Fe3+ ions.
This can also be extended to the negative ion. Iron(II) sulphate is FeSO4.
There is also a compound FeSO3 with the old name of iron(II) sulphite.
The modern names reflect the oxidation states of the sulphur in the two
compounds.
The sulphate ion is SO42-. The oxidation state of the sulphur is +6 (work
it out!). The ion is more properly called the sulphate(VI) ion.
The sulphite ion is SO32-. The oxidation state of the sulphur is +4 (work
that out as well!). This ion is more properly called the sulphate(IV) ion.
The ate ending simply shows that the sulphur is in a negative ion.
So FeSO4 is properly called iron(II) sulphate(VI), and FeSO3 is iron(II)
sulphate(IV). In fact, because of the easy confusion between these
names, the old names sulphate and sulphite are normally still used in
introductory chemistry courses.

Note: Even these aren't the full name! The oxygens in the
negative ions should also be identified. FeSO4 is properly
called iron(II) tetraoxosulphate(VI). It all gets a bit out of hand
for everyday use for common ions.

http://www.chemguide.co.uk/inorganic/redox/oxidnstates.html (8 of 12)30/12/2004 11:11:32

Oxidation states (oxidation numbers)

Using oxidation states to identify what's been oxidised and what's


been reduced
This is easily the most common use of oxidation states.
Remember:
Oxidation involves an increase in oxidation state
Reduction involves a decrease in oxidation state
In each of the following examples, we have to decide whether the
reaction involves redox, and if so what has been oxidised and what
reduced.
Example 1:
This is the reaction between magnesium and hydrochloric acid or
hydrogen chloride gas:

Have the oxidation states of anything changed? Yes they have - you
have two elements which are in compounds on one side of the equation
and as uncombined elements on the other. Check all the oxidation states
to be sure:.

The magnesium's oxidation state has increased - it has been oxidised.


The hydrogen's oxidation state has fallen - it has been reduced. The
chlorine is in the same oxidation state on both sides of the equation - it
hasn't been oxidised or reduced.
Example 2:
http://www.chemguide.co.uk/inorganic/redox/oxidnstates.html (9 of 12)30/12/2004 11:11:32

Oxidation states (oxidation numbers)

The reaction between sodium hydroxide and hydrochloric acid is:

Checking all the oxidation states:

Nothing has changed. This isn't a redox reaction.


Example 3:
This is a sneaky one! The reaction between chlorine and cold dilute
sodium hydroxide solution is:

Obviously the chlorine has changed oxidation state because it has ended
up in compounds starting from the original element. Checking all the
oxidation states shows:

The chlorine is the only thing to have changed oxidation state. Has it
been oxidised or reduced? Yes! Both! One atom has been reduced
because its oxidation state has fallen. The other has been oxidised.
This is a good example of a disproportionation reaction. A
disproportionation reaction is one in which a single substance is both
oxidised and reduced.

http://www.chemguide.co.uk/inorganic/redox/oxidnstates.html (10 of 12)30/12/2004 11:11:32

Oxidation states (oxidation numbers)

Using oxidation states to work out reacting proportions


This is sometimes useful where you have to work out reacting
proportions for use in titration reactions where you don't have enough
information to work out the complete ionic equation.
Remember that each time an oxidation state changes by one unit, one
electron has been transferred. If one substance's oxidation state in a
reaction falls by 2, that means that it has gained 2 electrons.
Something else in the reaction must be losing those electrons. Any
oxidation state fall by one substance must be accompanied by an equal
oxidation state increase by something else.
This example is based on information in an old AQA A' level question.
Ions containing cerium in the +4 oxidation state are oxidising agents.
(They are more complicated than just Ce4+.) They can oxidise ions
containing molybdenum from the +2 to the +6 oxidation state (from Mo2+
to MoO42-). In the process the cerium is reduced to the +3 oxidation state
(Ce3+). What are the reacting proportions?
The oxidation state of the molybdenum is increasing by 4. That means
that the oxidation state of the cerium must fall by 4 to compensate.
But the oxidation state of the cerium in each of its ions only falls from +4
to +3 - a fall of 1. So there must obviously be 4 cerium ions involved for
each molybdenum ion.
The reacting proportions are 4 cerium-containing ions to 1 molybdenum
ion.

Where would you like to go now?


To the Redox menu . . .
To the Inorganic Chemistry menu . . .
http://www.chemguide.co.uk/inorganic/redox/oxidnstates.html (11 of 12)30/12/2004 11:11:32

Oxidation states (oxidation numbers)

To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/inorganic/redox/oxidnstates.html (12 of 12)30/12/2004 11:11:32

calculations in as / a level chemistry

Understanding Chemistry

CALCULATIONS IN AS / A LEVEL CHEMISTRY


There are virtually no calculations on
this site, but you might be interested
in the author's book covering all the
calculations required by the UK A'
level syllabuses.
If you have found this site helpful,
you should find the book will help
you as well.
Students working in other countries
should also find the book useful.
Compare what you are expected to
know with the list of things that the
book covers (see below).
On this page you will find a description of how the book is organised,
together with a full list of the contents. You will also find a direct link to
the book on the Amazon.co.uk site.
School Science Review, March 2001
"This is an excellent book, . . ." " . . . a 'must have' for teachers and
students alike."
Times Educational Supplement, 27 April 2001
". . fills a large gap in the market and provides excellent
coverage . ."
"I shall recommend this to my students as a self-study aid and will
certainly use it myself."

http://www.chemguide.co.uk/book.html (1 of 5)30/12/2004 11:11:35

calculations in as / a level chemistry

How to get hold of the book


You can, of course, buy the book through normal book sellers, but if if
you want to buy online, you will find a direct link to Amazon.co.uk coming
up. (Note that Amazon mistakenly refers to the book as "Calculations in
A Level Chemistry".)
Non-UK students can also buy the book from Amazon.co.uk, but will
obviously have to pay a slightly higher delivery charge.

Note: If your usual source of books is Amazon.com, you


should compare the price for the book (including delivery)
from Amazon.com with the price from Amazon.co.uk - even if
you live in North America.
You may well find that it is significantly cheaper to buy from
Amazon.co.uk and have it sent by air mail across the Atlantic,
than to buy it in America!

Have a look at the book on the Amazon site

How the book works


Each type of calculation is introduced in a very gentle way, making no
great assumptions about your chemistry knowledge or maths ability.
There are then lots of worked examples, gradually getting more difficult
and showing as many variations on a calculation as possible.
At the end of each section there are a set of problems for you to do,
based firmly on what has gone before. At the back of the book, you will
http://www.chemguide.co.uk/book.html (2 of 5)30/12/2004 11:11:35

calculations in as / a level chemistry

find complete worked solutions to these problems.


At the end of each chapter, you will find another set of problems covering
the ground again. This time, numerical answers are given for the
problems - but no worked solutions.

What the book covers


Chapter 1: Formulae, Equations and Oxidation States
This chapter starts from a very simple level and teaches you how to write
formulae and to work out equations, including ionic equations. It also
covers oxidation states and their uses. Although it starts simply, by the
end of this chapter you will be able to write equations for some very
complicated reactions.
Chapter 2: Basic Calculations Involving Formulae and Equations
Covers relative atomic mass (and how it can be found from a mass
spectrum), relative formula mass, moles (including the Avogadro
Constant), finding empirical and molecular formulae, calculations from
equations, and percentage yield calculations.
Chapter 3: Basic Calculations Involving Gases
Includes Avogadro's Law and its uses, the molar volume of a gas
(including calculations from equations involving gases), and the ideal gas
equation.
Chapter 4: Basic Calculations Involving Solutions
Describes how to do calculations from equations involving solutions
(including how to derive equations from experiments), and then looks in
detail at titration calculations. It starts with simple acid-base titrations and
then gradually progresses to more difficult examples, including back
titrations and "double indicator" titrations.
Chapter 5: Thermochemistry

http://www.chemguide.co.uk/book.html (3 of 5)30/12/2004 11:11:35

calculations in as / a level chemistry

Starts by looking at how you process the results of thermochemistry


experiments, and then goes on to look at all the variations on
thermochemistry calculations that you are likely to meet at either AS or
A'level.
Chapter 6: Orders of Reaction
Starts by explaining what is meant by orders of reaction, and then looks
in detail at how you can work them out from experimental data - by
exploring tabulated data from initial rate experiments or by processing
results graphically. It also considers the effect of temperature on rates of
reaction, and how this can be used to find activation energies.
Chapter 7: Chemical Equilibria
Looks at calculations involving both Kc and Kp for both homogeneous
and heterogeneous equilibria. The work on Kp necessarily covers such
things as mole fractions and partial pressures.
Chapter 8: Acid - Base Equilibria
Defines pH and shows how to work out the pH of strong acids and bases
and of weak acids and bases. As a part of this, it necessarily covers
things like Kw, Ka and pKa. Also shows how to calculate pH changes
during simple acid-base titrations. Discusses how buffer solutions work
and how to calculate the pH's of both acidic and basic buffers.
Chapter 9: Other Equilibria
Covers solubility product, partition coefficients and Henry's Law.
Chapter 10: Redox Equilibria
Starts with a long introductory section to help you to understand the
origin of electrode potentials (redox potentials) - because most text
books try to do this too quickly and leave you confused! Looks at simple
calculations involving electrode potentials, and how they can be used to
predict the feasibility of redox reactions.
Chapter 11: Entropy and Free Energy

http://www.chemguide.co.uk/book.html (4 of 5)30/12/2004 11:11:35

calculations in as / a level chemistry

Discusses entropy and entropy changes in as simple a way as possible and looks at the way total entropy change governs the feasibility of a
reaction. Goes on to look at free energy changes and how they can be
calculated and used to predict feasibility.

Go to Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/book.html (5 of 5)30/12/2004 11:11:35

Understanding Chemistry - Inorganic Chemistry Menu

Understanding Chemistry

INORGANIC CHEMISTRY MENU

Basic ideas about oxidation and reduction . . .


Covers oxidation states (oxidation numbers), definitions of
oxidation and reduction, and how you work out equations for redox
reactions.
Some chemistry of the Group 2 elements . . .
Covers the elements beryllium (Be), magnesium (Mg), calcium
(Ca), strontium (Sr) and barium (Ba). Includes trends in atomic and
physical properties, trends in reactivity, the solubility patterns in
the hydroxides and sulphates, trends in the thermal decomposition
of the nitrates and carbonates, and some of the atypical properties
of beryllium.
Some chemistry of the Group 4 elements . . .
Covers the limited amount of Group 4 chemistry (carbon, silicon,
germanium, tin and lead) required by UK A level syllabuses. In
particular, it covers the trend from non-metal to metal as you go
down the group, and the increasing tendency towards an oxidation
state of +2. Also a certain amount of chemistry of the chlorides
and oxides.
Some chemistry of the Group 7 elements (the halogens) . . .
Covers the elements fluorine (F), chlorine (Cl), bromine (Br) and
iodine (I). Includes trends in atomic and physical properties, the
redox properties of the halogens and their ions, the acidity of the
hydrogen halides, and the tests for the halide ions.

http://www.chemguide.co.uk/inorgmenu.html (1 of 2)30/12/2004 11:11:37

Understanding Chemistry - Inorganic Chemistry Menu

Some essential complex ion chemistry . . .


An introduction to the complex ions formed by transition and other
metals. Includes bonding, shapes and names, and a simple look at
the origin of colour. Looks in detail at various ligand exchange
reactions and the chemistry of common hexaaqua ions.
Some essential transition metal chemistry . . .
A summary of the general features of transition metals together
with most of the reactions required for UK A' level syllabuses.

Go to Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/inorgmenu.html (2 of 2)30/12/2004 11:11:37

Periodic Table Group 2 menu

Understanding Chemistry

PERIODIC TABLE GROUP 2 MENU

Atomic and physical properties . . .


Discusses trends in atomic radius, ionisation energy,
electronegativity and melting point of the Group 2 elements.
Reactions with water . . .
Looks at the trends in the reactions between the Group 2 elements
and water.
Reactions with oxygen . . .
Looks at the trends in the reactions between the Group 2 elements
and oxygen.
The solubility of the hydroxides, sulphates and carbonates. . .
Describes the patterns in the solubilities of the hydroxides,
sulphates and carbonates of the Group 2 elements.
The thermal stability of the nitrates and carbonates . . .
Describes and explains patterns in the effect of heat on the
nitrates and carbonates of the Group 2 elements.
Some atypical properties of beryllium compounds . . .
Describes and explains some cases where beryllium compounds
don't fit the patterns shown by the rest of the Group.

http://www.chemguide.co.uk/inorganic/group2menu.html (1 of 2)30/12/2004 11:11:38

Periodic Table Group 2 menu

Go to inorganic chemistry menu . . .


Go to Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/inorganic/group2menu.html (2 of 2)30/12/2004 11:11:38

Atomic and physical properties of Periodic Table Group 2

ATOMIC AND PHYSICAL PROPERTIES OF THE


GROUP 2 ELEMENTS

This page explores the trends in some atomic and physical properties of
the Group 2 elements - beryllium, magnesium, calcium, strontium and
barium. You will find separate sections below covering the trends in
atomic radius, first ionisation energy, electronegativity and melting point.
Even if you aren't currently interested in all these things, it would
probably pay you to read the whole page. The same ideas tend to recur
throughout the atomic properties, and you may find that earlier
explanations help to you understand later ones.

Trends in Atomic Radius

Note: You will find atomic radius covered in detail in another


part of this site. If you choose to follow this link, use the
BACK button on your browser to return quickly to this page.

http://www.chemguide.co.uk/inorganic/group2/properties.html (1 of 11)30/12/2004 11:11:51

Atomic and physical properties of Periodic Table Group 2

You can see that the atomic radius increases as you go down the Group.
Notice that beryllium has a particularly small atom compared with the rest
of the Group.
Explaining the increase in atomic radius
The radius of an atom is governed by

the number of layers of electrons around the nucleus

the pull the outer electrons feel from the nucleus.

Compare beryllium and magnesium:


Be

1s22s2

Mg

1s22s22p63s2
Note: If you aren't sure about writing electronic structures
using s and p notation it might be a good idea to follow this
link before you go on. Use the BACK button on your browser
to return quickly to this page.

In each case, the two outer electrons feel a net pull of 2+ from the
nucleus. The positive charge on the nucleus is cut down by the
negativeness of the inner electrons.

http://www.chemguide.co.uk/inorganic/group2/properties.html (2 of 11)30/12/2004 11:11:51

Atomic and physical properties of Periodic Table Group 2

This is equally true for all the other atoms in Group 2. Work it out for
calcium if you aren't convinced.
The only factor which is going to affect the size of the atom is therefore
the number of layers of inner electrons which have to be fitted in around
the atom. Obviously, the more layers of electrons you have, the more
space they will take up - electrons repel each other. That means that the
atoms are bound to get bigger as you go down the Group.

Note: You may think that this is all a bit long-winded! It is,
after all, fairly obvious that atoms will get bigger if you add
more layers of electrons. Why, then, bother about exploring
the net pull on the electrons from the centre of the atom?
It is a matter of setting up good habits. If you are talking
about atoms in the same Group, the net pull from the centre
will always be the same - and you could ignore it without
creating problems. That isn't true if you try to compare atoms
from different parts of the Periodic Table. If you don't get into
the habit of thinking about all the possible factors, you are
going to make mistakes.

http://www.chemguide.co.uk/inorganic/group2/properties.html (3 of 11)30/12/2004 11:11:51

Atomic and physical properties of Periodic Table Group 2

Trends in First Ionisation Energy


First ionisation energy is the energy needed to remove the most loosely
held electrons from one mole of gaseous atoms to make one mole of
singly charged gaseous ions - in other words, for 1 mole of this process:

Note: You will find ionisation energy covered in detail in


another part of this site. If you choose to follow this link, use
the BACK button on your browser to return quickly to this
page.

Notice that first ionisation energy falls as you go down the group.
Explaining the decrease in first ionisation energy
Ionisation energy is governed by

the charge on the nucleus,

http://www.chemguide.co.uk/inorganic/group2/properties.html (4 of 11)30/12/2004 11:11:51

Atomic and physical properties of Periodic Table Group 2

the amount of screening by the inner electrons,

the distance between the outer electrons and the nucleus.

As you go down the Group, the increase in nuclear charge is exactly


offset by the increase in the number of inner electrons. Just as when we
were talking about atomic radius further up this page, in each of the
elements in this Group, the outer electrons feel a net attraction of 2+ from
the centre.
However, as you go down the Group, the distance between the nucleus
and the outer electrons increases and so they become easier to remove the ionisation energy falls.

Trends in Electronegativity
Electronegativity is a measure of the tendency of an atom to attract a
bonding pair of electrons. It is usually measured on the Pauling scale, on
which the most electronegative element (fluorine) is given an
electronegativity of 4.0.

Note: You will find electronegativity covered in detail in


another part of this site. If you choose to follow this link, use
the BACK button on your browser to return quickly to this
page.

http://www.chemguide.co.uk/inorganic/group2/properties.html (5 of 11)30/12/2004 11:11:51

Atomic and physical properties of Periodic Table Group 2

All of these elements have a low electronegativity. (Remember that the


most electronegative element, fluorine, has an electronegativity of 4.0.)
Notice that electronegativity falls as you go down the Group. The atoms
become less and less good at attracting bonding pairs of electrons.

Note: You might argue that the fall doesn't apply throughout
the Group because both calcium and strontium have an
electronegativity of 1.0. This is probably most easily
explained by the fact that electronegativities are only
recorded to 1 decimal place. If, for example, calcium was
1.04 and strontium was 0.95 (made-up numbers!), both
would appear to have an electronegativity of 1.0.

http://www.chemguide.co.uk/inorganic/group2/properties.html (6 of 11)30/12/2004 11:11:51

Atomic and physical properties of Periodic Table Group 2

Explaining the decrease in electronegativity


Imagine a bond between a magnesium atom and a chlorine atom. Think
of it to start with as a covalent bond - a pair of shared electrons. The
electron pair will be dragged towards the chlorine end because there is a
much greater net pull from the chlorine nucleus than from the
magnesium one.

The electron pair ends up so close to the chlorine that there is essentially
a transfer of an electron to the chlorine - ions are formed.
The large pull from the chlorine nucleus is why chlorine is much more
electronegative than magnesium is.
Now compare this with the beryllium-chlorine bond.
The net pull from each end of the bond is the same as before, but you
have to remember that the beryllium atom is smaller than a magnesium
atom. That means that the electron pair is going to be closer to the net 2
+ charge from the beryllium end, and so more strongly attracted to it.

http://www.chemguide.co.uk/inorganic/group2/properties.html (7 of 11)30/12/2004 11:11:51

Atomic and physical properties of Periodic Table Group 2

In this case, the electron pair doesn't get attracted close enough to the
chlorine for an ionic bond to be formed. Because of its small size,
beryllium forms covalent bonds, not ionic ones. The attraction between
the beryllium nucleus and a bonding pair is always too great for ions to
be formed.
Summarising the trend down the Group
As the metal atoms get bigger, any bonding pair gets further and further
away from the metal nucleus, and so is less strongly attracted towards it.
In other words, as you go down the Group, the elements become less
electronegative.
As you go down the Group, the bonds formed between these elements
and other things such as chlorine become more and more ionic. The
bonding pair is increasingly attracted away from the Group 2 element
towards the chlorine (or whatever).

Trends in Melting Point

http://www.chemguide.co.uk/inorganic/group2/properties.html (8 of 11)30/12/2004 11:11:51

Atomic and physical properties of Periodic Table Group 2

You will see that (with the exception of magnesium) the melting point
falls as you go down the Group.
Explaining the trends in melting point (or not!)
Thankfully, none of the current A'level syllabuses ask you to explain the
pattern in the melting points. Trying to explain patterns in melting point is
notoriously difficult, and simple explanations usually come to grief if you
look at them closely.
It would be tempting to say that (apart from magnesium) the lower
melting points as you go down the Group reflect weaker metallic bonds,
but this is dangerous. The metallic bond isn't fully broken on melting.
Boiling point is usually a better guide to the strengths of the forces
involved.
You will sometimes come across an explanation for the low melting point
of magnesium in terms of the way the atoms pack in the crystal. It isn't
very difficult to pick holes in this argument and, in any case, the boiling
point of magnesium is also surprisingly low - and packing is completely
irrelevant when you are thinking about a liquid. In the case of
magnesium, there must be some factor producing a relatively weak
metallic bond. (I have no idea what that might be!)

http://www.chemguide.co.uk/inorganic/group2/properties.html (9 of 11)30/12/2004 11:11:51

Atomic and physical properties of Periodic Table Group 2

You will see that there is no obvious pattern in the boiling points implying that there is no simple pattern in the strengths of the metallic
bonds.
Another measure of the strengths of the metallic bonds might be
atomisation energy. This is the energy needed to produce 1 mole of
separated atoms in the gas state starting from the element in its standard
state (the state you would expect it to be in at approximately room
temperature and pressure).

And again there is no simple pattern. Don't worry about any of this for
A'level purposes, but be very wary of any explanation which looks
http://www.chemguide.co.uk/inorganic/group2/properties.html (10 of 11)30/12/2004 11:11:51

Atomic and physical properties of Periodic Table Group 2

deceptively simple!

Help please! As I understand it (from Inorganic Chemistry


by A.G.Sharpe) there is no easy explanation for the
variations in the physical data in Group 2. However, if
anybody reading this knows of a simple (words of one
syllable!) and reliable explanation for these melting point,
boiling point and atomisation energy figures (particularly the
very low magnesium values), please contact me via the link
on the about this site page.

Where would you like to go now?


To the Group 2 menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/inorganic/group2/properties.html (11 of 11)30/12/2004 11:11:51

Understanding Chemistry - About this site

Understanding Chemistry

ABOUT THIS SITE


The Site
Aims
This site is intended to cover the needs of UK A level chemistry students,
although it is actually being used by students on equivalent (16 to 18
year old) courses worldwide. It is also being used by students at the
beginning of university level courses.
I started by concentrating on the bits that textbooks tend to do too quickly
and that students often find difficulty with. My over-riding aim is to try to
increase your understanding of these difficult bits so that you gain
confidence.
More recently I have been adding more and more factual content, but
always with a concentration on understanding. I'm simply not interested
in the "learn this parrot fashion and you will pass" approach - it is much
more of a slog than taking a bit of time to understand what is going on.

The structure of the site - an apology!


The site covers a large amount of material spread over lots of pages
(and still growing for the foreseeable future). One of the down-sides of
developing the site starting from the hard bits and then adding to it later
is that you will often find related topics in different parts of the site.
I have tried to cross-refer everything as far as possible, but I know the
structure of the site isn't as ideal as it could be in places. Unfortunately, I
am now stuck with it! There are so many other sites linking to specific
pages that I can't now reorganise things without creating chaos.

http://www.chemguide.co.uk/about.html (1 of 4)30/12/2004 11:11:55

Understanding Chemistry - About this site

You should be able to find what you want via the menu systems or using
the Google site search in the Main Menu. However, it may well be that I
haven't written exactly what you want yet - in which case, you will need to
look elsewhere. Please bear with me - writing new stuff takes a long time!

Helping yourself
To get the best from the site you must have a clear idea of what you
need to know. You will often find comments suggesting that you should
refer to your syllabus or to recent exam papers. It is important that you
have these to hand.
UK A level students can easily download syllabuses from their Exam
Boards, and should be able to find other useful things on those websites.
You can find links to these sites via the Main Menu.
If you are working in another examination system, you still need to know
exactly what is expected of you, but will have to find that information for
yourself.

About the author


Jim Clark

is a Cambridge graduate with over 30 years of teaching


experience at this level.
was Head of Chemistry and then Head of Science at Truro School
in Cornwall.
took early retirement in 1997 and has since concentrated on
writing.
is the author of Calculations in AS/A Level Chemistry published by
Longman (September 2000). Follow the link to find out more about
the book.

http://www.chemguide.co.uk/about.html (2 of 4)30/12/2004 11:11:55

Understanding Chemistry - About this site

is the author of Longman GCSE Chemistry published by Longman


(May 2002). Follow this link to find out more about the book.

Contacting the author


If you have problems understanding anything you have found on this site,
then please get in touch so that I can put it right. If you e-mail me I will do
my best to answer within 24 hours. Obviously there will be times when
this is impossible (if I am on holiday, for example). Be patient - I will reply
eventually!
Because of the huge and growing number of people now visiting the site
every month, I'm afraid it is no longer possible to answer questions about
things which aren't specifically related to what I have already written.
Please use the address:

Copying from this site


You may use the information on this site in any way you wish - including
making multiple copies for other people's use - provided that

you make clear where the information came from;

nobody makes any financial gain from it.

Go to Main Menu . . .

Jim Clark 2004


http://www.chemguide.co.uk/about.html (3 of 4)30/12/2004 11:11:55

Understanding Chemistry - About this site

http://www.chemguide.co.uk/about.html (4 of 4)30/12/2004 11:11:55

Periodic Table Group 4 menu

Understanding Chemistry

PERIODIC TABLE GROUP 4 MENU

The trend from non-metal to metal . . .


Discusses the trend from non-metallic to metallic behaviour as you
go down Group 4. Describes the structures and physical properties
of the elements, and some important atomic properties.
Oxidation state trends . . .
Looks at the increasing tendency towards an oxidation state of +2
as you go down the Group.
Group 4 chlorides . . .
Describes and explains the differences between the chlorides of
carbon, silicon and lead as required by UK A level chemistry
syllabuses.
Group 4 oxides . . .
Describes and explains the differences between the dioxides of
carbon and silicon, and looks at the trend in acid-base behaviour
of the oxides as you go down the Group.
Some chemistry of aqueous lead(II) ions . . .
Describes the reactions of lead(II) ions in solution with hydroxide
ions, chloride ions, iodide ions and sulphate ions.

http://www.chemguide.co.uk/inorganic/group4menu.html (1 of 2)30/12/2004 11:11:58

Periodic Table Group 4 menu

Go to inorganic chemistry menu . . .


Go to Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/inorganic/group4menu.html (2 of 2)30/12/2004 11:11:58

The trend from non-metal to metal in Group 4

THE TREND FROM NON-METAL TO METAL IN


THE GROUP 4 ELEMENTS

This page explores the trend from non-metallic to metallic behaviour in


the Group 4 elements - carbon (C), silicon (Si), germanium (Ge), tin (Sn)
and lead (Pb). It describes how this trend is shown in the structures and
physical properties of the elements, and finally makes a not entirely
successful attempt to explain the trend.

Structures and Physical Properties


Structures of the elements
The trend from non-metal to metal as you go down the Group is clearly
seen in the structures of the elements themselves.
Carbon at the top of the Group has giant covalent structures in its two
most familiar allotropes - diamond and graphite.

Allotropes: Two or more forms of the same element in the


same physical state.
The structures of diamond and graphite are explored in more
detail on a page about giant covalent structures in another
part of this site. It would probably be worth your while to read
this page before you go any further.
Use the BACK button on your browser to return quickly to this
page.

http://www.chemguide.co.uk/inorganic/group4/properties.html (1 of 12)30/12/2004 11:12:10

The trend from non-metal to metal in Group 4

Diamond has a three-dimensional structure of carbon atoms each joined


covalently to 4 other atoms. The diagram shows a small part of that
structure.

Exactly this same structure is found in silicon and germanium and in one
of the allotropes of tin - "grey tin" or "alpha-tin".
The common allotrope of tin ("white tin" or "beta-tin") is metallic and has
its atoms held together by metallic bonds. The structure is a distorted
close-packed arrangement. In close-packing, each atom is surrounded
by 12 near-neighbours.
By the time you get to lead, the atoms are arranged in a straightforward
12-co-ordinated metallic structure.

Note: If you aren't sure about metallic bonding or metallic


structures, you should follow these links before you go any
further. The first link will actually lead you to the second one if
you want to explore both of these topics.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/inorganic/group4/properties.html (2 of 12)30/12/2004 11:12:10

The trend from non-metal to metal in Group 4

There is therefore a clear trend from the typical covalency found in nonmetals to the metallic bonding in metals, with the change-over obvious in
the two entirely different structures found in tin.

Physical properties of the elements


Melting points and boiling points
If you look at the trends in melting and boiling points as you go down
Group 4, it is very difficult to make any sensible comments about the shift
from covalent to metallic bonding. The trends reflect the increasing
weakness of the covalent or metallic bonds as the atoms get bigger and
the bonds get longer.

The low value for tin's melting point compared with lead is presumably
due to tin forming a distorted 12-co-ordinated structure rather than a pure
one. The tin values in the chart refer to metallic white tin.

Note: The data in this chart comes from the University of


Sheffield's excellent Webelements site. There is an awful lot
of variability in the data depending on where you get it from. I
have to admit to choosing this set because it shows simple,
largely unbroken patterns!

http://www.chemguide.co.uk/inorganic/group4/properties.html (3 of 12)30/12/2004 11:12:10

The trend from non-metal to metal in Group 4

Brittleness
There is a much clearer non-metal / metal difference shown if you look at
the brittleness of the elements.
Carbon as diamond is, of course, very hard - reflecting the strength of the
covalent bonds. However, if you hit it with a hammer, it shatters. Once
you apply enough energy to break the existing carbon-carbon bonds,
that's it!
Silicon, germanium and grey tin (all with the same structure as diamond)
are also brittle solids.
However, white tin and lead have metallic structures. The atoms can roll
over each other without any permanent disruption of the metallic bonds leading to typical metallic properties like being malleable and ductile.
Lead in particular is a fairly soft metal.

Electrical conductivity
Carbon as diamond doesn't conduct electricity. In diamond the electrons
are all tightly bound and not free to move.

Note: In graphite, each atom donates one electron to a


delocalised system of electrons which takes in the whole of
its layer. These electrons are free to move around, and so
graphite conducts electricity - but this is a special case.
If you are interested, the bonding in graphite is like a vastly
extended version of the bonding in benzene. Each carbon
atom undergoes sp2 hybridisation, and then the unhybridised
p orbitals on each carbon atom overlap sideways to give a
massive pi system above and below the plane of the sheet of
atoms.

http://www.chemguide.co.uk/inorganic/group4/properties.html (4 of 12)30/12/2004 11:12:11

The trend from non-metal to metal in Group 4

Unlike diamond (which doesn't conduct electricity), silicon, germanium


and grey tin are semiconductors.

Semiconductors: The theory of semiconductors lies


outside A level chemistry, but briefly . . .
When lots of atoms come together to make a giant structure,
their atomic orbitals merge to produce a huge number of
molecular orbitals, which arrange themselves in bands of
increasing energy. One of these is often described as a
valence band. The molecular orbitals in this band hold the
electrons which make up the normal covalent (or metallic)
bonding.
The other band is called the conductance band. This usually
has a higher energy than the valence band, and in something
like diamond or silicon at absolute zero, the conductance
band is empty of electrons.
However, as electrons gain thermal energy as the
temperature increases, some electrons may jump from the
valence band into the conductance band - especially if the
gap between the two is small. Once they are in the
conductance band, they are delocalised from their original
atoms and are free to move and conduct electricity.
In diamond, the energy gap between the valence band and
conductance band is too high for this to happen. In silicon,
the band gap is small enough for electrons to jump, and so
silicon is a semiconductor.
If you are interested in this, you might like to try a Google
search on silicon semiconductors band theory (or similar).
There is a Google search box on the Main Menu page (link at
the bottom of this page). Don't forget to check the "search
www" box if you do this - otherwise you will end up back here
again!

http://www.chemguide.co.uk/inorganic/group4/properties.html (5 of 12)30/12/2004 11:12:11

The trend from non-metal to metal in Group 4

White tin and lead are normal metallic conductors of electricity.


There is therefore a clear trend from the typically non-metallic
conductivity behaviour of carbon as diamond, and the typically metallic
behaviour of white tin and lead.

Trying to explain the trends


The main characteristic of metals is that they form positive ions. What we
need to do is to look at the factors which increase the likelihood of
positive ions being formed as you go down Group 4.
Electronegativity
Electronegativity is a measure of the tendency of an atom to attract a
bonding pair of electrons. It is usually measured on the Pauling scale,
where the most electronegative element (fluorine) is given an
electronegativity of 4.
The lower the electronegativity of an atom, the less strongly the atom
attracts a bonding pair of electrons. That means that this atom will tend
to lose the electron pair towards whatever else it is attached to. The atom
we are interested in will therefore tend to carry either a partial positive
charge or form a positive ion.
Metallic behaviour is usually associated with a low electronegativity.

Note: If you aren't sure about electronegativity you really


ought to read about it before you go any further.
Use the BACK button on your browser to return quickly to this
page.

http://www.chemguide.co.uk/inorganic/group4/properties.html (6 of 12)30/12/2004 11:12:11

The trend from non-metal to metal in Group 4

So what happens to electronegativity in Group 4? Does it decrease as


you go down the Group, suggesting a trend towards metallic behaviour?

Well! It certainly falls from carbon to silicon, but from there on it is a


complete mess!
There therefore seems to be no relationship between the non-metal to
metal trend and the electronegativity values. Assuming the
electronegativity values are correct, I am completely at a loss to
understand this!

Note: The data in this chart again comes from the University
of Sheffield's Webelements site. Again, there is an awful lot
of variability in the data depending on where you get it from.
But in no case that I have found is there any trend to lower
electronegativities as you go down the Group. Older data
sources give a fall from carbon (2.5) to silicon (1.8), but then
give all the other elements in the Group the same value (all
1.8).
If anyone reading this has a simple explanation for the lack of
correlation between the trend to metallic behaviour and the
electronegativity values, could you please contact me via the
address on the about this site page.

http://www.chemguide.co.uk/inorganic/group4/properties.html (7 of 12)30/12/2004 11:12:11

The trend from non-metal to metal in Group 4

Ionisation energies
If you are thinking about the formation of positive ions, the obvious place
to start looking is how ionisation energies change as you go down Group
4.
Ionisation energies are defined as the energy needed to carry out each
of the following changes. They are quoted in kJ mol-1.
First ionisation energy:

Second ionisation energy:

. . . and so on.

Note: If you aren't sure about ionisation energies it would


pay you to follow this link before you go any further.
Use the BACK button on your browser to return quickly to this
page.

http://www.chemguide.co.uk/inorganic/group4/properties.html (8 of 12)30/12/2004 11:12:11

The trend from non-metal to metal in Group 4

None of the Group 4 elements form 1+ ions, so looking at the first


ionisation energy alone isn't very helpful. Some of the elements do,
however, form 2+ and (to some extent) 4+ ions.
The first chart shows how the total ionisation energy needed to form the 2
+ ions varies as you go down the Group. The values are all in kJ mol-1.

You can see that the ionisation energies tend to fall as you go down the
Group - although there is a slight increase at lead. The main trend is
because:

The atoms are getting bigger because of the extra layers of


electrons. The further the outer electrons are from the nucleus, the
less they are attracted - and so the easier they are to remove.
The outer electrons are screened from the full effect of the nucleus
by the increasing number of inner electrons.
These two effects outweigh the effect of the increasing nuclear
charge.

Note: The reason for the oddity at lead is discussed in some


more detail on a page about the oxidation states shown by
elements in Group 4. It isn't particularly important to the
present discussion.

http://www.chemguide.co.uk/inorganic/group4/properties.html (9 of 12)30/12/2004 11:12:11

The trend from non-metal to metal in Group 4

If you look at the amount of ionisation energy needed to form 4+ ions, the
pattern is similar, but not entirely clear cut. Again, the values are all in kJ
mol-1.

Note: The increase in total ionisation energy at lead is even


more obvious in the case of the possible formation of 4+ ions.
This is important when it comes to looking at the preferred
oxidation states of lead.

http://www.chemguide.co.uk/inorganic/group4/properties.html (10 of 12)30/12/2004 11:12:11

The trend from non-metal to metal in Group 4

What is clear looking at these two charts is that you have to put in large
amounts of ionisation energy to form 2+ ions, and huge amounts to form
4+ ions.
However, in each case there is a fall in ionisation energy as you go down
the Group which makes it more likely that tin and lead could form positive
ions - however, there is no indication from these figures that they are
likely to form positive ions.
The ionisation energies of carbon at the top of the Group are so huge
that there is no possibility of it forming simple positive ions.

Note: Even for tin and lead, you have to put in huge
amounts of energy to form either 2+ or 4+ ions. So why do
they form ions at all?
You have to remember that there are lots of other energy
terms involved in the formation of an ionic compound apart
from ionisation energy. Some of these give out large amounts
of energy - for example, lattice enthalpy if you are forming an
ionic solid, or hydration enthalpy if you are forming a solution.
You will need to read about Born-Haber cycles in order to
understand this fully. These aren't covered formally on this
site, but you will find them discussed in detail in my chemistry
calculations book.

Where would you like to go now?


To the Group 4 menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

http://www.chemguide.co.uk/inorganic/group4/properties.html (11 of 12)30/12/2004 11:12:11

The trend from non-metal to metal in Group 4

Jim Clark 2004

http://www.chemguide.co.uk/inorganic/group4/properties.html (12 of 12)30/12/2004 11:12:11

giant covalent structures

GIANT COVALENT STRUCTURES

This page decribes the structures of giant covalent substances like


diamond, graphite and silicon dioxide (silicon(IV) oxide), and relates
those structures to the physical properties of the substances.

The structure of diamond


The giant covalent structure of diamond
Carbon has an electronic arrangement of 2,4. In diamond, each carbon
shares electrons with four other carbon atoms - forming four single bonds.

In the diagram some carbon atoms only seem to be forming two bonds
(or even one bond), but that's not really the case. We are only showing a
small bit of the whole structure.
This is a giant covalent structure - it continues on and on in three
dimensions. It is not a molecule, because the number of atoms joined up
in a real diamond is completely variable - depending on the size of the
crystal.

http://www.chemguide.co.uk/atoms/structures/giantcov.html (1 of 8)30/12/2004 11:12:21

giant covalent structures

Note: We quoted the electronic structure of carbon as 2,4.


That simple view is perfectly adequate to explain the bonding
in diamond. If you are interested in a more modern view, you
could read the page on bonding in methane and ethane in
the organic section of this site. In the case of diamond, each
carbon is bonded to 4 other carbons rather than hydrogens,
but that makes no essential difference.

How to draw the structure of diamond


Don't try to be too clever by trying to draw too much of the structure!
Learn to draw the diagram given above. Do it in the following stages:

Practise until you can do a reasonable free-hand sketch in about 30


seconds.

The physical properties of diamond


Diamond

has a very high melting point (almost 4000C). Very strong carboncarbon covalent bonds have to be broken throughout the structure
before melting occurs.
is very hard. This is again due to the need to break very strong
covalent bonds operating in 3-dimensions.
doesn't conduct electricity. All the electrons are held tightly
between the atoms, and aren't free to move.
is insoluble in water and organic solvents. There are no possible

http://www.chemguide.co.uk/atoms/structures/giantcov.html (2 of 8)30/12/2004 11:12:21

giant covalent structures

attractions which could occur between solvent molecules and


carbon atoms which could outweigh the attractions between the
covalently bound carbon atoms.

The structure of graphite


The giant covalent structure of graphite
Graphite has a layer structure which is quite difficult to draw
convincingly in three dimensions. The diagram below shows the
arrangement of the atoms in each layer, and the way the layers are
spaced.

Notice that you can't really draw the side view of the layers to the same
scale as the atoms in the layer without one or other part of the diagram
being either very spread out or very squashed.
In that case, it is important to give some idea of the distances involved.
The distance between the layers is about 2.5 times the distance between
the atoms within each layer.
The layers, of course, extend over huge numbers of atoms - not just the
few shown above.

http://www.chemguide.co.uk/atoms/structures/giantcov.html (3 of 8)30/12/2004 11:12:21

giant covalent structures

You might argue that carbon has to form 4 bonds because of its 4
unpaired electrons, whereas in this diagram it only seems to be forming 3
bonds to the neighbouring carbons. This diagram is something of a
simplification, and shows the arrangement of atoms rather than the
bonding.

The bonding in graphite


Each carbon atom uses three of its electrons to form simple bonds to its
three close neighbours. That leaves a fourth electron in the bonding
level. These "spare" electrons in each carbon atom become delocalised
over the whole of the sheet of atoms in one layer. They are no longer
associated directly with any particular atom or pair of atoms, but are free
to wander throughout the whole sheet.

If you are interested (beyond A'level): The bonding in


graphite is like a vastly extended version of the bonding in
benzene. Each carbon atom undergoes sp2 hybridisation,
and then the unhybridised p orbitals on each carbon atom
overlap sideways to give a massive pi system above and
below the plane of the sheet of atoms.

The important thing is that the delocalised electrons are free to move
anywhere within the sheet - each electron is no longer fixed to a
particular carbon atom. There is, however, no direct contact between the
delocalised electrons in one sheet and those in the neighbouring sheets.
The atoms within a sheet are held together by strong covalent bonds stronger, in fact, than in diamond because of the additional bonding
caused by the delocalised electrons. So what holds the sheets together?
In graphite you have the ultimate example of van der Waals dispersion
forces. As the delocalised electrons move around in the sheet, very large
temporary dipoles can be set up which will induce opposite dipoles in the
sheets above and below - and so on throughout the whole graphite
crystal.

http://www.chemguide.co.uk/atoms/structures/giantcov.html (4 of 8)30/12/2004 11:12:21

giant covalent structures

Note: If you aren't sure about van der Waals forces follow
this link before you go on. Use the BACK button on your
browser to return to this page.

The physical properties of graphite


Graphite

has a high melting point, similar to that of diamond. In order to melt


graphite, it isn't enough to loosen one sheet from another. You
have to break the covalent bonding throughout the whole structure.
has a soft, slippery feel, and is used in pencils and as a dry
lubricant for things like locks. You can think of graphite rather like
a pack of cards - each card is strong, but the cards will slide over
each other, or even fall off the pack altogether. When you use a
pencil, sheets are rubbed off and stick to the paper.
has a lower density than diamond. This is because of the relatively
large amount of space that is "wasted" between the sheets.
is insoluble in water and organic solvents - for the same reason
that diamond is insoluble. Attractions between solvent molecules
and carbon atoms will never be strong enough to overcome the
strong covalent bonds in graphite.
conducts electricity. The delocalised electrons are free to move
throughout the sheets. If a piece of graphite is connected into a
circuit, electrons can fall off one end of the sheet and be replaced
with new ones at the other end.

http://www.chemguide.co.uk/atoms/structures/giantcov.html (5 of 8)30/12/2004 11:12:21

giant covalent structures

Note: The logic of this is that a piece of graphite ought only


to conduct electricity in 2-dimensions because electrons can
only move around in the sheets - and not from one sheet to
its neighbours.
In practice, a real piece of graphite isn't a perfect crystal, but
a host of small crystals stuck together at all sorts of angles.
Electrons will be able to find a route through the large piece
of graphite in all directions by moving from one small crystal
to the next.

The structure of silicon dioxide, SiO2


Silicon dioxide is also known as silicon(IV) oxide.
The giant covalent structure of silicon dioxide
There are three different crystal forms of silicon dioxide. The easiest one
to remember and draw is based on the diamond structure.
Crystalline silicon has the same structure as diamond. To turn it into
silicon dioxide, all you need to do is to modify the silicon structure by
including some oxygen atoms.

http://www.chemguide.co.uk/atoms/structures/giantcov.html (6 of 8)30/12/2004 11:12:21

giant covalent structures

Notice that each silicon atom is bridged to its neighbours by an oxygen


atom. Don't forget that this is just a tiny part of a giant structure extending
on all 3 dimensions.

Note: If you want to be fussy, the Si-O-Si bond angles are


wrong in this diagram. In reality the "bridge" from one silicon
atom to its neighbour isn't in a straight line, but via a "V"
shape (similar to the shape around the oxygen atom in a
water molecule). It's extremely difficult to draw that
convincingly and tidily in a diagram involving this number of
atoms. The simplification is perfectly acceptable.

The physical properties of silicon dioxide


Silicon dioxide

has a high melting point - varying depending on what the particular


structure is (remember that the structure given is only one of three
possible structures), but around 1700C. Very strong siliconoxygen covalent bonds have to be broken throughout the structure
before melting occurs.
is hard. This is due to the need to break the very strong covalent
bonds.
doesn't conduct electricity. There aren't any delocalised electrons.
All the electrons are held tightly between the atoms, and aren't
free to move.
is insoluble in water and organic solvents. There are no possible
attractions which could occur between solvent molecules and the
silicon or oxygen atoms which could overcome the covalent bonds
in the giant structure.

Where would you like to go now?

http://www.chemguide.co.uk/atoms/structures/giantcov.html (7 of 8)30/12/2004 11:12:21

giant covalent structures

To the structures menu . . .


To the atomic structure and bonding menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/atoms/structures/giantcov.html (8 of 8)30/12/2004 11:12:21

metal structures

METALLIC STRUCTURES

This page decribes the structure of metals, and relates that structure to
the physical properties of the metal.

The structure of metals


The arrangement of the atoms
Metals are giant structures of atoms held together by metallic bonds.
"Giant" implies that large but variable numbers of atoms are involved depending on the size of the bit of metal.

Note: Before you go on, it might be a good idea to read the


page on bonding in metals unless you are reasonably happy
about the idea of the delocalised electrons ("sea of
electrons") in metals.

12-co-ordination
Most metals are close packed - that is, they fit as many atoms as
possible into the available volume. Each atom in the structure has 12
touching neighbours. Such a metal is described as 12-co-ordinated.
Each atom has 6 other atoms touching it in each layer.

There are also 3 atoms touching any particular atom in the layer above
and another 3 in the layer underneath.
http://www.chemguide.co.uk/atoms/structures/metals.html (1 of 6)30/12/2004 11:12:26

metal structures

This second diagram shows the layer immediately above the first layer.
There will be a corresponding layer underneath. (There are actually two
different ways of placing the third layer in a close packed structure, but
that goes beyond the requirements of current A'level syllabuses.)
8-co-ordination
Some metals (notably those in Group 1 of the Periodic Table) are packed
less efficiently, having only 8 touching neighbours. These are 8-coordinated.

The left hand diagram shows that no atoms are touching each other
within a particular layer . They are only touched by the atoms in the
layers above and below. The right hand diagram shows the 8 atoms (4
above and 4 below) touching the darker coloured one.
Dislocations
It would be misleading to suppose that all the atoms in a piece of metal
are arranged in a regular way. Any piece of metal is made up of a large
number of "crystal grains", which are regions of perfect regularity. At the
grain boundaries atoms have become misaligned.

http://www.chemguide.co.uk/atoms/structures/metals.html (2 of 6)30/12/2004 11:12:26

metal structures

The grain boundaries are also known as dislocations.

The physical properties of metals


Melting points and boiling points
Metals tend to have high melting and boiling points because of the
strength of the metallic bond. The strength of the bond varies from metal
to metal and depends on the number of electrons which each atom
delocalises into the sea of electrons, and on the packing.
Group 1 metals like sodium and potassium have relatively low melting
and boiling points mainly because each atom only has one electron to
contribute to the bond - but there are other problems as well:

Group 1 elements are also inefficiently packed (8-co-ordinated), so


that they aren't forming as many bonds as most metals.
They have relatively large atoms (meaning that the nuclei are
some distance from the delocalised electrons) which also weakens
the bond.

Electrical conductivity
Metals conduct electricity. The delocalised electrons are free to move
throughout the structure in 3-dimensions. They can cross grain
boundaries. Even though the pattern may be disrupted at the boundary,
as long as atoms are touching each other, the metallic bond is still
present.
Liquid metals also conduct electricity, showing that although the metal
atoms may be free to move, the delocalisation remains in force until the
metal boils.

http://www.chemguide.co.uk/atoms/structures/metals.html (3 of 6)30/12/2004 11:12:26

metal structures

Thermal conductivity
Metals are good conductors of heat. Heat energy is picked up by the
electrons as additional kinetic energy (it makes them move faster). The
energy is transferred throughout the rest of the metal by the moving
electrons.

Strength and workability


Malleability and ductility
Metals are described as malleable (can be beaten into sheets) and
ductile (can be pulled out into wires). This is because of the ability of the
atoms to roll over each other into new positions without breaking the
metallic bond.
If a small stress is put onto the metal, the layers of atoms will start to roll
over each other. If the stress is released again, they will fall back to their
original positions. Under these circumstances, the metal is said to be
elastic.

If a larger stress is put on, the atoms roll over each other into a new
position, and the metal is permanently changed.

The hardness of metals

http://www.chemguide.co.uk/atoms/structures/metals.html (4 of 6)30/12/2004 11:12:26

metal structures

This rolling of layers of atoms over each other is hindered by grain


boundaries because the rows of atoms don't line up properly. It follows
that the more grain boundaries there are (the smaller the individual
crystal grains), the harder the metal becomes.
Offsetting this, because the grain boundaries are areas where the atoms
aren't in such good contact with each other, metals tend to fracture at
grain boundaries. Increasing the number of grain boundaries not only
makes the metal harder, but also makes it more brittle.
Controlling the size of the crystal grains
If you have a pure piece of metal, you can control the size of the grains
by heat treatment or by working the metal.
Heating a metal tends to shake the atoms into a more regular
arrangement - decreasing the number of grain boundaries, and so
making the metal softer. Banging the metal around when it is cold tends
to produce lots of small grains. Cold working therefore makes a metal
harder. To restore its workability, you would need to reheat it.
You can also break up the regular arrangement of the atoms by inserting
atoms of a slightly different size into the structure. Alloys such as brass
(a mixture of copper and zinc) are harder than the original metals
because the irregularity in the structure helps to stop rows of atoms from
slipping over each other.

Where would you like to go now?


To the structures menu . . .
To the atomic structure and bonding menu . . .
To Main Menu . . .

http://www.chemguide.co.uk/atoms/structures/metals.html (5 of 6)30/12/2004 11:12:26

metal structures

Jim Clark 2000

http://www.chemguide.co.uk/atoms/structures/metals.html (6 of 6)30/12/2004 11:12:26

metallic bonding

METALLIC BONDING

This page introduces the bonding in metals. It explains how the metallic
bond arises and why its strength varies from metal to metal.

What is a metallic bond?


Metallic bonding in sodium
Metals tend to have high melting points and boiling points suggesting
strong bonds between the atoms. Even a metal like sodium (melting
point 97.8C) melts at a considerably higher temperature than the
element (neon) which precedes it in the Periodic Table.
Sodium has the electronic structure 1s22s22p63s1. When sodium atoms
come together, the electron in the 3s atomic orbital of one sodium atom
shares space with the corresponding electron on a neighbouring atom to
form a molecular orbital - in much the same sort of way that a covalent
bond is formed.
The difference, however, is that each sodium atom is being touched by
eight other sodium atoms - and the sharing occurs between the central
atom and the 3s orbitals on all of the eight other atoms. And each of
these eight is in turn being touched by eight sodium atoms, which in turn
are touched by eight atoms - and so on and so on, until you have taken
in all the atoms in that lump of sodium.
All of the 3s orbitals on all of the atoms overlap to give a vast number of
molecular orbitals which extend over the whole piece of metal. There
have to be huge numbers of molecular orbitals, of course, because any
orbital can only hold two electrons.
The electrons can move freely within these molecular orbitals, and so
each electron becomes detached from its parent atom. The electrons are
said to be delocalised. The metal is held together by the strong forces of
attraction between the positive nuclei and the delocalised electrons.

http://www.chemguide.co.uk/atoms/bonding/metallic.html (1 of 3)30/12/2004 11:12:31

metallic bonding

This is sometimes described as "an array of positive ions in a sea of


electrons".
If you are going to use this view, beware! Is a metal made up of atoms or
ions? It is made of atoms.
Each positive centre in the diagram represents all the rest of the atom
apart from the outer electron, but that electron hasn't been lost - it may
no longer have an attachment to a particular atom, but it's still there in
the structure. Sodium metal is therefore written as Na - not Na+.
Metallic bonding in magnesium
If you work through the same argument with magnesium, you end up
with stronger bonds and so a higher melting point.
Magnesium has the outer electronic structure 3s2. Both of these
electrons become delocalised, so the "sea" has twice the electron
density as it does in sodium. The remaining "ions" also have twice the
charge (if you are going to use this particular view of the metal bond) and
so there will be more attraction between "ions" and "sea".
More realistically, each magnesium atom has one more proton in the
nucleus than a sodium atom has, and so not only will there be a greater
number of delocalised electrons, but there will also be a greater
attraction for them.
Magnesium atoms have a slightly smaller radius than sodium atoms, and
so the delocalised electrons are closer to the nuclei. Each magnesium
atom also has twelve near neighbours rather than sodium's eight. Both of
these factors increase the strength of the bond still further.

http://www.chemguide.co.uk/atoms/bonding/metallic.html (2 of 3)30/12/2004 11:12:31

metallic bonding

Metallic bonding in transition elements


Transition metals tend to have particularly high melting points and boiling
points. The reason is that they can involve the 3d electrons in the
delocalisation as well as the 4s. The more electrons you can involve, the
stronger the attractions tend to be.

Note: If you aren't happy about the electronic structure of


transition metals, then you might like to follow this link to
revise it.

The metallic bond in molten metals


In a molten metal, the metallic bond is still present, although the ordered
structure has been broken down. The metallic bond isn't fully broken until
the metal boils. That means that boiling point is actually a better guide to
the strength of the metallic bond than melting point is. On melting, the
bond is loosened, not broken.

Where would you like to go now?


To explore the structure of metals . . .
To the bonding menu . . .
To the atomic structure and bonding menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/atoms/bonding/metallic.html (3 of 3)30/12/2004 11:12:31

deciding bond type from physical properties

DECIDING WHAT TYPE OF STRUCTURE A


SUBSTANCE HAS

This page explains how you can decide what sort of structure a
substance has by looking at its physical properties. It starts with a very
simple look at solids, liquids and gases.

Changes of state as a guide to forces between particles


The arrangements of particles in solids, liquids and gases
A simple view of the arrangement of the particles in solids, liquids and
gases looks like this:

Solids
In the solid, the particles are touching, and the only motion allowed to
them is vibration. The particles may be arranged regularly (in which case,
the solid is crystalline), or at random (giving waxy solids like candles or
some forms of polythene, for example).
The particles are held in the solid by forces which depend on the actual
substance - ionic bonds, covalent bonds, metallic bonds, hydrogen
bonds or van der Waals attractions.
Melting and freezing
If energy is supplied by heating the solid, the heat energy causes greater
http://www.chemguide.co.uk/atoms/structures/whichsort.html (1 of 5)30/12/2004 11:12:35

deciding bond type from physical properties

vibrations until the particles eventually loosen from each other to form a
liquid. The heat energy required to convert 1 mole of solid into a liquid at
its melting point is called the enthalpy of fusion.
When a liquid freezes, the reverse happens. At some temperature, the
motion of the particles is slow enough for the forces of attraction to be
able to hold the particles as a solid. As the new bonds are formed, heat
energy is evolved.
Liquids
In a liquid, the particles are mainly touching, but some gaps have
appeared in the structure. These gaps allow the particles to move, and
so the particles are arranged randomly. Unless melting has broken up a
substance consisting only of covalent bonds (a giant covalent structure),
the forces that held the solid particles together are also present in the
liquid, but in a somewhat loosened form.
Boiling and condensing
If more heat energy is supplied, the particles eventually move fast
enough to break all the attractions between them, and the liquid boils.
The heat energy required to convert 1 mole of liquid into a gas at its
boiling point is called the enthalpy of vaporisation.
If the gas is cooled, at some temperature the gas particles will slow down
enough for the attractions to become effective enough to condense it
back into a liquid. Again, as those forces are re-established, heat energy
is released.

Remember: Breaking bonds needs energy; making bonds


releases it.

http://www.chemguide.co.uk/atoms/structures/whichsort.html (2 of 5)30/12/2004 11:12:35

deciding bond type from physical properties

Gases
In a gas, the particles are entirely free to move. At ordinary pressures,
the distance between individual particles is of the order of ten times the
diameter of the particles. At that distance, any attractions between the
particles are fairly negligible.

Deducing the type of bonding from physical properties


The physical state and other properties
The best place to start is usually the physical state.
Melting point isn't always a good guide to the size of the attractions
between particles, because the attractive forces have only been
loosened on melting - not broken entirely. Boiling point is a much better
guide, because enough heat has now been supplied to break the
attractive forces completely. The stronger the attractions, the higher the
boiling point.
That being said, melting points are often used to judge the size of
attractive forces between particles in solids, but you will find the
occasional oddity. Those oddities usually disappear if you consider
boiling points instead.

As an example: You would expect stronger metallic bonding


in aluminium than in magnesium, because aluminium has 3
electrons to delocalise into the "sea of electrons" rather than
magnesium's 2. The boiling points reflect this: Al 2470C, Mg
1110C. However, aluminium's melting point is only 10C
higher than magnesium's: Al 660C, Mg 650C. (I've never
found a good explanation for this!)
If you need some more background on metallic bonding, you
could follow this link.

http://www.chemguide.co.uk/atoms/structures/whichsort.html (3 of 5)30/12/2004 11:12:35

deciding bond type from physical properties

So . . . If it is a gas, liquid or low melting point solid, it will consist of


covalently bound molecules (except the noble gases which have
molecules consisting of single atoms).
The size of the melting point or boiling point gives a guide to the strength
of the intermolecular forces. If it is also soluble in water (without
reacting), that suggests a small molecule capable of hydrogen bonding or, at least, a small very polar molecule.
If it is a high melting point solid, it will be a giant structure - either ionic,
metallic or giant covalent.
Solubility in water (without reaction) suggests it's ionic. If the substance
also undergoes electrolysis when it is molten, that would confirm that it
was ionic.

Note: Electrolysis is the splitting up of a compound using


electricity. For example, molten sodium chloride conducts
electricity and is split into sodium and chlorine in the process.

Conductivity of electricity in the solid state suggests delocalised


electrons, and therefore either a metal or graphite. The clue as to which
you had would usually come from other data - appearance, malleability,
etc.
Note: Semi-conductors like silicon - a giant covalent
structure with the same arrangement of atoms as diamond also conduct electricity. The theory of semi-conductors is
beyond the A'level syllabuses.

http://www.chemguide.co.uk/atoms/structures/whichsort.html (4 of 5)30/12/2004 11:12:35

deciding bond type from physical properties

Where would you like to go now?


To the structures menu . . .
To the atomic structure and bonding menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/atoms/structures/whichsort.html (5 of 5)30/12/2004 11:12:35

structures and physical properties of period 3 elements

THE STRUCTURES OF THE PERIOD 3


ELEMENTS

This page describes the structures of the Period 3 elements from sodium
to argon, and shows how these structures can be used to explain the
physical properties of the elements.

Variation in physical properties in period 3


Melting and boiling points

In a moment we shall explain all the ups and downs in this graph.

Electrical conductivity
Sodium, magnesium and aluminium are all good conductors of electricity.
Silicon is a semiconductor. None of the rest conduct electricity.

Explaining the trends

http://www.chemguide.co.uk/atoms/structures/period3.html (1 of 5)30/12/2004 11:12:38

structures and physical properties of period 3 elements

Warning! To understand this section you must be familiar


with metallic bonding and the structure of metals, giant
covalent structures, simple molecular structures and van der
Waals forces. If you are uncertain of any of this, now is the
time to revise it.

Three metallic structures


Sodium, magnesium and aluminium all have metallic structures, which
accounts for their electrical conductivity and relatively high melting and
boiling points.
Melting and boiling points rise across the three metals because of the
increasing number of electrons which each atom can contribute to the
delocalised "sea of electrons". The atoms also get smaller and have
more protons as you go from sodium to magnesium to aluminium.
The attractions and therefore the melting and boiling points increase
because:

The nuclei of the atoms are getting more positively charged.

The sea is getting more negatively charged.

The sea is getting progressively nearer to the nuclei and so more


strongly attracted.

Silicon - a giant covalent structure

http://www.chemguide.co.uk/atoms/structures/period3.html (2 of 5)30/12/2004 11:12:38

structures and physical properties of period 3 elements

Silicon is a non-metal, and has a giant covalent structure exactly the


same as carbon in diamond - hence the high melting point. You have to
break strong covalent bonds in order to melt it.
There are no obviously free electrons in the structure, and although it
conducts electricity, it doesn't do so in the same way as metals. Silicon is
a semiconductor.

Note: Explaining how semiconductors conduct electricity is


beyond the scope of A'level chemistry syllabuses.

Four molecular elements


Phosphorus, sulphur, chlorine and argon are simple molecular
substances with only van der Waals attractions between the molecules.
Their melting or boiling points will be lower than those of the first four
members of the period which have giant structures. The presence of
individual molecules prevents any possibility of electrons flowing, and so
none of them conduct electricity.
The sizes of the melting and boiling points are governed entirely by the
sizes of the molecules:

http://www.chemguide.co.uk/atoms/structures/period3.html (3 of 5)30/12/2004 11:12:38

structures and physical properties of period 3 elements

Argon molecules consist of single argon atoms.


Phosphorus
There are several forms of phosphorus. The data in the graph at the top
of the page applies to white phosphorus which contains P4 molecules. To
melt phosphorus you don't have to break any covalent bonds - just the
much weaker van der Waals forces between the molecules.
Sulphur
Sulphur consists of S8 rings of atoms. The molecules are bigger than
phosphorus molecules, and so the van der Waals attractions will be
stronger, leading to a higher melting and boiling point.
Chlorine
Chlorine, Cl2, is a much smaller molecule with comparatively weak van
der Waals attractions, and so chlorine will have a lower melting and
boiling point than sulphur or phosphorus.
Argon
Argon molecules are just single argon atoms, Ar. The scope for van der
Waals attractions between these is very limited and so the melting and
boiling points of argon are lower again.

http://www.chemguide.co.uk/atoms/structures/period3.html (4 of 5)30/12/2004 11:12:38

structures and physical properties of period 3 elements

Note: You might also be interested in the trends in ionisation


energy, atomic radius and electronegativity in this period.
You will find relevant descriptions and explanations if you
follow these links - or they are available via the menus below.

Where would you like to go now?


To the structures menu . . .
To the atomic structure and bonding menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/atoms/structures/period3.html (5 of 5)30/12/2004 11:12:38

Oxidation state trends in Group 4

OXIDATION STATE TRENDS IN GROUP 4

This page explores the oxidation states (oxidation numbers) shown by


the Group 4 elements - carbon (C), silicon (Si), germanium (Ge), tin (Sn)
and lead (Pb). It looks at the increasing tendency of the elements to form
compounds in which their oxidation states are +2, particularly with
reference to tin and lead.

Note: If you aren't happy about oxidation and reduction


(including the use of oxidation states), it is essential to follow
this link before you go any further.
Use the BACK button on your browser to return quickly to this
page.

Some examples of the trends in oxidation states


The overall trend
The typical oxidation state shown by elements in Group 4 is +4, found in
compounds like CCl4, SiCl4 and SnO2.

Warning: Don't fall into the trap of quoting CH4 as an


example of carbon with a typical oxidation state of +4.
Because carbon is more electronegative than hydrogen, its
oxidation state in this instance is -4!

http://www.chemguide.co.uk/inorganic/group4/oxstates.html (1 of 9)30/12/2004 11:12:46

Oxidation state trends in Group 4

However, as you go down the Group, there are more and more examples
where the oxidation state is +2, such as SnCl2, PbO, and Pb2+.
With tin, the +4 state is still more stable than the +2, but by the time you
get to lead, the +2 state is the more stable - and dominates the chemistry
of lead.

An example from carbon chemistry


The only common example of the +2 oxidation state in carbon chemistry
occurs in carbon monoxide, CO. Carbon monoxide is a strong reducing
agent because it is easily oxidised to carbon dioxide - where the
oxidation state is the more thermodynamically stable +4.
For example, carbon monoxide reduces many hot metal oxides to the
metal - a reaction which is used, for example, in the extraction of iron in a
blast furnace.

Examples from tin chemistry


By the time you get down the Group as far as tin, the +2 state has
become increasingly common, and there is a good range of both tin(II)
and tin(IV) compounds. However, tin(IV) is still the more stable oxidation
state of tin.
That means that it will be fairly easy to convert tin(II) compounds into tin
(IV) compounds. This is best shown in the fact that Sn2+ ions in solution
are good reducing agents.
For example, a solution containing tin(II) ions (for example, tin(II) chloride
solution) will reduce a solution of iodine to iodide ions. In the process, the
tin(II) ions are oxidised to tin(IV) ions.
http://www.chemguide.co.uk/inorganic/group4/oxstates.html (2 of 9)30/12/2004 11:12:46

Oxidation state trends in Group 4

Note: For simplicity, I am writing this equation (and the next


few) as if the product contained simple tin(IV) ions. In fact,
simple tin(IV) ions don't exist in solution. In these examples,
they will usually be a part of a much larger complex ion. Don't
worry about this at this level.

Tin(II) ions also reduce iron(III) ions to iron(II) ions. For example, tin(II)
chloride solution will reduce iron(III) chloride solution to iron(II) chloride
solution. In the process, the tin(II) ions are oxidised to the more stable tin
(IV) ions.

Tin(II) ions will also, of course, be easily oxidised by powerful oxidising


agents like acidified potassium manganate(VII) solution (potassium
permanganate solution). This reaction could be used as a titration to find
the concentration of tin(II) ions in a solution.

http://www.chemguide.co.uk/inorganic/group4/oxstates.html (3 of 9)30/12/2004 11:12:46

Oxidation state trends in Group 4

Note: If you aren't happy about titration calculations


(including those involving potassium manganate(VII) ), you
might be interested in my chemistry calculations book.

And as a final example . . .


In organic chemistry, tin and concentrated hydrochloric acid are
traditionally used to reduce nitrobenzene to phenylamine (aniline). This
reaction involves the tin first being oxidised to tin(II) ions and then further
to the preferred tin(IV) ions.

Note: This reaction is dealt with in some detail in the organic


chemistry section of the site on a page about the preparation
of phenylamine.
Use the BACK button on your browser to return to this page if
you choose to follow this link.

Examples from lead chemistry


With lead, the situation is reversed. This time, the lead(II) oxidation state
is the more stable, and there is a strong tendency for lead(IV)
compounds to react to give lead(II) compounds.
Lead(IV) chloride, for example, decomposes at room temperature to give
lead(II) chloride and chlorine gas:

. . . and lead(IV) oxide decomposes on heating to give lead(II) oxide and


oxygen.

http://www.chemguide.co.uk/inorganic/group4/oxstates.html (4 of 9)30/12/2004 11:12:46

Oxidation state trends in Group 4

Lead(IV) oxide also reacts with concentrated hydrochloric acid, oxidising


some of the chloride ions in the acid to chlorine gas. Once again, the
lead is reduced from the +4 to the more stable +2 state.

Trying to explain the trends in oxidation states


There's nothing surprising about the normal Group oxidation state of +4.
All of the elements in the group have the outer electronic structure
ns2npx1npy1, where n varies from 2 (for carbon) to 6 (for lead). The
oxidation state of +4 is where all these outer electrons are directly
involved in the bonding.
As you get closer to the bottom of the Group, there is an increasing
tendency for the s2 pair not to be used in the bonding. This is often
known as the inert pair effect - and is dominant in lead chemistry.
However, just giving it a name like "inert pair effect" explains nothing.
You need to look at two different explanations depending on whether you
are talking about the formation of ionic or covalent bonds.

http://www.chemguide.co.uk/inorganic/group4/oxstates.html (5 of 9)30/12/2004 11:12:46

Oxidation state trends in Group 4

Note: What follows is quite likely to be way beyond what you


need for UK A level purposes - and is there mainly for
interest. To be sure, refer to your syllabus and, more
importantly, to past exam papers and mark schemes. If you
are a UK A level chemistry student and haven't got these,
follow this link to the syllabuses page to find out how to get
hold of them.

The inert pair effect in the formation of ionic bonds


If the elements in Group 4 form 2+ ions, they will lose the p electrons,
leaving the s2 pair unused. For example, to form a lead(II) ion, lead will
lose the two 6p electrons, but the 6s electrons will be left unchanged - an
"inert pair".
You would normally expect ionisation energies to fall as you go down a
Group as the electrons get further from the nucleus. That doesn't quite
happen in Group 4.
This first chart shows how the total ionisation energy needed to form the 2
+ ions varies as you go down the Group. The values are all in kJ mol-1.

Notice the slight increase between tin and lead.


This means that it is slightly more difficult to remove the p electrons from
lead than from tin.
However, if you look at the pattern for the loss of all four electrons, the
discrepancy between tin and lead is much more marked. The relatively
http://www.chemguide.co.uk/inorganic/group4/oxstates.html (6 of 9)30/12/2004 11:12:46

Oxidation state trends in Group 4

large increase between tin and lead must be because the 6s2 pair is
significantly more difficult to remove in lead than the corresponding 5s2
pair in tin.

Again, the values are all in kJ mol-1, and the two charts are to
approximately the same scale.
The reasons for all this lie in the Theory of Relativity. With the heavier
elements like lead, there is what is known as a relativistic contraction
of the electrons which tends to draw the electrons closer to the nucleus
than you would expect. Because they are closer to the nucleus, they are
more difficult to remove. The heavier the element, the greater this effect.
This affects s electrons much more than p electrons.
In the case of lead, the relativistic contraction makes it energetically more
difficult to remove the 6s electrons than you might expect. The energy
releasing terms when ions are formed (like lattice enthalpy or hydration
enthalpy) obviously aren't enough to compensate for this extra energy.

http://www.chemguide.co.uk/inorganic/group4/oxstates.html (7 of 9)30/12/2004 11:12:46

Oxidation state trends in Group 4

That means that it doesn't make energetic sense for lead to form 4+ ions.

Note: If you want to find out more about the relativistic


contraction, try a Google search on relativistic contraction
electrons - but expect to get involved in some heavy
reading! You will find a Google search box on the Main Menu
(link at bottom of this page). Don't forget to check the "search
www" button otherwise you will just end up back here again!

The inert pair effect in the formation of covalent bonds


You need to think about why carbon normally forms four covalent bonds
rather than two.
Using the electrons-in-boxes notation, the outer electronic structure of
carbon looks like this:

There are only two unpaired electrons. Before carbon forms bonds,
though, it normally promotes one of the s electrons to the empty p orbital.

That leaves 4 unpaired electrons which (after hybridisation) can go on to


http://www.chemguide.co.uk/inorganic/group4/oxstates.html (8 of 9)30/12/2004 11:12:46

Oxidation state trends in Group 4

form 4 covalent bonds.


It is worth supplying the energy to promote the s electron, because the
carbon can then form twice as many covalent bonds. Each covalent bond
that forms releases energy, and this is more than enough to supply the
energy needed for the promotion.
One possible explanation for the reluctance of lead to do the same thing
lies in falling bond energies as you go down the Group. Bond energies
tend to fall as atoms get bigger and the bonding pair is further from the
two nuclei and better screened from them.
For example, the energy released when two extra Pb-X bonds (where X
is H or Cl or whatever) are formed may no longer be enough to
compensate for the extra energy needed to promote a 6s electron into
the empty 6p orbital.
This would would be made worse, of course, if the energy gap between
the 6s and 6p orbitals was increased by the relativistic contraction of the
6s orbital.

Where would you like to go now?


To the Group 4 menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/inorganic/group4/oxstates.html (9 of 9)30/12/2004 11:12:46

The chlorides of carbon, silicon and lead

THE CHLORIDES OF CARBON, SILICON AND


LEAD

This page takes a brief look at the tetrachlorides of carbon, silicon and
lead, and also at lead(II) chloride. It looks at their structures, stability and
reactions with water.

Structures
Carbon, silicon and lead tetrachlorides
These all have the formula XCl4.
They are all simple covalent molecules with a typical tetrahedral shape.
All of them are liquids at room temperature. (Although at room
temperature, lead(IV) chloride will tend to decompose to give lead(II)
chloride and chlorine gas - see below.)
Lead(II) chloride, PbCl2
Lead(II) chloride is a white solid, melting at 501C. It is very slightly
soluble in cold water, but more soluble in hot water. You can think of lead
(II) chloride as being mainly ionic in character.

Stability
At the top of Group 4, the most stable oxidation state shown by the
elements is +4. This is the oxidation state shown by carbon and silicon in
CCl4 and SiCl4. These therefore have no tendency to split up to give
dichlorides.
However, the relative stability of the +4 oxidation state falls as you go
down the Group, and the +2 oxidation state becomes the most stable by
http://www.chemguide.co.uk/inorganic/group4/chlorides.html (1 of 5)30/12/2004 11:12:55

The chlorides of carbon, silicon and lead

the time you get to lead.


Lead(IV) chloride decomposes at room temperature to give the more
stable lead(II) chloride and chlorine gas.

Note: This oxidation state trend in Group 4 is dealt with in


more detail on another page in this section.
Use the BACK button on your browser to return to this page if
you choose to follow this link.

Reaction with water (hydrolysis)


Carbon tetrachloride (tetrachloromethane)
Carbon tetrachloride has no reaction with water. If you add it to water, it
simply forms a separate layer underneath the layer of water.
Suppose a water molecule is going to react with the carbon tetrachloride.
The reaction would have to start by the water molecule's oxygen
attaching itself to the carbon atom via the oxygen's lone pair. A chlorine
atom would get pushed off the carbon in the process.
There are two problems with this.
First, the chlorines are so bulky and the carbon atom so small, that the
oxygen can't easily get at the carbon atom.

http://www.chemguide.co.uk/inorganic/group4/chlorides.html (2 of 5)30/12/2004 11:12:55

The chlorides of carbon, silicon and lead

. . . and even if it did, there will be a stage where there is considerable


cluttering around that carbon atom before the chlorine atom breaks away
completely. There is going to be a lot of repulsion between the various
lone pairs on all the atoms surrounding the carbon.

That cluttering is going to make this half-way stage (properly called a


"transition state") very unstable. A very unstable transition state means a
very high activation energy for the reaction.
The other problem is that there isn't a convenient empty orbital on the
carbon that the oxygen lone pair can attach to.
If it could attach before the chlorine starts to break away, that would be
an advantage. Forming a bond releases energy, and that energy would
therefore be readily available for breaking a carbon-chlorine bond. But in
the case of a carbon atom, that isn't possible.

Silicon tetrachloride
The situation is different with silicon tetrachloride.
The silicon atom is bigger, and so there is more room around it for the
water molecule to attack, and the transition state will be less cluttered.
But silicon has the additional advantage that there are empty 3d orbitals
available to accept a lone pair from the water molecule. Carbon doesn't
have 2d orbitals because there are no such things. There are no empty 2-

http://www.chemguide.co.uk/inorganic/group4/chlorides.html (3 of 5)30/12/2004 11:12:55

The chlorides of carbon, silicon and lead

level orbitals available in the carbon case.


This means that the oxygen can bond to the silicon before the need to
break a silicon-chlorine bond. This makes the whole process
energetically easier.
So . . . silicon tetrachloride reacts violently with water to give white solid
silicon dioxide and steamy fumes of HCl.

Liquid SiCl4 fumes in moist air for this reason - it is reacting with water
vapour in the air.

Lead tetrachloride (lead(IV) chloride)


The reaction of lead(IV) chloride with water is just like the silicon
tetrachloride one. You will get lead(IV) oxide produced as a brown solid
and fumes of hydrogen chloride given off. (This will also, of course, be
confused by the decomposition of the lead(IV) chloride to give lead(II)
chloride and chlorine gas - see above.)

Lead(II) chloride
Unlike the tetrachlorides, lead(II) chloride can be thought of as ionic. It is
sparingly soluble in cold water, but more soluble in hot water. Looked at
simply, solubility in water involves the break-up of the ionic lattice and the
hydration of the lead(II) and chloride ions to give Pb2+(aq) and Cl-(aq).

Where would you like to go now?


To the Group 4 menu . . .

http://www.chemguide.co.uk/inorganic/group4/chlorides.html (4 of 5)30/12/2004 11:12:55

The chlorides of carbon, silicon and lead

To the Inorganic Chemistry menu . . .


To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/inorganic/group4/chlorides.html (5 of 5)30/12/2004 11:12:55

The oxides of carbon, silicon, germanium, tin and lead

THE OXIDES OF GROUP 4

This page takes a brief look at the oxides of carbon, silicon, germanium,
tin and lead. It concentrates on the structural differences between carbon
dioxide and silicon dioxide, and on the trends in acid-base behaviour of
the oxides as you go down Group 4.

The structures of carbon dioxide and silicon dioxide


There is an enormous difference between the physical properties of
carbon dioxide and silicon dioxide (also known as silicon(IV) oxide or
silica). Carbon dioxide is a gas whereas silicon dioxide is a hard highmelting solid. The other dioxides in Group 4 are also solids.
This obviously reflects a difference in structure between carbon dioxide
and the dioxides of the rest of the Group.
The structure of carbon dioxide
The fact that carbon dioxide is a gas means that it must consist of simple
molecules. Carbon can form simple molecules with oxygen because it
can form double bonds with the oxygen.

None of the other elements in Group 4 form double bonds with oxygen,
and so that forces completely different structures on them.

http://www.chemguide.co.uk/inorganic/group4/oxides.html (1 of 11)30/12/2004 11:13:13

The oxides of carbon, silicon, germanium, tin and lead

Note: The explanation for this is quite probably beyond what


you need for UK A level chemistry purposes, but I am
including it anyway. It isn't very difficult to understand and, to
be honest, there isn't anything else remotely interesting on
the rest of this page!
If you can follow it OK, well done! If not, skip over it to the
structure of silicon dioxide. If you haven't met the concept of
hybridisation, it would probably be better to miss it out - but
give it a try and see what happens.

When carbon forms bonds with oxygen, it first promotes one of the
electrons in the 2s level into the empty 2p level. This produces 4
unpaired electrons.

It now reshuffles those electrons slightly by hybridising the 2s electron


and one of the 2p electrons to make two sp1 hybrid orbitals of equal
energy. The other 2p electrons are left alone for the time being.

What these look like in the atom (using the same colour coding) is:
http://www.chemguide.co.uk/inorganic/group4/oxides.html (2 of 11)30/12/2004 11:13:13

The oxides of carbon, silicon, germanium, tin and lead

Notice that the two green lobes are two different hybrid orbitals arranged as far apart from each other as possible. Don't confuse them
with the shape of a p orbital.
So that's how the carbon is organised just before it bonds. Now we need
to look at the oxygen.
Oxygen's electronic structure is 1s22s22px22py12pz1.
Hybridisation occurs in the oxygen as well. This time, sp2 hybrids are
formed with the s orbital and two of the p orbitals being rearranged to
give 3 orbitals of equal energy - leaving a temporarily unaffected p orbital.

This time two of the sp2 hybrid orbitals contain lone pairs of electrons.

Now line up the two oxygens and the carbon prior to bonding them.
Notice that the left-hand oxygen has been rotated through 90:
http://www.chemguide.co.uk/inorganic/group4/oxides.html (3 of 11)30/12/2004 11:13:13

The oxides of carbon, silicon, germanium, tin and lead

Then bring them together so that the pale green hybrid orbitals overlap
end-to-end to form simple covalent bonds. These are properly called
sigma bonds, and are shown as orange in the next diagram.
This brings the various p orbitals close enough together that they overlap
sideways.

Sideways overlap between the two sets of p orbitals produces two pi


bonds - similar to the pi bond found in, say, ethene. These pi bonds are
twisted at 90 to each other in the final molecule.

http://www.chemguide.co.uk/inorganic/group4/oxides.html (4 of 11)30/12/2004 11:13:13

The oxides of carbon, silicon, germanium, tin and lead

So . . . in order to form a carbon-oxygen double bond, it is necessary for


the p orbitals on the carbon and the oxygen to overlap sideways.

The structure of silicon dioxide


Silicon doesn't double bond with oxygen.
Silicon atoms are bigger than carbon. That means that silicon-oxygen
bonds will be longer than carbon-oxygen bonds.
Imagine trying to make a silicon-oxygen double bond in the same way as
we did for a carbon-oxygen double bond. With the longer silicon-oxygen
bonds, the p orbitals on the silicon and the oxygen aren't quite close
enough together to allow enough sideways overlap to give a stable pi
bond.
So, silicon bonds with oxygen in such a way that only single bonds are
formed.
There are various different structures for silicon dioxide. The easiest to
remember and draw is:

This is based on a diamond structure with each of the silicon atoms


being bridged to its other four neighbours via an oxygen atom.

http://www.chemguide.co.uk/inorganic/group4/oxides.html (5 of 11)30/12/2004 11:13:13

The oxides of carbon, silicon, germanium, tin and lead

Note: If you want to be fussy, the Si-O-Si bond angles are


wrong in this diagram. In reality the "bridge" from one silicon
atom to its neighbour isn't in a straight line, but via a "V"
shape (similar to the shape around the oxygen atom in a
water molecule). It's extremely difficult to draw that
convincingly and tidily in a diagram involving this number of
atoms. The simplification is perfectly acceptable.

This means that silicon dioxide is a giant covalent structure. The strong
bonds in three dimensions make it a hard, high melting point solid.

Note: If you want a more detailed discussion of the silicon


dioxide structure (including a guide to drawing the diamond
structure!) and how it affects its physical properties, you will
find it on a page about giant covalent structures.
If you choose to follow this link, use the BACK button on your
browser to return to this page.

The acid-base behaviour of the Group 4 oxides


The oxides of the elements at the top of Group 4 are acidic, but acidity of
the oxides falls as you go down the Group. Towards the bottom of the
Group, the oxides become more basic - although without ever losing their
acidic character completely.
An oxide which can show both acidic and basic properties is said to be
amphoteric.
The trend is therefore from acidic oxides at the top of the Group towards
amphoteric ones at the bottom.

Carbon and silicon oxides


http://www.chemguide.co.uk/inorganic/group4/oxides.html (6 of 11)30/12/2004 11:13:13

The oxides of carbon, silicon, germanium, tin and lead

Carbon monoxide
Carbon monoxide is usually treated as if it was a neutral oxide, but in fact
it is very, very slightly acidic. It doesn't react with water, but it will react
with hot concentrated sodium hydroxide solution to give a solution of
sodium methanoate.

The fact that the carbon monoxide reacts with the basic hydroxide ion
shows that it must be acidic.

Carbon and silicon dioxides


These are both weakly acidic.
With water
Silicon dioxide doesn't react with water, because of the difficulty of
breaking up the giant covalent structure.
Carbon dioxide does react with water to a slight extent to produce
hydrogen ions (strictly, hydroxonium ions) and hydrogencarbonate ions.
Overall, this reaction is:

The solution of carbon dioxide in water is sometimes known as carbonic


acid, but in fact only about 0.1% of the carbon dioxide has actually
reacted. The position of equilibrium is well to the left-hand side.
With bases
Carbon dioxide reacts with sodium hydroxide solution in the cold to give
either sodium carbonate or sodium hydrogencarbonate solution depending on the reacting proportions.

http://www.chemguide.co.uk/inorganic/group4/oxides.html (7 of 11)30/12/2004 11:13:13

The oxides of carbon, silicon, germanium, tin and lead

Silicon dioxide also reacts with sodium hydroxide solution, but only if it is
hot and concentrated. Sodium silicate solution is formed.

You may also be familiar with one of the reactions happening in the Blast
Furnace extraction of iron - in which calcium oxide (from the limestone
which is one of the raw materials) reacts with silicon dioxide to produce a
liquid slag, calcium silicate. This is also an example of the acidic silicon
dioxide reacting with a base.

Gernamium, tin and lead oxides


The monoxides
All of these oxides are amphoteric - they show both basic and acidic
properties.
The basic nature of the oxides
These oxides all react with acids to form salts.
For example, they all react with concentrated hydrochloric acid. This can
be summarised as:

. . . where X can be Ge and Sn, but unfortunately needs modifying a bit


for lead.
Lead(II) chloride is fairly insoluble in water and, instead of getting a
solution, it would form an insoluble layer over the lead(II) oxide if you
were to use dilute hydrochloric acid - stopping the reaction from going on.

http://www.chemguide.co.uk/inorganic/group4/oxides.html (8 of 11)30/12/2004 11:13:13

The oxides of carbon, silicon, germanium, tin and lead

However, in this example we are talking about using concentrated


hydrochloric acid.
The large excess of chloride ions in the concentrated acid react with the
lead(II) chloride to produce soluble complexes such as PbCl42-. These
ionic complexes are soluble in water and so the problem disappears.

Unfortunately, it means that you have more to remember!

Note: There are almost certainly going to be similar


complexes formed in the germanium and tin cases in the
presence of excess concentrated hydrochloric acid, but
because they aren't important to the reaction happening, they
tend to be ignored at this level.

The acidic nature of the oxides


All of these oxides also react with bases like sodium hydroxide solution.
This time we can generalise without exception:

Lead(II) oxide, for example, would react to give PbO22- - plumbate(II)


ions.

Note: This reaction is written using a simplified version of


the formula of the product. This is adequate for this level.

http://www.chemguide.co.uk/inorganic/group4/oxides.html (9 of 11)30/12/2004 11:13:13

The oxides of carbon, silicon, germanium, tin and lead

The dioxides
These dioxides are again amphoteric - showing both basic and acidic
properties.
The basic nature of the dioxides
The dioxides react with concentrated hydrochloric acid first to give
compounds of the type XCl4:

These will react with excess chloride ions in the hydrochloric acid to give
complexes such as XCl62-.

In the case of lead(IV) oxide, the reaction has to be done with ice-cold
hydrochloric acid. If the reaction is done any warmer, the lead(IV)
chloride decomposes to give lead(II) chloride and chlorine gas. This is an
effect of the preferred oxidation state of lead being +2 rather than +4.

Note: You will find more about this (including an overall


equation for the reaction of lead(IV) oxide with concentrated
hydrochloric acid at ordinary temperatures) on a page about
the oxidation state trends in Group 4.
If you choose to follow this link, use the BACK button on your
browser to return to this page.

http://www.chemguide.co.uk/inorganic/group4/oxides.html (10 of 11)30/12/2004 11:13:13

The oxides of carbon, silicon, germanium, tin and lead

The acidic nature of the dioxides


The dioxides will react with hot concentrated sodium hydroxide solution
to give soluble complexes of the form [X(OH)6]2-.

Some sources suggest that the lead(IV) oxide needs molten sodium
hydroxide. In that case, the equation is different.

Note: If you are a UK student doing the Edexcel syllabus,


you might like to know that their chief examiner for A level
chemistry at the time of writing uses this second equation for
the lead(IV) oxide reaction on his website, and the first
equation for the other two dioxides.

Where would you like to go now?


To the Group 4 menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/inorganic/group4/oxides.html (11 of 11)30/12/2004 11:13:13

some insoluble lead(II) compounds

SOME INSOLUBLE LEAD(II) COMPOUNDS

This page looks at the formation of some insoluble lead(II) compounds


from aqueous lead(II) ions using precipitation reactions. It describes the
reactions to form lead(II) hydroxide, lead(II) chloride, lead(II) iodide and
lead(II) sulphate.
Because of the insolubility of so many lead(II) compounds, the usual
source of lead(II) ions in solution is lead(II) nitrate solution - and that will
be assumed in all the following examples.

Making lead(II) hydroxide


If a little sodium hydroxide solution is added to colourless lead(II) nitrate
solution, a white precipitate of lead(II) hydroxide is produced.

If more sodium hydroxide solution is added, the precipitate redissolves to


give a colourless solution which might be called sodium plumbate(II)
solution - but could be called by a lot of alternative names depending on
exactly how the formula is written!

Note: These equations are simplifications. You will get


complexes formed involving hydroxide ions, but the formulae
of these aren't very clear-cut. I am using these particular
versions of the equations to keep them in line with the
corresponding reaction between lead(II) oxide and sodium
hydroxide solution on the oxides of Group 4 page - also a
simplification!

http://www.chemguide.co.uk/inorganic/group4/lead.html (1 of 3)30/12/2004 11:13:22

some insoluble lead(II) compounds

Making lead(II) chloride


Lead(II) chloride can be made as a white precipitate by adding a solution
containing chloride ions to lead(II) nitrate solution. You could use things
like sodium chloride solution to provide the chloride ions, but it is usually
easier just to add some dilute hydrochloric acid.

Note: If you add concentrated hydrochloric acid to excess,


the lead(II) chloride precipitate will dissolve again. Complex
ions like PbCl42- are produced, and these are soluble in
water.

Making lead(II) iodide


If you add colourless potassium iodide solution (or any other source of
iodide ions in solution) to a solution of lead(II) nitrate, a bright yellow
precipitate of lead(II) iodide is produced.

Making lead(II) sulphate


Adding a source of aqueous sulphate ions to a solution of lead(II) nitrate
results in a white precipitate of lead(II) sulphate. The easiest thing to add
is usually dilute sulphuric acid - but any other soluble sulphate would do.

Where would you like to go now?


To the Group 4 menu . . .
http://www.chemguide.co.uk/inorganic/group4/lead.html (2 of 3)30/12/2004 11:13:22

some insoluble lead(II) compounds

To the Inorganic Chemistry menu . . .


To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/inorganic/group4/lead.html (3 of 3)30/12/2004 11:13:22

Periodic Table Group 7 menu

Understanding Chemistry

PERIODIC TABLE GROUP 7 (HALOGENS) MENU

Atomic and physical properties . . .


Discusses trends in atomic radius, electronegativity, electron
affinity and melting and boiling points of the Group 7 elements. It
also looks at the bond strengths of halogen-halogen bonds and of
hydrogen-halogen bonds.
Halogens as oxidising agents . . .
Describes and explains the trend in oxidising ability of the Group 7
elements based on the reactions between one halogen and the
ions of another one - for example, between Cl2 and I- ions from
salts like KI.
The acidity of the hydrogen halides . . .
Discusses the acidity of the hydrogen halides (like hydrogen
chloride), and explains why HF is a weak acid.
Halide ions as reducing agents . . .
Describes and explains the trend in reducing ability of the halide
ions based on their reactions with concentrated sulphuric acid.
Testing for halide ions . . .
Describes and explains the tests for halide ions using silver nitrate
solution followed by ammonia solution.

http://www.chemguide.co.uk/inorganic/group7menu.html (1 of 2)30/12/2004 11:13:27

Periodic Table Group 7 menu

Go to inorganic chemistry menu . . .


Go to Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/inorganic/group7menu.html (2 of 2)30/12/2004 11:13:27

Atomic and physical properties of Periodic Table Group 7 (the halogens)

ATOMIC AND PHYSICAL PROPERTIES OF THE


GROUP 7 ELEMENTS (THE HALOGENS)

This page explores the trends in some atomic and physical properties of
the Group 7 elements (the halogens) - fluorine, chlorine, bromine and
iodine. You will find separate sections below covering the trends in
atomic radius, electronegativity, electron affinity, and melting and boiling
points. There is also a section on the bond enthalpies (strengths) of
halogen-halogen bonds (for example, Cl-Cl) and of hydrogen-halogen
bonds (e.g. H-Cl)
Even if you aren't currently interested in all these things, it would
probably pay you to read the whole page. The same ideas tend to recur
throughout the atomic properties, and you may find that earlier
explanations help to you understand later ones.

Trends in Atomic Radius

Note: You will find atomic radius covered in detail in another


part of this site. If you choose to follow this link, use the
BACK button on your browser to return quickly to this page.

http://www.chemguide.co.uk/inorganic/group7/properties.html (1 of 13)30/12/2004 11:13:51

Atomic and physical properties of Periodic Table Group 7 (the halogens)

You can see that the atomic radius increases as you go down the Group.
Explaining the increase in atomic radius
The radius of an atom is governed by

the number of layers of electrons around the nucleus

the pull the outer electrons feel from the nucleus.

Compare fluorine and chlorine:


F

2,7

Cl

2,8,7

In each case, the outer electrons feel a net pull of 7+ from the nucleus.
The positive charge on the nucleus is cut down by the negativeness of
the inner electrons.

http://www.chemguide.co.uk/inorganic/group7/properties.html (2 of 13)30/12/2004 11:13:51

Atomic and physical properties of Periodic Table Group 7 (the halogens)

This is equally true for all the other atoms in Group 7. The outer electrons
always feel a net pull of 7+ from the centre.
The only factor which is going to affect the size of the atom is therefore
the number of layers of inner electrons which have to be fitted in around
the atom. Obviously, the more layers of electrons you have, the more
space they will take up - electrons repel each other. That means that the
atoms are bound to get bigger as you go down the Group.

Trends in Electronegativity
Electronegativity is a measure of the tendency of an atom to attract a
bonding pair of electrons. It is usually measured on the Pauling scale, on
which the most electronegative element (fluorine) is given an
electronegativity of 4.0.

Note: You will find electronegativity covered in detail in


another part of this site. If you choose to follow this link, use
the BACK button on your browser to return quickly to this
page.

http://www.chemguide.co.uk/inorganic/group7/properties.html (3 of 13)30/12/2004 11:13:51

Atomic and physical properties of Periodic Table Group 7 (the halogens)

Notice that electronegativity falls as you go down the Group. The atoms
become less good at attracting bonding pairs of electrons.
Explaining the decrease in electronegativity
This is easily shown using simple dots-and-crosses diagrams for
hydrogen fluoride and hydrogen chloride.

The bonding pair of electrons between the hydrogen and the halogen
feels the same net pull of 7+ from both the fluorine and the chlorine. (This
is exactly the same sort of argument as you have seen in the atomic
radius section above.) However, in the chlorine case, the nucleus is
further away from that bonding pair. That means that it won't be as
strongly attracted as in the fluorine case.
The larger pull from the closer fluorine nucleus is why fluorine is more
electronegative than chlorine is.
http://www.chemguide.co.uk/inorganic/group7/properties.html (4 of 13)30/12/2004 11:13:51

Atomic and physical properties of Periodic Table Group 7 (the halogens)

Summarising the trend down the Group


As the halogen atoms get bigger, any bonding pair gets further and
further away from the halogen nucleus, and so is less strongly attracted
towards it. In other words, as you go down the Group, the elements
become less electronegative.

Trends in First Electron Affinity


Defining first electron affinity
The first electron affinity is the energy released when 1 mole of gaseous
atoms each acquire an electron to form 1 mole of gaseous 1- ions.
This is more easily seen in symbol terms.

It is the energy released (per mole of X) when this change happens.


First electron affinities have negative values. For example, the first
electron affinity of chlorine is -364 kJ mol-1. By convention, the negative
sign shows a release of energy.

The first electron affinities of the Group 7 elements

http://www.chemguide.co.uk/inorganic/group7/properties.html (5 of 13)30/12/2004 11:13:51

Atomic and physical properties of Periodic Table Group 7 (the halogens)

Note: You will find electron affinity covered in detail in


another part of this site. The current page duplicates much of
that material, but you might like to read it again in different
words.
If you choose to follow this link, use the BACK button on your
browser to return quickly to this page.

Notice that the trend down the Group isn't tidy. The tendency is for the
electron affinities to decrease (in the sense that less heat is given out),
but the fluorine value is out of line.
The electron affinity is a measure of the attraction between the incoming
electron and the nucleus. The higher the attraction, the higher the
electron affinity.

http://www.chemguide.co.uk/inorganic/group7/properties.html (6 of 13)30/12/2004 11:13:51

Atomic and physical properties of Periodic Table Group 7 (the halogens)

In the bigger atom, the attraction from the more positive nucleus is offset
by the additional screening electrons, so each incoming electron feels
the effect of a net 7+ charges from the centre - exactly as when you are
thinking about atomic radius or electronegativity.
As the atom gets bigger, the incoming electron is further from the
nucleus and so feels less attraction. The electron affinity therefore falls
as you go down the Group.
But what about fluorine? That is a very small atom, with the incoming
electron quite close to the nucleus. Why isn't its electron affinity bigger
than chlorine's?
There is another effect operating. When the new electron comes into the
atom, it is entering a region of space already very negatively charged
because of the existing electrons. There is bound to be some repulsion,
offsetting some of the attraction from the nucleus.
In the case of fluorine, because the atom is very small, the existing
electron density is very high. That means that the extra repulsion is
particularly great and lessens the attraction from the nucleus enough to
lower the electron affinity below that of chlorine.

Trends in Melting Point and Boiling Point

http://www.chemguide.co.uk/inorganic/group7/properties.html (7 of 13)30/12/2004 11:13:51

Atomic and physical properties of Periodic Table Group 7 (the halogens)

You will see that both melting points and boiling points rise as you go
down the Group.
If you explore the graphs, you will find that fluorine and chlorine are
gases at room temperature, bromine is a liquid and iodine a solid.
Nothing very surprising there!
Explaining the trends in melting point and boiling point
All of the halogens exist as diatomic molecules - F2, Cl2, and so on. The
intermolecular attractions between one molecule and its neighbours are
van der Waals dispersion forces.

http://www.chemguide.co.uk/inorganic/group7/properties.html (8 of 13)30/12/2004 11:13:51

Atomic and physical properties of Periodic Table Group 7 (the halogens)

Note: If you aren't sure about van der Waals dispersion


forces, you will find them covered in detail in another part of
this site. You won't understand the next bit unless you are
happy about dispersion forces and how they vary with the
size of the molecule.
Use the BACK button on your browser to return quickly to this
page.

As the molecules get bigger there are obviously more electrons which
can move around and set up the temporary dipoles which create these
attractions.
The stronger intermolecular attractions as the molecules get bigger
means that you have to supply more heat energy to turn them into either
a liquid or a gas - and so their melting and boiling points rise.

Bond enthalpies (bond energies or bond strengths)


Bond enthalpy is the heat needed to break one mole of a covalent bond
to produce individual atoms, starting from the original substance in the
gas state, and ending with gaseous atoms.
So for chlorine, Cl2(g), it is the heat energy needed to carry out this
change per mole of bond:

For bromine, the reaction is still from gaseous bromine molecules to


separate gaseous atoms.

http://www.chemguide.co.uk/inorganic/group7/properties.html (9 of 13)30/12/2004 11:13:51

Atomic and physical properties of Periodic Table Group 7 (the halogens)

Bond enthalpy in the halogens, X2(g)


A covalent bond works because the bonding pair is attracted to both the
nuclei at either side of it. It is that attraction which holds the molecule
together. The size of the attraction will depend, amongst other things, on
the distance from the bonding pair to the two nuclei.

As with all halogens, the bonding pair will feel a net pull of 7+ from both
ends of the bond - the charge on the nucleus offset by the inner
electrons. That will still be the same whatever the size of the halogen
atoms.
As the atoms get bigger, the bonding pair gets further from the nuclei and
so you would expect the strength of the bond to fall.

http://www.chemguide.co.uk/inorganic/group7/properties.html (10 of 13)30/12/2004 11:13:51

Atomic and physical properties of Periodic Table Group 7 (the halogens)

So . . . are the actual bond enthalpies in line with this prediction?

The bond enthalpies of the Cl-Cl, Br-Br and I-I bonds fall just as you
would expect, but the F-F bond is way out of line!
Because fluorine atoms are so small, you might expect a very strong
bond - in fact, it is remarkably weak. There must be another factor at
work as well.
As well as the bonding pair of electrons between the two atoms, each
atom has 3 non-bonding pairs of electrons in the outer level - lone pairs.
Where the bond gets very short (as in F-F), the lone pairs on the two
atoms get close enough together to set up a significant amount of
repulsion.

http://www.chemguide.co.uk/inorganic/group7/properties.html (11 of 13)30/12/2004 11:13:51

Atomic and physical properties of Periodic Table Group 7 (the halogens)

In the case of fluorine, this repulsion is great enough to counteract quite


a lot of the attraction between the bonding pair and the two nuclei. This
obviously weakens the bond.

Bond enthalpies in the hydrogen halides, HX(g)


Where the halogen atom is attached to a hydrogen atom, this effect
doesn't happen. There are no lone pairs on a hydrogen atom!

http://www.chemguide.co.uk/inorganic/group7/properties.html (12 of 13)30/12/2004 11:13:51

Atomic and physical properties of Periodic Table Group 7 (the halogens)

As the halogen atom gets bigger, the bonding pair gets more and more
distant from the nucleus. The attraction is less, and the bond gets weaker
- exactly what is shown by the data. There is nothing complicated
happening in this case.

Where would you like to go now?


To the Group 7 menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/inorganic/group7/properties.html (13 of 13)30/12/2004 11:13:51

electron affinity

ELECTRON AFFINITY

This page explains what electron affinity is, and then looks at the factors
that affect its size. It assumes that you know about simple atomic
orbitals, and can write electronic structures for simple atoms.

Important! If you aren't reasonable happy about atomic


orbitals and electronic structures you should follow these
links before you go any further.

First electron affinity


Ionisation energies are always concerned with the formation of positive
ions. Electron affinities are the negative ion equivalent, and their use is
almost always confined to elements in groups 6 and 7 of the Periodic
Table.
Defining first electron affinity
The first electron affinity is the energy released when 1 mole of gaseous
atoms each acquire an electron to form 1 mole of gaseous 1- ions.
This is more easily seen in symbol terms.

It is the energy released (per mole of X) when this change happens.


First electron affinities have negative values. For example, the first
electron affinity of chlorine is -349 kJ mol-1. By convention, the negative
sign shows a release of energy.

http://www.chemguide.co.uk/atoms/properties/eas.html (1 of 6)30/12/2004 11:14:00

electron affinity

The first electron affinities of the group 7 elements


F

-328 kJ mol-1

Cl

-349 kJ mol-1

Br

-324 kJ mol-1

-295 kJ mol-1

Note: These values are based on the most recent research.


If you are using a different data source, you may have slightly
different numbers. That doesn't matter - the pattern will still
be the same.

Is there a pattern?
Yes - as you go down the group, first electron affinities become less (in
the sense that less energy is evolved when the negative ions are
formed). Fluorine breaks that pattern, and will have to be accounted for
separately.
The electron affinity is a measure of the attraction between the incoming
electron and the nucleus - the stronger the attraction, the more energy is
released.
The factors which affect this attraction are exactly the same as those
relating to ionisation energies - nuclear charge, distance and screening.

Note: If you haven't read about ionisation energy recently, it


might be a good idea to follow this link before you go on.
These factors are discussed in more detail on that page than
they are on this one.

http://www.chemguide.co.uk/atoms/properties/eas.html (2 of 6)30/12/2004 11:14:00

electron affinity

The increased nuclear charge as you go down the group is offset by


extra screening electrons. Each outer electron in effect feels a pull of 7+
from the centre of the atom, irrespective of which element you are talking
about.
For example, a fluorine atom has an electronic structure of
1s22s22px22py22pz1. It has 9 protons in the nucleus.
The incoming electron enters the 2-level, and is screened from the
nucleus by the two 1s2 electrons. It therefore feels a net attraction from
the nucleus of 7+ (9 protons less the 2 screening electrons).
By contrast, chlorine has the electronic structure
1s22s22p63s23px23py23pz1. It has 17 protons in the nucleus.
But again the incoming electron feels a net attraction from the nucleus of
7+ (17 protons less the 10 screening electrons in the first and second
levels).

Note: If you want to be fussy, there is also a small amount of


screening by the 2s electrons in fluorine and by the 3s
electrons in chlorine. This will be approximately the same in
both these cases and so doesn't affect the argument in any
way (apart from complicating it!).

The over-riding factor is therefore the increased distance that the


incoming electron finds itself from the nucleus as you go down the group.
The greater the distance, the less the attraction and so the less energy is
released as electron affinity.

http://www.chemguide.co.uk/atoms/properties/eas.html (3 of 6)30/12/2004 11:14:00

electron affinity

Note: Comparing fluorine and chlorine isn't ideal, because


fluorine breaks the trend in the group. However, comparing
chlorine and bromine, say, makes things seem more difficult
because of the more complicated electronic structures
involved.
What we have said so far is perfectly true and applies to the
fluorine-chlorine case as much as to anything else in the
group, but there's another factor which operates as well
which we haven't considered yet - and that over-rides the
effect of distance in the case of fluorine.

Why is fluorine out of line?


The incoming electron is going to be closer to the nucleus in fluorine than
in any other of these elements, so you would expect a high value of
electron affinity.
However, because fluorine is such a small atom, you are putting the new
electron into a region of space already crowded with electrons and there
is a significant amount of repulsion. This repulsion lessens the attraction
the incoming electron feels and so lessens the electron affinity.
A similar reversal of the expected trend happens between oxygen and
sulphur in Group 6. The first electron affinity of oxygen (-142 kJ mol-1) is
smaller than that of sulphur (-200 kJ mol-1) for exactly the same reason
that fluorine's is smaller than chlorine's.

Comparing Group 6 and Group 7 values


As you might have noticed, the first electron affinity of oxygen (-142 kJ
mol-1) is less than that of fluorine (-328 kJ mol-1). Similarly sulphur's (200 kJ mol-1) is less than chlorine's (-349 kJ mol-1). Why?
It's simply that the Group 6 element has 1 less proton in the nucleus than
its next door neighbour in Group 7. The amount of screening is the same
http://www.chemguide.co.uk/atoms/properties/eas.html (4 of 6)30/12/2004 11:14:00

electron affinity

in both.
That means that the net pull from the nucleus is less in Group 6 than in
Group 7, and so the electron affinities are less.

First electron affinity and reactivity


The reactivity of the elements in group 7 falls as you go down the group fluorine is the most reactive and iodine the least.
Often in their reactions these elements form their negative ions. At GCSE
the impression is sometimes given that the fall in reactivity is because
the incoming electron is held less strongly as you go down the group and
so the negative ion is less likely to form. That explanation looks
reasonable until you include fluorine!
An overall reaction will be made up of lots of different steps all involving
energy changes, and you cannot safely try to explain a trend in terms of
just one of those steps. Fluorine is much more reactive than chlorine
(despite the lower electron affinity) because the energy released in other
steps in its reactions more than makes up for the lower amount of energy
released as electron affinity.

Second electron affinity


You are only ever likely to meet this with respect to the group 6 elements
oxygen and sulphur which both form 2- ions.
Defining second electron affinity
The second electron affinity is the energy required to add an electron to
each ion in 1 mole of gaseous 1- ions to produce 1 mole of gaseous 2ions.
This is more easily seen in symbol terms.

http://www.chemguide.co.uk/atoms/properties/eas.html (5 of 6)30/12/2004 11:14:00

electron affinity

It is the energy needed to carry out this change per mole of X-.
Why is energy needed to do this?
You are forcing an electron into an already negative ion. It's not going to
go in willingly!
1st EA = -142 kJ mol-1
2nd EA = +844 kJ mol-1
The positive sign shows that you have to put in energy to perform this
change. The second electron affinity of oxygen is particularly high
because the electron is being forced into a small, very electron-dense
space.

Where would you like to go now?


To the atomic properties menu . . .
To the atomic structure and bonding menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/atoms/properties/eas.html (6 of 6)30/12/2004 11:14:00

Halogens as oxidising agents

THE OXIDISING ABILITY OF THE GROUP 7


ELEMENTS (THE HALOGENS)

This page explores the trend in oxidising ability of the Group 7 elements
(the halogens) - fluorine, chlorine, bromine and iodine. We are going to
look at the ability of one halogen to oxidise the ions of another one, and
how that changes as you go down the Group.

Note: If you aren't comfortable with terms like oxidation and


oxidising agent in terms of electron transfer, then you should
explore the area of the site dealing with redox reactions
before you go on.

The facts
We are going to look at the reactions between one halogen (chlorine,
say) and the ions of another one (iodide ions, perhaps). The iodide ions
will be in a solution of a salt like sodium or potassium iodide. The sodium
or potassium ions will be spectator ions, and are completely irrelevant to
the reaction.
In the chlorine and iodide ion case, the reaction would be:

The iodide ions have lost electrons to form iodine molecules. They have
been oxidised.
The chlorine molecules have gained electrons to form chloride ions. They
have been reduced.
This is obviously a redox reaction in which chlorine is acting as an
oxidising agent.
http://www.chemguide.co.uk/inorganic/group7/halogensasoas.html (1 of 9)30/12/2004 11:14:08

Halogens as oxidising agents

Fluorine
We'll have to exclude fluorine from this descriptive bit, because it is too
strong an oxidising agent. Fluorine oxidises water to oxygen and so it is
impossible to do simple solution reactions with it.

Chlorine, bromine and iodine


In each case, a halogen higher in the Group can oxidise the ions of one
lower down. For example, chlorine can oxidise the bromide ions (in, for
example, potassium bromide solution) to bromine:

The bromine appears as an orange solution.


As you have seen above, chlorine can also oxidise iodide ions (in, for
example, potassium iodide solution) to iodine:

The iodine appears either as a red solution if you are mean with the
amount of chlorine you use, or as a dark grey precipitate if the chlorine is
in excess.

Note: The reason for the red solution is that iodine dissolves
in potassium iodide (or other soluble iodides) by reacting to
give a red ion, I3-. If the chlorine is in excess, obviously there
isn't anything left for the iodine to react with, and so it
remains as a dark grey precipitate.

http://www.chemguide.co.uk/inorganic/group7/halogensasoas.html (2 of 9)30/12/2004 11:14:08

Halogens as oxidising agents

Bromine can only oxidise iodide ions to iodine. It isn't a strong enough
oxidising agent to convert chloride ions into chlorine. (You have just seen
exactly the reverse of that happening.)
A red solution of iodine is formed (see the note above) until the bromine
is in excess. Then you get a dark grey precipitate.

Iodine won't oxidise any of the other halide ions (unless you happened to
have some extremely radioactive and amazingly rare astatide ions astatine is at the bottom of this Group).
To summarise

Oxidation is loss of electrons. Each of the elements (for example,


chlorine) could potentially take electrons from something else to
make their ions (e.g. Cl-). That means that they are all potentially
oxidising agents.
Fluorine is such a powerful oxidising agent that you can't
reasonably do solution reactions with it.
Chlorine has the ability to take electrons from both bromide ions
and iodide ions. Bromine and iodine can't get those electrons back
from the chloride ions formed.
That means that chlorine is a more powerful oxidising agent than
either bromine or iodine.

Similarly bromine is a more powerful oxidising agent than iodine.


Bromine can remove electrons from iodide ions to give iodine and the iodine can't get them back from the bromide ions formed.

This all means that oxidising ability falls as you go down the Group.

Explaining the trend

http://www.chemguide.co.uk/inorganic/group7/halogensasoas.html (3 of 9)30/12/2004 11:14:08

Halogens as oxidising agents

Whenever one of these halogens is involved in oxidising something in


solution, the halogen ends up as halide ions with water molecules
attached to them. Looking at all four of the common halogens:

As you go down the Group, the ease with which these hydrated ions are
formed falls, and so the halogens become less good as oxidising agents
- less ready to take electrons from something else.
The reason that the hydrated ions form less readily as you go down the
Group is a fairly complicated mixture of several factors. Unfortunately,
this is often over-simplified to give what is actually a faulty and
misleading explanation. We'll deal with this first before giving a proper
explanation.
The faulty explanation
This is normally given for the trend in oxidising ability of chlorine, bromine
and iodine, and goes like this:
How easily the element forms its ions depends on how strongly the new
electrons are attracted. As the atoms get bigger, the new electrons find
themselves further from the nucleus, and more and more screened from
it by the inner electrons (offsetting the effect of the greater nuclear
charge). The bigger atoms are therefore less good at attracting new
electrons and forming ions.
That sounds reasonable! What's wrong with it?
What we are describing is the trend in electron affinity as you go from
chlorine to bromine to iodine. Electron affinity tends to fall as you go
down the Group. This is described in detail on another page.

http://www.chemguide.co.uk/inorganic/group7/halogensasoas.html (4 of 9)30/12/2004 11:14:08

Halogens as oxidising agents

Note: If you haven't recently read about the electron


affinities of the halogens, you ought to follow this link before
you go on.
Use the BACK button on your browser to return to this page.

The snag comes if you try to expand the argument to include fluorine.
Fluorine has a much higher tendency to form its hydrated ion than
chlorine does. BUT . . . the tendency of the fluorine atom to gain an
electron is less than that of chlorine - as measured by its electron affinity!
That makes a nonsense of the whole argument.
So, what is going wrong? The mistake is to look at only one part of a
much more complicated process. The argument about atoms accepting
electrons applies to isolated atoms in the gas state picking up electrons
to make isolated ions - also in the gas state. That's not what we should
be talking about.
In reality:

The halogen starts as diatomic molecules, X2 - which may be gas,


liquid or solid, depending on the halogen.
These have to be split apart to make individual atoms.
Those atoms each gain an electron. (That's the stage of the
process we've been concentrating on in the faulty explanation.)
The isolated ions become wrapped in water molecules to form
hydrated ions.

The proper explanation


The table below looks at how much energy is involved in each of these
changes. To be sure that you understand the various terms:
http://www.chemguide.co.uk/inorganic/group7/halogensasoas.html (5 of 9)30/12/2004 11:14:08

Halogens as oxidising agents

Atomisation energy
This is the energy needed to produce 1 mole of isolated gaseous
atoms starting from an element in its standard state (gas for
chlorine, and liquid for bromine, for example - both of them as X2).
For a gas like chlorine, this is simply half of the bond enthalpy
(because breaking a Cl-Cl bond produces 2 chlorine atoms, not 1).
For a liquid like bromine or a solid like iodine, it also includes the
energy that is needed to convert them into gases.
Electron affinity
The first electron affinity is the energy released when 1 mole of
gaseous atoms each acquire an electron to form 1 mole of
gaseous 1- ions.
In symbol terms:

Hydration enthalpy (hydration energy)


This is the energy released when 1 mole of gaseous ions
dissolves in water to produce hydrated ions.

atomisation
energy
(kJ mol-1)

electron
affinity
(kJ mol-1)

hydration
enthalpy
(kJ mol-1)

overall
(kJ mol1)

+79

-328

-506

-755

Cl

+121

-349

-364

-592

Br

+112

-324

-335

-547

http://www.chemguide.co.uk/inorganic/group7/halogensasoas.html (6 of 9)30/12/2004 11:14:08

Halogens as oxidising agents

+107

-295

-293

-481

There's quite a lot of data here to look at. Concentrate first on the final
column which shows the overall heat evolved when all the other
processes happen. It is calculated by adding the figures in the previous 3
columns.
You can see that the amount of heat evolved falls quite dramatically from
the top to the bottom of the Group, with the biggest fall from fluorine to
chlorine.
Fluorine produces a lot of heat when it forms its hydrated ion, chlorine
less so, and so on down the Group.

Note: Don't forget that we are only talking about half of a


redox reaction in each case. There will be other energy terms
involving whatever the halogen is oxidising. Those changes
will be overall endothermic. For example, if chlorine oxidises
iodide ions to iodine, that half of the total reaction would need
+481 kJ mol-1, giving an enthalpy change of reaction of (-592
+ 481) = -111 kJ per mole of I- oxidised.

Why is fluorine a much stronger oxidising agent than chlorine?


What produces the very negative value for the enthalpy change when
fluorine turns into its hydrated ions? There are two main factors.
The atomisation energy of fluorine is abnormally low. This reflects the
low bond enthalpy of fluorine.

Note: The reason for fluorine's low bond enthalpy is


described on another page.

http://www.chemguide.co.uk/inorganic/group7/halogensasoas.html (7 of 9)30/12/2004 11:14:08

Halogens as oxidising agents

The main reason, though, is the very high hydration enthalpy of the
fluoride ion. That is because the ion is very small. There is a very strong
attraction between the fluoride ions and water molecules. The stronger
the attraction, the more heat is evolved when the hydrated ions are
formed.
Why the fall in oxidising ability from chlorine to bromine to iodine?
The fall in atomisation energy between these three elements is fairly
slight, and would tend to make the overall change more negative as you
go down the Group. The explanation doesn't lie there!
It is helpful to look at the changes in electron affinity and hydration
enthalpy as you go down the Group. Using the figures from the previous
table:

change in electron
affinity
(kJ mol-1)

change in hydration
enthalpy
(kJ mol-1)

Cl to Br

+25

+29

Br to I

+29

+42

going from

You can see that both of these effects matter, but that the more
important one - the one that changes the most - is the change in the
hydration enthalpy.
As you go down the Group, the ions become less attractive to water
molecules as they get bigger. Although the ease with which an atom
attracts an electron matters, it isn't actually as important as the hydration
enthalpy of the negative ion formed.
The faulty explanation misses the mark even if you restrict it to chlorine,
bromine and iodine!

http://www.chemguide.co.uk/inorganic/group7/halogensasoas.html (8 of 9)30/12/2004 11:14:08

Halogens as oxidising agents

Warning! You really need to find out what (if any)


explanation your examiners expect you to give for this. If their
mark schemes (or the way they phrase their questions)
suggest that they want the faulty explanation, there isn't
much you can do about it. Unfortunately, there are times in
exams when you have to grit your teeth and give technically
wrong answers because that's what your examiners want. It
shouldn't happen like this, but it does!
UK A' level students should search their syllabuses, past
exam papers, mark schemes and any other support material
available from their Exam Board. If you haven't got any of
this, you can find your Exam Board's web address by
following this link. Students elsewhere should find out the
equivalent information from their own sources.

Where would you like to go now?


To the Group 7 menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/inorganic/group7/halogensasoas.html (9 of 9)30/12/2004 11:14:08

Hydrogen halides as acids

THE ACIDITY OF THE HYDROGEN HALIDES

This page looks at the acidity of the hydrogen halides - hydrogen


fluoride, hydrogen chloride, hydrogen bromide and hydrogen iodide. It
starts by describing their physical properties and how they might be
made and then explains what happens when they react with water to
make acids like hydrofluoric acid and hydrochloric acid.
There is then a lengthy discussion (should you need it!) of why
hydrofluoric acid is a weak acid.

The hydrogen halides - background information


Physical properties
The hydrogen halides are colourless gases at room temperature,
producing steamy fumes in moist air. Hydrogen fluoride has an
abnormally high boiling point for the size of the molecule (293 K or 20C),
and could condense to a liquid on a cool day.

Hydrogen fluoride's boiling point is higher than you might expect because
http://www.chemguide.co.uk/inorganic/group7/acidityhx.html (1 of 15)30/12/2004 11:14:22

Hydrogen halides as acids

it forms hydrogen bonds.

Note: If you aren't happy about hydrogen bonding, you


ought to follow this link before you go on.
Use the BACK button on your browser to return to this page.

Fluorine is the most electronegative of all the elements and the bond
between it and hydrogen is very polar. The hydrogen atom carries quite a
lot of positive charge ( +); the fluorine is fairly negatively charged ( -).
In addition, each fluorine atom has 3 very active lone pairs of electrons.
Fluorine's outer electrons are at the 2-level, and the lone pairs represent
small highly charged regions of space. Hydrogen bonds form between
the + hydrogen on one HF molecule and a lone pair on the fluorine of
another one.

The other hydrogen halides don't form hydrogen bonds. The other
halogens aren't as electronegative as fluorine, and so the bonds in HX
are less polar. As well as that, their lone pairs are at higher energy
levels. That makes the lone pairs bigger, and so they don't carry such an
intensely concentrated negative charge for the hydrogens to be attracted
to.

http://www.chemguide.co.uk/inorganic/group7/acidityhx.html (2 of 15)30/12/2004 11:14:22

Hydrogen halides as acids

Note: If you aren't sure about why the electronegativity of


the halogens changes as you go down the Group, you could
follow this link.
Use the BACK button on your browser to return to this page.

Making the hydrogen halides


There are several ways of making hydrogen halides, but the only one of
interest at A' level is the reaction between an ionic halide (like sodium
chloride) and an acid like concentrated phosphoric(V) acid, H3PO4, or
concentrated sulphuric acid.
Making hydrogen chloride
You can add concentrated sulphuric acid to a solid chloride like sodium
chloride in the cold. The concentrated sulphuric acid donates a hydrogen
ion to a chloride ion to make hydrogen chloride. Because this is a gas, it
immediately escapes from the system.

The full equation for the reaction is:

Sodium hydrogensulphate is also formed.


Concentrated phosphoric(V) acid behaves similarly. You would again
add it to solid sodium choride. Once again, as soon as any hydrogen
chloride is formed, it escapes as a gas. The ionic equation is:

. . . and the full one showing the formation of the salt, sodium
dihydrogenphosphate(V) is:

http://www.chemguide.co.uk/inorganic/group7/acidityhx.html (3 of 15)30/12/2004 11:14:22

Hydrogen halides as acids

Making the other hydrogen halides


All of the hydrogen halides can be made in an exactly similar way using
concentrated phosphoric(V) acid. All you would need to do is swap the
symbol Cl in the two equations for whichever other halogen you were
interested in.
The situation is more complicated with concentrated sulphuric acid.
Hydrogen fluoride can be made in exactly the same way as hydrogen
chloride using concentrated sulphuric acid, but hydrogen bromide and
hydrogen iodide can't.
The problem is that concentrated sulphuric acid is a reasonably strong
oxidising agent, and as well as producing hydrogen bromide or hydrogen
iodide, some of the halide ions are oxidised to bromine or iodine. This
problem doesn't happen with phosphoric(V) acid because it isn't an
oxidising agent.

Note: Redox reactions involving halide ions and conc


sulphuric acid are covered on a separate page.

The acidity of the hydrogen halides


Hydrogen chloride as an acid
We are going to use the Bronsted-Lowry definition of an acid as a proton
donor. Hydrogen chloride is an acid because it gives protons (hydrogen
ions) to other things. We are going to concentrate on its reaction with
water.
Hydrogen chloride gas is very soluble in water, reacting with it to produce
hydrochloric acid. The familiar steamy fumes of hydrogen chloride in
moist air are caused by the hydrogen chloride reacting with water vapour
http://www.chemguide.co.uk/inorganic/group7/acidityhx.html (4 of 15)30/12/2004 11:14:22

Hydrogen halides as acids

in the air to produce a fog of concentrated hydrochloric acid.


A proton is donated from the hydrogen chloride to one of the lone pairs
on a water molecule.

A co-ordinate (dative covalent) bond is formed between the oxygen and


the transferred hydrogen ion.

Note: If you need to revise co-ordinate (dative covalent)


bonding, you could follow this link. That page also describes
the reaction between hydrogen chloride and ammonia another reaction of hydrogen chloride as an acid.
Use the BACK button on your browser to return to this page.

The equation for the reaction is:

The H3O+ ion is the hydroxonium ion (also known as the hydronium ion
or the oxonium ion). This is the ion that we are actually talking about
when we write H+(aq).
When hydrogen chloride dissolves in water (to produce hydrochloric
acid), almost 100% of the hydrogen chloride molecules react in this way.
Hydrochloric acid is therefore a strong acid. A strong acid is one which
is fully ionised in solution.
Hydrobromic acid and hydriodic acid as strong acids
http://www.chemguide.co.uk/inorganic/group7/acidityhx.html (5 of 15)30/12/2004 11:14:22

Hydrogen halides as acids

Hydrogen bromide and hydrogen iodide dissolve in (and react with) water
in exactly the same way as hydrogen chloride does. Hydrogen bromide
reacts to give hydrobromic acid; hydrogen iodide gives hydriodic acid.
Both of these are also strong acids.
Hydrofluoric acid as an exception
By contrast, although hydrogen fluoride dissolves freely in water,
hydrofluoric acid is only a weak acid - similar in strength to organic acids
like methanoic acid. The reason for this is quite complicated.

Warning! Are you sure you need to know the reason for
this? The explanation is far more complicated than books at
this level tend to suggest. To understand it properly you need
to be familiar with energy cycles, entropy, free energy and
equilibrium constants.
The explanation starts fairly easily, but gets harder and
harder as you go on!

Looking at the bond enthalpy of the H-F bond


Because the fluorine atom is so small, the bond enthalpy (bond energy)
of the hydrogen-fluorine bond is very high. Comparing all the hydrogenhalogen bond enthalpies:

bond enthalpy
(kJ mol-1)
H-F

+562

H-Cl

+431

H-Br

+366

http://www.chemguide.co.uk/inorganic/group7/acidityhx.html (6 of 15)30/12/2004 11:14:22

Hydrogen halides as acids

H-I

+299

In order for ions to form when the hydrogen fluoride reacts with water,
you first have to break the H-F bond. It would seem reasonable to say
that the relative reluctance of hydrogen fluoride to react with water is due
to the large amount of energy needed to break that bond.
It might seem reasonable, but if you dig a little deeper, that explanation
falls apart!
Looking at the energetics of the process from HX(g) to X-(aq)
We need to consider the energetics of this sequence:

All of these terms are involved in the overall enthalpy change as you
convert HX(g) into its ions in water.
However, the terms involving the hydrogen will be the same for every
hydrogen halide. So if we are just trying to draw comparisons, we only
need to look at the terms shown in red in the diagram.

http://www.chemguide.co.uk/inorganic/group7/acidityhx.html (7 of 15)30/12/2004 11:14:22

Hydrogen halides as acids

bond
enthalpy of
HX
(kJ mol-1)

electron
affinity of X
(kJ mol-1)

hydration
enthalpy of X(kJ mol-1)

sum of
these
(kJ mol1)

HF

+562

-328

-506

-272

HCl

+431

-349

-364

-282

HBr

+366

-324

-335

-293

HI

+299

-295

-293

-289

If you compare the total HF and HCl values, there is virtually no


difference.

Note: There is so much discrepancy between the numbers


we are using if you obtain them from different sources that
you couldn't count a 10 kJ difference in the totals as being
significant. These particular figures come from a variety of
sources of varying reliability!

The large bond enthalpy of the H-F bond is compensated for by the large
hydration enthalpy of the fluoride ion. There is a very strong attraction
between the very small fluoride ion and the water molecules. This
releases a lot of heat (the hydration enthalpy) when the fluoride ion
becomes wrapped in water molecules.
The fact that there isn't much difference between the overall values
means that we still aren't looking in the right place for the explanation of
why hydrofluoric acid is weak!
Looking at other attractions in the system
The mistake we are making so far is in starting from the wrong place!
The energy terms we have been looking at start from HX as a gas. In
fact, we should be thinking of it starting in solution - but not yet reacted
http://www.chemguide.co.uk/inorganic/group7/acidityhx.html (8 of 15)30/12/2004 11:14:22

Hydrogen halides as acids

with the water.


The proper equation we should be working from is:

That then needs to be incorporated into an improved energy cycle:

Unfortunately, at this point I've been unable to find any figures for that
first stage. However, in each case, that initial separation of the HX from
water molecules will be endothermic. Energy is needed to break the
intermolecular attractions between the HX molecules and water.
That energy will be much greater in the case of hydrogen fluoride
because it forms hydrogen bonds with the water molecules. The other
hydrogen halides only form the weaker van der Waals dispersion forces
or dipole-dipole attractions.

http://www.chemguide.co.uk/inorganic/group7/acidityhx.html (9 of 15)30/12/2004 11:14:22

Hydrogen halides as acids

Note: If you aren't sure about van der Waals forces, you
could follow this link.
Use the BACK button on your browser to return to this page.

Putting all this together, here are the overall enthalpy changes (including
all the stages in the energy cycle) for the reactions:

enthalpy change
(kJ mol-1)
HF

-13

HCl

-59

HBr

-63

H-I

-57

Note: These figures are taken from Modern Physical


Chemistry by Liptrot, Thompson and Walker. They are
quoted in that book without any explanation of where they
came from. I have no way of checking their accuracy.

http://www.chemguide.co.uk/inorganic/group7/acidityhx.html (10 of 15)30/12/2004 11:14:22

Hydrogen halides as acids

You can see that the enthalpy change for HF is much lower than that for
the other three hydrogen halides - but it is still an exothermic change. We
still haven't got fully to the bottom of why hydrofluoric acid is a weak acid!
Entropy and free energy considerations
What decides the extent to which a reaction happens isn't enthalpy
change but is a term called free energy change.
Free energy change is calculated from the enthalpy change, the
temperature of the reaction and the entropy change during the reaction.
The only way of making sense of entropy without getting bogged down in
some serious maths is to think of it as a measure of the amount of
disorder in a system. Entropy is given the symbol S. If a system becomes
more disordered, then its entropy increases. If it becomes more ordered,
its entropy decreases.
The key equation is:

As an approximation, for a reaction to happen, the free energy change


must be negative. But more accurately, free energy change can be used
to calculate a value for the equilibrium constant for a reaction using the
expression:

http://www.chemguide.co.uk/inorganic/group7/acidityhx.html (11 of 15)30/12/2004 11:14:22

Hydrogen halides as acids

The equilibrium constant, Ka, refers to the reaction:

The values for T S (needed to calculate G) for the four reactions at a


temperature of 298 K are:

T S
(kJ mol-1)
HF

-29

HCl

-13

HBr

-4

H-I

+4

Note: These figures are again taken from Modern Physical


Chemistry by Liptrot, Thompson and Walker. Again, I don't
know how they were calculated, and have no way of
checking their accuracy.

http://www.chemguide.co.uk/inorganic/group7/acidityhx.html (12 of 15)30/12/2004 11:14:22

Hydrogen halides as acids

Notice that at the top of the Group, the systems are becoming more
ordered when the HX reacts with the water. The entropy of the system
(the amount of disorder) is decreasing - particularly with the hydrogen
fluoride.
The reason for this is that the very strong attractions between H3O+ and
F-(aq) ions imposes a lot of order on the system, as also does the
attraction between the water molecules and the various ions present.
These attractions (of both kinds) will be at their greatest in the case of
the very small fluoride ions.
Putting all this together, what is the effect on the free energy change, and
therefore the value of the equilibrium constant for the reaction?

Ka

H
(kJ mol-1)

T S
(kJ mol-1)

G
(kJ mol-1)

(mol dm-3)

HF

-13

-29

+16

1.6 x 10-3

HCl

-59

-13

-46

1.2 x 108

HBr

-63

-4

-59

2.2 x 1010

HI

-57

+4

-61

5.0 x 1010

The values for these estimated equilibrium constants for HCl, HBr and HI
are so high that you can think of the reaction as being essentially "oneway". The ionisation is virtually 100% complete. These are all strong
acids - but getting somewhat stronger as you go down the Group.
By contrast, the estimated equilibrium constant of hydrofluoric acid is
pretty small. Hydrofluoric acid only ionises to a very small extent in water.
It is a weak acid.
It is possible to check the accuracy of our explanation by comparing the
real value of the equilibrium constant with the estimated one.

http://www.chemguide.co.uk/inorganic/group7/acidityhx.html (13 of 15)30/12/2004 11:14:22

Hydrogen halides as acids

Real value: 5.6 x 10-4 mol dm-3

Estimated value: 1.6 x 10-3 mol dm-3

You might think that these values aren't all that similar, but in fact they
are in remarkable agreement! Because of the way the maths works, a
very small change in G has a very large effect on Ka.
To have the values in close agreement, G would only have to increase
from +16 to +18.5 kJ mol-1. Given the uncertainty about the values used
to calculate G, our calculated value could easily be that much out.
To be honest, the calculated values for Ka shouldn't really have been
quoted beyond 1 significant figure. I've quoted them slightly more
accurately than they deserve in order to pick out the trend in the three
strong acids.
Summary: Why is hydrofluoric acid a weak acid?
The two main factors are:

Very strong hydrogen bonding between the un-ionised hydrogen


fluoride molecules and water molecules. This costs quite a lot of
energy to break, and doesn't occur in the other hydrogen halides.
A large decrease in entropy when the hydrogen fluoride molecules
react with water. This is particularly noticeable with hydrogen
fluoride because the attractiveness of the very small fluoride ions
produced imposes a lot of order on the surrounding water
molecules, and also on nearby hydroxonium ions. The effect falls
as the halide ions get bigger.

A final thought
A Peanuts cartoon showed one of the others (Linus?) asking the awful
Lucy why the sky was blue. The conversation went something like this:
Linus: "Why is the sky blue?"

http://www.chemguide.co.uk/inorganic/group7/acidityhx.html (14 of 15)30/12/2004 11:14:22

Hydrogen halides as acids

Lucy: "BECAUSE IT ISN'T GREEN!"


Linus: "Oh . . . I thought the reason would be much more
complicated than that."
Make of that what you will!

Where would you like to go now?


To the Group 7 menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/inorganic/group7/acidityhx.html (15 of 15)30/12/2004 11:14:22

redox reactions involving halide ions and sulphuric acid

THE REDOX REACTIONS BETWEEN HALIDE


IONS AND CONCENTRATED SULPHURIC ACID

This page describes and explains the redox reactions involving halide
ions and concentrated sulphuric acid. It uses these reactions to discuss
the trend in reducing ability of the ions as you go from fluoride to chloride
to bromide to iodide.

The Facts
There are two different types of reaction which might go on when
concentrated sulphuric acid is added to a solid ionic halide like sodium
fluoride, chloride, bromide or iodide. The concentrated sulphuric acid can
act both as an acid and as an oxidising agent.
Concentrated sulphuric acid acting as an acid
The concentrated sulphuric acid gives a hydrogen ion to the halide ion to
produce a hydrogen halide. Because this is a gas, it immediately
escapes from the system. If the hydrogen halide is exposed to moist air,
you see it as steamy fumes.
As an example, concentrated sulphuric acid reacts with solid sodium
chloride in the cold to produce hydrogen chloride and sodium
hydrogensulphate.

All of the halide ions (fluoride, chloride, bromide and iodide) behave
similarly.

http://www.chemguide.co.uk/inorganic/group7/halideions.html (1 of 12)30/12/2004 11:14:33

redox reactions involving halide ions and sulphuric acid

Note: These reactions to make the hydrogen halides are


dealt with on a separate page.
If you want to read a bit more about them, follow this link and
use the BACK button on your browser to return to this page.

Concentrated sulphuric acid acting as an oxidising agent


With fluoride or chloride ions
Concentrated sulphuric acid isn't a strong enough oxidising agent to
oxidise fluoride or chloride ions. In those cases, all you get produced are
the steamy fumes of the hydrogen halide - hydrogen fluoride or hydrogen
chloride.
You can look at this another way - from the point of view of the halide
ions. The fluoride and chloride ions aren't strong enough reducing agents
to reduce the sulphuric acid.
Whichever way you look at it, all you get is the hydrogen halide!
That isn't true, though, with bromides and iodides.
With bromide ions
The bromide ions are strong enough reducing agents to reduce the
concentrated sulphuric acid. In the process the bromide ions are oxidised
to bromine.

The bromide ions reduce the sulphuric acid to sulphur dioxide gas. This
is a decrease of oxidation state of the sulphur from +6 in the sulphuric
acid to +4 in the sulphur dioxide.

http://www.chemguide.co.uk/inorganic/group7/halideions.html (2 of 12)30/12/2004 11:14:33

redox reactions involving halide ions and sulphuric acid

You can combine these two half-equations to give the overall ionic
equation for the reaction:

Note: If you aren't confident about redox reactions, electronhalf equations, and oxidation states you really ought to follow
this link before you go any further.

What you see in this reaction are the steamy fumes of hydrogen bromide
contaminated with the brown colour of bromine vapour. The sulphur
dioxide is a colourless gas, so you couldn't observe its presence directly.
With iodide ions
Iodide ions are stronger reducing agents than bromide ions are. They are
oxidised to iodine by the concentrated sulphuric acid.

The reduction of the sulphuric acid is more complicated than before. The
iodide ions are powerful enough reducing agents to reduce it

first to sulphur dioxide (sulphur oxidation state = +4)

then to sulphur itself (oxidation state = 0)

and all the way to hydrogen sulphide (sulphur oxidation state = -2).

The most important of this mixture of reduction products is probably the


hydrogen sulphide. The half-equation for its formation is:

Combining these last two half-equations gives:

http://www.chemguide.co.uk/inorganic/group7/halideions.html (3 of 12)30/12/2004 11:14:33

redox reactions involving halide ions and sulphuric acid

Important! Don't try to remember this equation - the


chances of you ever needing it in an exam are tiny. Learn
how to work out electron-half-equations and combine them to
make the overall equation. A bit of time acquiring that skill will
save you a lot of pointless learning.

This time what you see is a trace of steamy fumes of hydrogen iodide,
but mainly lots of iodine. The reaction is exothermic and so purple iodine
vapour is formed, and probably dark grey solid iodine condensing around
the top of the tube. There will also be red colours where the iodine
comes into contact with the solid iodide.
The red colour is due to the I3- ion formed by reaction between I2
molecules and I- ions.
You won't see the colourless hydrogen sulphide gas, but might pick up its
"bad egg" smell if you were foolish enough to smell the intensely
poisonous gases evolved!

Summary of the trend in reducing ability

Fluoride and chloride ions won't reduce concentrated sulphuric


acid.
Bromide ions reduce the sulphuric acid to sulphur dioxide. In the
process, the bromide ions are oxidised to bromine.
Iodide ions reduce the sulphuric acid to a mixture of products
including hydrogen sulphide. The iodide ions are oxidised to iodine.
Reducing ability of the halide ions increases as you go down the
Group.

Explaining the trend


http://www.chemguide.co.uk/inorganic/group7/halideions.html (4 of 12)30/12/2004 11:14:33

redox reactions involving halide ions and sulphuric acid

An over-simplified explanation
This only works (and even then, not very well!) if you ignore fluoride ions.
The argument goes like this:
When a halide ion acts as a reducing agent, it gives electrons to
something else. That means that the halide ion itself has to lose
electrons.
The bigger the halide ion, the further the outer electrons are from the
nucleus, and the more they are screened from it by inner electrons. It
therefore gets easier for the halide ions to lose electrons as you go down
the Group because there is less attraction between the outer electrons
and the nucleus.
It sounds convincing, but it only tells part of the story. We need to look in
some detail at the energetics of the change.

Important! You really need to find out what (if any)


explanation your examiners expect you to give for this. If their
mark schemes (or the way they phrase their questions)
suggest that they want this simplified explanation, then that's
what you will have to give them.
The rest of this page is going to get quite complicated. It
would be worth while finding out whether it is something you
need to know. (Although it is always more satisfying the
closer you can get to the truth!)
UK A' level students should search their syllabuses, past
exam papers, mark schemes and any other support material
available from their Exam Board. If you haven't got any of
this, you can find your Exam Board's web address by
following this link. Students elsewhere should find out the
equivalent information from their own sources.

http://www.chemguide.co.uk/inorganic/group7/halideions.html (5 of 12)30/12/2004 11:14:33

redox reactions involving halide ions and sulphuric acid

A more detailed explanation


Looking at how the enthalpy changes vary from halogen to halogen
We need to compare the amount of heat evolved or absorbed when you
convert a solid halide (like sodium chloride) into molecules of the halogen.
Taking sodium chloride as an example:

We need to supply the energy to break the attractions between the


ions in the sodium chloride. In other words, we need to supply the
lattice enthalpy.
We need to supply the energy to remove the electron from the
chloride ion. This is the reverse of the electron affinity of the
chlorine. You can get this figure by looking up the electron affinity
in a Data Book and giving it a positive rather than a negative sign.
We then recover some energy when the chlorine atoms turn into
chlorine molecules. Energy is released when the bonds are formed.
Chlorine is simple because it is a gas. With bromine and iodine,
heat will also be released when they condense to a liquid or solid.
To take account of this, it is better to think of this in terms of
atomisation energy rather than bond energy. The number we want
is the reverse of atomisation energy.
Atomisation energy is the energy needed to produce 1 mole of
isolated gaseous atoms starting from an element in its standard
state (gas for chlorine, and liquid for bromine, for example - both of
them as X2).

Look carefully at the diagram so that you see how this all fits together:

http://www.chemguide.co.uk/inorganic/group7/halideions.html (6 of 12)30/12/2004 11:14:33

redox reactions involving halide ions and sulphuric acid

Note: The term "lattice enthalpy" used here should more


accurately be described as "lattice dissociation enthalpy".
If you aren't confident about energy cycles and the logic
behind them (Hess's Law), you might be interested in my
chemistry calculations book where they are explained in a lot
of detail.

http://www.chemguide.co.uk/inorganic/group7/halideions.html (7 of 12)30/12/2004 11:14:33

redox reactions involving halide ions and sulphuric acid

What we need to do is calculate the enthalpy change shown by the green


arrow in the diagram for each of the halogens so that we can make a
comparison. The diagram shows that the overall change involving the
halide ions is endothermic - the green arrow is pointing upwards towards
a higher energy.
This isn't the total enthalpy change for the whole reaction. Heat will be
given out when the changes involving the sulphuric acid occur. That will
be the same irrespective of which halogen you are talking about. The
total enthalpy change will be the sum of the enthalpy changes for the
halide ion half-reaction and the sulphuric acid half-reaction.
The table shows the energy changes which vary from halogen to
halogen. We are assuming that you start from solid sodium halide. The
values for the lattice enthalpies for other solid halides would be different,
but the pattern will still be the same.

heat needed
to break up
NaX lattice
(kJ mol-1)

heat needed
to remove
electron from
halide ion
(kJ mol-1)

heat released in
forming
halogen
molecules
(kJ mol-1)

sum of
these
(kJ mol-1)

+902

+328

-79

+1151

Cl

+771

+349

-121

+999

Br

+733

+324

-112

+945

+684

+295

-107

+872

Note: There is likely to be some error in these figures. They


come from a variety of sources - some more reliable than
others!

http://www.chemguide.co.uk/inorganic/group7/halideions.html (8 of 12)30/12/2004 11:14:33

redox reactions involving halide ions and sulphuric acid

The overall enthalpy change for the halide half-reaction:


Look at the final column of figures.
Notice that the sum of these enthalpy changes gets less endothermic as
you go down the Group. That means that the total change (including the
sulphuric acid) will become easier as you go down the Group.
The amount of heat given out by the half-reaction involving the sulphuric
acid must be great enough to make the reactions with the bromide or
iodide feasible, but not enough to compensate for the more positive
values produced by the fluoride and chloride half-reactions.
I don't know what the real value for the sulphuric acid half-reaction to
produce sulphur dioxide is, but it must be something like -980 kJ mol-1.
Try the effect of combining that value with the overall values in the table
to see what happens to the total enthalpy change of reaction for each
halogen.

Exploring the changes in the various energy terms


Which individual energy terms in the table are most important in making
the halogen half-reaction less endothermic as you go down the Group?
Chlorine to iodine
Considering the halogens from chlorine to iodine, it is the lattice enthalpy
which has fallen most. It falls by 87 kJ mol-1. By contrast, the heat
needed to remove the electron has only fallen by 54 kJ mol-1.
Both of these terms matter, but the fall in lattice enthalpy is the more
important. This falls because the ions are getting bigger. That means that
they aren't as close to each other, and so the attractions between
positive and negative ions in the solid lattice get less.
The simplified explanation that we mentioned earlier concentrates on the
less important fall in the amount of energy needed to remove the electron
from the ion. That's misleading!

http://www.chemguide.co.uk/inorganic/group7/halideions.html (9 of 12)30/12/2004 11:14:33

redox reactions involving halide ions and sulphuric acid

Fluorine
Fluoride ions are very difficult to oxidise to fluorine. The table shows that
this isn't anything to do with the amount of energy needed to remove an
electron from a fluoride ion. It is actually easier to remove an electron
from a fluoride ion than from a chloride ion. In this case, to make the
generalisation that an electron gets easier to remove as the ion gets
bigger is just plain wrong!
Fluoride ions are so small that the electrons feel an abnormal amount of
repulsion from each other. This outweighs the effect of their closeness to
the nucleus and makes them easier to remove than you might expect.
There are two important reasons why fluoride ions are so difficult to
oxidise.
The first is the comparatively very high lattice enthalpy of the solid
fluoride. This is due to the small size of the fluoride ion, which means that
the positive and negative ions are very close together and so strongly
attracted to each other.
The other factor is the small amount of heat which is released when the
fluorine atoms combine to make fluorine molecules. (Scroll back and look
at the table again.)
This is because of the low bond enthalpy of the F-F bond. The reason for
this low bond enthalpy is discusssed on a separate page.

Note: If you haven't read about this recently, you will find it
on the page about atomic and physical properties of the
halogens

http://www.chemguide.co.uk/inorganic/group7/halideions.html (10 of 12)30/12/2004 11:14:33

redox reactions involving halide ions and sulphuric acid

What if the halide ions were in solution rather than in a solid?


We have concentrated on the energetics of the process starting from
solid halide ions because that's what you use if you try to oxidise them
using concentrated sulphuric acid. What about oxidising them in solution
using some different oxidising agent?
The trend is exactly the same. Fluoride ions are difficult to oxidise and it
gets easier as you go down the Group towards iodide ions. Looked at
another way, fluoride ions aren't good reducing agents, but iodide ions
are.
The explanation this time has to start from the hydrated ions in solution
rather than solid ions. In a sense, this has already been done on another
page.
Fluorine is a very powerful oxidising agent because it very readily forms
its negative ion in solution. That means that it will be energetically difficult
to reverse the process.
By contrast, for the energetic reasons you will find discussed, iodine is
relatively reluctant to form its negative ion in solution. That means that it
will be relatively easy to persuade it to revert to iodine molecules again.

Note: You will find the oxidising ability of the halogens


explained in detail by following this link.
Because you would now be thinking about the reverse of the
processes described on that page, you will have to reverse
the sign of all the energy changes explored. If I were you, I
wouldn't bother to follow this up unless there is some
overwhelming reason to!

http://www.chemguide.co.uk/inorganic/group7/halideions.html (11 of 12)30/12/2004 11:14:33

redox reactions involving halide ions and sulphuric acid

Where would you like to go now?


To the Group 7 menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/inorganic/group7/halideions.html (12 of 12)30/12/2004 11:14:33

Testing for halide ions

TESTING FOR HALIDE IONS

This page describes and explains the tests for halide ions (fluoride,
chloride, bromide and iodide) using silver nitrate solution followed by
ammonia solution.

Using silver nitrate solution


Carrying out the test
This test has to be done in solution. If you start from a solid, it must first
be dissolved in pure water.
The solution is acidified by adding dilute nitric acid. (Remember: silver
nitrate + dilute nitric acid.) The nitric acid reacts with, and removes,
other ions that might also give a confusing precipitate with silver nitrate.
Silver nitrate solution is then added to give:

ion present

observation

F-

no precipitate

Cl-

white precipitate

Br-

very pale cream precipitate

I-

very pale yellow precipitate

The chloride, bromide and iodide precipitates are shown in the


photograph:

http://www.chemguide.co.uk/inorganic/group7/testing.html (1 of 7)30/12/2004 11:14:39

Testing for halide ions

The chloride precipitate is obviously white, but the other two aren't really
very different from each other. You couldn't be sure which you had
unless you compared them side-by-side.
All of the precipitates change colour if they are exposed to light - taking
on grey or purplish tints.
The absence of a precipitate with fluoride ions doesn't prove anything
unless you already know that you must have a halogen present and are
simply trying to find out which one. All the absence of a precipitate shows
is that you haven't got chloride, bromide or iodide ions present.
The chemistry of the test
The precipitates are the insoluble silver halides - silver chloride, silver
bromide or silver iodide.

Silver fluoride is soluble, and so you don't get a precipitate.

http://www.chemguide.co.uk/inorganic/group7/testing.html (2 of 7)30/12/2004 11:14:39

Testing for halide ions

Confirming the precipitate using ammonia solution


Carrying out the confirmation
Ammonia solution is added to the precipitates.

original precipitate

observation

AgCl

precipitate dissolves to give a colourless


solution

AgBr

precipitate is almost unchanged using dilute


ammonia solution, but dissolves in
concentrated ammonia solution to give a
colourless solution

AgI

precipitate is insoluble in ammonia solution


of any concentration

Explaining what happens


Background
There is no such thing as an absolutely insoluble ionic compound. A
precipitate will only form if the concentrations of the ions in solution in
water exceed a certain value - different for every different compound.
This value can be quoted as a solubility product. For the silver halides,
the solubility product is given by the expression:
Ksp = [Ag+(aq)][X-(aq)]
The square brackets have their normal meaning, showing concentrations
in mol dm-3.
If the actual concentrations of the ions in solution produce a value less
http://www.chemguide.co.uk/inorganic/group7/testing.html (3 of 7)30/12/2004 11:14:39

Testing for halide ions

than the solubility product, you don't get a precipitate. If the product of
the concentrations would exceed this value, you do get a precipitate.
Essentially, the product of the ionic concentrations can never be greater
than the solubility product value. Enough of the solid is precipitated so
that the ionic product is lowered to the value of the solubility product.

Note: If your syllabus says that you need to know about


solubility product calculations, you might be interested in my
chemistry calculations book where they are explained in
detail.

Look at the way the solubility products vary from silver chloride to silver
iodide. (You can't quote a solubility product value for silver fluoride
because it is too soluble. Solubility products only work with compounds
which are very, very sparingly soluble.)

Ksp
(mol2dm-6)
AgCl

1.8 x 10-10

AgBr

7.7 x 10-13

AgI

8.3 x 10-17

Note: These figures come from the Chemistry Data Book by


Stark and Wallace.

http://www.chemguide.co.uk/inorganic/group7/testing.html (4 of 7)30/12/2004 11:14:39

Testing for halide ions

You can see that the compounds are all pretty insoluble, but become
even less soluble as you go from the chloride to the bromide to the iodide.
What is the ammonia doing?
The ammonia combines with silver ions to produce a complex ion called
the diamminesilver(I) ion, [Ag(NH3)2]+. This is a reversible reaction, but
the complex is very stable, and the position of equilibrium lies well to the
right.

A solution in contact with one of the silver halide precipitates will contain
a very small concentration of dissolved silver ions. The effect of adding
the ammonia is to lower this concentration still further.
What happens if you multiply this new silver ion concentration by the
halide ion concentration? If the answer is less than the solubility product,
the precipitate will dissolve.
That happens with the silver chloride, and with the silver bromide if
concentrated ammonia is used. The more concentrated ammonia tips the
equilibrium even further to the right, lowering the silver ion concentration
even more.
The silver iodide is so insoluble that the ammonia won't lower the silver
ion concentration enough for the precipitate to dissolve.

An alternative test using concentrated sulphuric acid


If you add concentrated sulphuric acid to a solid sample of one of the
halides you get these results:

http://www.chemguide.co.uk/inorganic/group7/testing.html (5 of 7)30/12/2004 11:14:39

Testing for halide ions

ion present

observation

F-

steamy acidic fumes (of HF)

Cl-

steamy acidic fumes (of HCl)

Br-

steamy acidic fumes (of HBr) contaminated with


brown bromine vapour

I-

Some steamy fumes (of HI), but lots of purple iodine


vapour (plus various red colours in the tube)

Note: The chemistry of this test is explained in detail on


another page.

The only possible confusion is between a fluoride and a chloride - they


would behave identically. You could distinguish between them by
dissolving the original solid in water and then testing with silver nitrate
solution. The chloride gives a white precipitate; the fluoride doesn't give a
precipitate.

Where would you like to go now?


To the Group 7 menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002


http://www.chemguide.co.uk/inorganic/group7/testing.html (6 of 7)30/12/2004 11:14:39

Testing for halide ions

http://www.chemguide.co.uk/inorganic/group7/testing.html (7 of 7)30/12/2004 11:14:39

Complex ions menu

Understanding Chemistry

COMPLEX IONS MENU

What is a complex ion? . . .


An introduction to complex ions with an explanation of what
ligands are and how they bond to the central metal ion.
Naming complex ions . . .
An introduction to the naming of common complex ions.
Shapes of complex ions . . .
An introduction to the shapes of complex ions (including some
simple stereoisomerism).
The origin of colour in complex ions . . .
A simple explanation of why complex ions formed by transition
metals tend to be coloured, and a look at the factors which cause
the colours to change during reactions.
Ligand exchange reactions (introduction) . . .
Looks at the chemistry of several simple ligand exchange
reactions required for UK A' level courses.
Ligand exchange reactions (stability constants) . . .
Explains what is meant by stability constants for complex ions.
This page assumes that you are familiar with simple ligand
exchange reactions.
http://www.chemguide.co.uk/inorganic/complexmenu.html (1 of 2)30/12/2004 11:14:42

Complex ions menu

The acidity of the hexaaqua ions . . .


Explains why ions of the type M(H2O)6n+ are acidic, and how and
why the acidity varies with the charge on the ion.
The reactions of the hexaaqua ions with hydroxide ions . . .
Looks in detail at what happens when you add sodium hydroxide
solution to solutions of the hexaaqua ions.
The reactions of the hexaaqua ions with ammonia . . .
Looks in detail at what happens when you add ammonia solution
to solutions of the hexaaqua ions.
The reactions of the hexaaqua ions with carbonate ions . . .
Looks in detail at what happens when you add sodium carbonate
solution to solutions of the hexaaqua ions.

Go to inorganic chemistry menu . . .


Go to Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/inorganic/complexmenu.html (2 of 2)30/12/2004 11:14:42

introducing complex ions - ligands and bonding

AN INTRODUCTION TO COMPLEX METAL IONS

This page explains the terms complex ion and ligand, and looks at the
bonding between the ligands and the central metal ion. It discusses
various sorts of ligand (including some quite complicated ones), and
describes what is meant by co-ordination number.

Complex metal ions containing simple ligands


What is a complex metal ion?
A complex ion has a metal ion at its centre with a number of other
molecules or ions surrounding it. These can be considered to be
attached to the central ion by co-ordinate (dative covalent) bonds. (In
some cases, the bonding is actually more complicated than that.)

Note: If you aren't sure about co-ordinate (dative covalent)


bonding, you aren't going to make much sense of what
follows without first following this link. Amongst other
examples of co-ordinate bonding, that page contains a
description of the bonding in the complex ion formed between
aluminium ions and water molecules, and that will be
repeated below on this page - so you needn't spend a lot of
time reading that bit.
If you follow this link, use the BACK button on your browser
to return quickly to this page.

http://www.chemguide.co.uk/inorganic/complexions/whatis.html (1 of 15)30/12/2004 11:15:06

introducing complex ions - ligands and bonding

The molecules or ions surrounding the central metal ion are called
ligands.
The nature of ligands
Simple ligands include water, ammonia and chloride ions.

What all these have got in common is active lone pairs of electrons in the
outer energy level. These are used to form co-ordinate bonds with the
metal ion.
All ligands are lone pair donors. In other words, all ligands function as
Lewis bases.

Note: If you haven't come across the term Lewis base, and
want to find out more, you could follow this link to a page on
theories of acids and bases. It isn't, however, particularly
important to the rest of this page that you know anything
more than the fact that a substance which forms a coordinate bond by donating a lone pair of electrons to
something else is known as a Lewis base.
If you follow this link, use the BACK button on your browser
to return quickly to this page.

http://www.chemguide.co.uk/inorganic/complexions/whatis.html (2 of 15)30/12/2004 11:15:06

introducing complex ions - ligands and bonding

Bonding in simple complex ions


Al(H2O)6 3+
We are going to look in detail at the bonding in the complex ion formed
when water molecules attach themselves to an aluminium ion to give Al
(H2O)63+.
Start by thinking about the structure of a naked aluminium ion before the
water molecules bond to it.
Aluminium has the electronic structure
1s22s22p63s23px1
When it forms an Al3+ ion it loses the 3-level electrons to leave
1s22s22p6

Warning: It is a complete waste of time going any further


with this page if you aren't confident about writing electronic
structures in this form for elements and ions (including the
first transition series). Explore the atomic properties menu,
following the links to atomic orbitals and electronic structures.
Come back to this page later using the BACK button, the
History file, or the Go menu on your browser.

http://www.chemguide.co.uk/inorganic/complexions/whatis.html (3 of 15)30/12/2004 11:15:06

introducing complex ions - ligands and bonding

That means that all the 3-level orbitals are now empty. The aluminium
uses of six of these to accept lone pairs from six water molecules.
It re-organises (hybridises) the 3s, the three 3p, and two of the 3d
orbitals to produce six new orbitals all with the same energy.
You might wonder why it chooses to use six orbitals rather than four or
eight or whatever. Six is the maximum number of water molecules it is
possible to fit around an aluminium ion (and most other metal ions). By
making the maximum number of bonds, it releases most energy and so
becomes most energetically stable.

Only one lone pair is shown on each water molecule. The other lone pair
is pointing away from the aluminium and so isn't involved in the bonding.
The resulting ion looks like this:

Note: Dotted arrows represent lone pairs coming from water


molecules behind the plane of the screen or paper. Wedge
shaped arrows represent bonds from water molecules in front
of the plane of the screen or paper.

http://www.chemguide.co.uk/inorganic/complexions/whatis.html (4 of 15)30/12/2004 11:15:06

introducing complex ions - ligands and bonding

Because of the movement of electrons towards the centre of the ion, the
3+ charge is no longer located entirely on the aluminium, but is now
spread over the whole of the ion.
Because the aluminium is forming 6 bonds, the co-ordination number
of the aluminium is said to be 6. The co-ordination number of a complex
ion counts the number of co-ordinate bonds being formed by the metal
ion at its centre.
In a simple case like this, that obviously also counts the number of
ligands - but that isn't necessarily so, as you will see later. Some ligands
can form more than one co-ordinate bond with the metal ion.

Fe(H2O)6 3+
This example is chosen because it is very similar to the last one - except
that it involves a transition metal.
Iron has the electronic structure
1s22s22p63s23p63d64s2
When it forms an Fe3+ ion it loses the 4s electrons and one of the 3d
electrons to leave
1s22s22p63s23p63d5
Looking at this as electrons-in-boxes, at the bonding level:

http://www.chemguide.co.uk/inorganic/complexions/whatis.html (5 of 15)30/12/2004 11:15:06

introducing complex ions - ligands and bonding

Now, be careful! The single electrons in the 3d level are NOT involved in
the bonding in any way. Instead, the ion uses 6 orbitals from the 4s, 4p
and 4d levels to accept lone pairs from the water molecules.
Before they are used, the orbitals are re-organised (hybridised) to
produce 6 orbitals of equal energy.

Once the co-ordinate bonds have been formed, the ion looks exactly the
same as the equivalent aluminium ion.

Because the iron is forming 6 bonds, the co-ordination number of the iron
http://www.chemguide.co.uk/inorganic/complexions/whatis.html (6 of 15)30/12/2004 11:15:06

introducing complex ions - ligands and bonding

is 6.

Warning: This example was deliberately chosen to avoid a


complication. With many ligands, the electrons in the 3d
levels get moved around before the co-ordinate bonds are
formed.
For example, if you use cyanide ions, CN-, as ligands around
an iron(III) ion, the d electrons end up as 2 pairs plus a single
one - instead of all being single. Because this produces two
empty 3d orbitals, it affects which orbitals are hybridised
ready to form bonds with the ligands.
You don't need to worry about this problem for A' level
purposes. All you need to know is that ligands bond by
forming co-ordinate bonds using available empty orbitals in
the metal ion. You don't need to know exactly what these
orbitals are.

CuCl4 2This is a simple example of the formation of a complex ion with a


negative charge.
Copper has the electronic structure
1s22s22p63s23p63d104s1
When it forms a Cu2+ ion it loses the 4s electron and one of the 3d
electrons to leave
1s22s22p63s23p63d9
To bond the four chloride ions as ligands, the empty 4s and 4p orbitals
are used (in a hybridised form) to accept a lone pair of electrons from
each chloride ion. Because chloride ions are bigger than water
molecules, you can't fit 6 of them around the central ion - that's why you
only use 4.
http://www.chemguide.co.uk/inorganic/complexions/whatis.html (7 of 15)30/12/2004 11:15:06

introducing complex ions - ligands and bonding

Only one of the 4 lone pairs on each chloride ion is shown. The other
three are pointing away from the copper ion, and aren't involved in the
bonding.
That gives you the complex ion:

The ion carries 2 negative charges overall. That comes from a


combination of the 2 positive charges on the copper ion and the 4
negative charges from the 4 chloride ions.
In this case, the co-ordination number of the copper is, of course, 4.

Complex metal ions containing more complicated ligands


http://www.chemguide.co.uk/inorganic/complexions/whatis.html (8 of 15)30/12/2004 11:15:06

introducing complex ions - ligands and bonding

In the examples we've already looked at, each ligand only forms one
bond with the central metal ion to give the complex ion. Such a ligand is
said to be unidentate. That means literally that it only has one tooth!
It only has one pair of electrons that it can use to bond to the metal - any
other lone pairs are pointing in the wrong direction.
Some ligands, however, have rather more teeth! These are known
generally as multidentate or polydentate ligands, but can be broken down
into a number of different types.
Bidentate ligands
Bidentate ligands have two lone pairs, both of which can bond to the
central metal ion.
The two commonly used examples are 1,2-diaminoethane (old name:
ethylenediamine - often given the abbreviation "en"), and the
ethanedioate ion (old name: oxalate).

In the ethanedioate ion, there are lots more lone pairs than the two
shown, but these are the only ones we are interested in.
You can think of these bidentate ligands rather as if they were a pair of
headphones, carrying lone pairs on each of the "ear pieces". These will
then fit snuggly around a metal ion.

http://www.chemguide.co.uk/inorganic/complexions/whatis.html (9 of 15)30/12/2004 11:15:06

introducing complex ions - ligands and bonding

Ni (NH2CH2CH2NH2)3 2+
You might find this abbreviated to [Ni(en)3] 2+.

Note: It is quite common to wrap square brackets around


complex ions to emphasise that the charge is carried by the
whole ion - as I have done in the abbreviated form, and in the
diagram below.
I have omitted them from the formula in the heading above
because they didn't come out well in the italic font I used. It
just confused things.

The structure of the ion looks like this:

http://www.chemguide.co.uk/inorganic/complexions/whatis.html (10 of 15)30/12/2004 11:15:06

introducing complex ions - ligands and bonding

In this case, the "ear pieces" are the nitrogen atoms of the NH2 groups and the "bit that goes over your head " is the -CH2CH2- group. If you
were going to draw this in an exam, you would obviously want to draw it
properly - but for learning purposes, drawing all the atoms makes the
diagram look unduly complicated!
Notice that the arrangement of the bonds around the central metal ion is
exactly the same as it was with the ions with 6 water molecules attached.
The only difference is that this time each ligand uses up two of the
positions - at right angles to each other.
Because the nickel is forming 6 co-ordinate bonds, the co-ordination
number of this ion is 6, despite the fact that it is only joined to 3 ligands.
Co-ordination number counts the number of bonds, not the number of
ligands.
Cr (C2O4)3 3This is the complex ion formed by attaching 3 ethanedioate (oxalate) ions
to a chromium(III) ion. The shape is exactly the same as the previous
nickel complex. The only real difference is the number of charges. The
original chromium ion carried 3+ charges, and each ethanedioate ion
carried 2-. (3+) + (3 x 2-) = 3-.
The structure of the ion looks like this:

Again, if you drew this in an exam, you would want to show all the atoms
properly. If you need to be able to do this, practice drawing it so that it
looks clear and tidy! Refer back to the diagram of the ethanedioate ion
further up the page to help you.

http://www.chemguide.co.uk/inorganic/complexions/whatis.html (11 of 15)30/12/2004 11:15:06

introducing complex ions - ligands and bonding

A quadridentate ligand
A quadridentate ligand has four lone pairs, all of which can bond to the
central metal ion.
An example of this occurs in haemoglobin (American: hemoglobin).
The functional part of this is an iron(II) ion surrounded by a complicated
molecule called haem (heme). This is a sort of hollow ring of carbon and
hydrogen atoms, at the centre of which are 4 nitrogen atoms with lone
pairs on them.
Haem is one of a group of similar compounds called porphyrins. They
all have the same sort of ring system, but with different groups attached
to the outside of the ring. You aren't going to need to know the exact
structure of the haem at this level.
We could simplify the haem with the trapped iron ion as:

Each of the lone pairs on the nitrogen can form a co-ordinate bond with
the iron(II) ion - holding it at the centre of the complicated ring of atoms.
The iron forms 4 co-ordinate bonds with the haem, but still has space to
form two more - one above and one below the plane of the ring.
The protein globin attaches to one of these positions using a lone pair
on one of the nitrogens in one of its amino acids. The interesting bit is the
other position.

http://www.chemguide.co.uk/inorganic/complexions/whatis.html (12 of 15)30/12/2004 11:15:06

introducing complex ions - ligands and bonding

Overall, the complex ion has a co-ordination number of 6 because the


central metal ion is forming 6 co-ordinate bonds.
The water molecule which is bonded to the bottom position in the
diagram is easily replaced by an oxygen molecule (again via a lone pair
on one of the oxygens in O2) - and this is how oxygen gets carried
around the blood by the haemoglobin.
When the oxygen gets to where it is needed, it breaks away from the
haemoglobin which returns to the lungs to get some more.
You probably know that carbon monoxide is poisonous because it reacts
with haemoglobin. It bonds to the same site that would otherwise be
used by the oxygen - but it forms a very stable complex. The carbon
monoxide doesn't break away again, and that makes that haemoglobin
molecule useless for any further oxygen transfer.

A hexadentate ligand
A hexadentate ligand has 6 lone pairs of electrons - all of which can form
co-ordinate bonds with the same metal ion. The best example is EDTA.
EDTA is used as a negative ion - EDTA4-. The diagram shows the
http://www.chemguide.co.uk/inorganic/complexions/whatis.html (13 of 15)30/12/2004 11:15:06

introducing complex ions - ligands and bonding

structure of the ion with the important atoms and lone pairs picked out.

Note: The abbreviation EDTA comes from an old name for


the parent acid of this - one where each of the negatively
charged oxygens has a hydrogen atom attached. This used
to be called ethylenediaminetetraacetic acid. Nobody ever
calls it anything other than EDTA!

The EDTA ion entirely wraps up a metal ion using all 6 of the positions
that we have seen before. The co-ordination number is again 6 because
of the 6 co-ordinate bonds being formed by the central metal ion. The
diagram below shows this happening with a copper(II) ion.
Drawing the product of this clearly while showing all the atoms defeats
me completely! Here is a simplified version. Make sure that you can see
how this relates to the full structure above.

The overall charge, of course, comes from the 2+ on the original copper
http://www.chemguide.co.uk/inorganic/complexions/whatis.html (14 of 15)30/12/2004 11:15:06

introducing complex ions - ligands and bonding

(II) ion and the 4- on the EDTA4- ion.

Where would you like to go now?


To the complex ion menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/inorganic/complexions/whatis.html (15 of 15)30/12/2004 11:15:06

Theories of acids and bases

THEORIES OF ACIDS AND BASES

This page describes the Arrhenius, Bronsted-Lowry, and Lewis theories


of acids and bases, and explains the relationships between them. It also
explains the concept of a conjugate pair - an acid and its conjugate base,
or a base and its conjugate acid.

Note: Current UK A' level syllabuses concentrate on the


Bronsted-Lowry theory, but you should also be aware of
Lewis acids and bases. The Arrhenius theory is of historical
interest only, and you are unlikely to need it unless you are
doing some work on the development of ideas in chemistry.

The Arrhenius Theory of acids and bases


The theory

Acids are substances which produce hydrogen ions in solution.

Bases are substances which produce hydroxide ions in solution.

Neutralisation happens because hydrogen ions and hydroxide ions react


to produce water.

Limitations of the theory


Hydrochloric acid is neutralised by both sodium hydroxide solution and
ammonia solution. In both cases, you get a colourless solution which you
can crystallise to get a white salt - either sodium chloride or ammonium
chloride.
These are clearly very similar reactions. The full equations are:

http://www.chemguide.co.uk/physical/acidbaseeqia/theories.html (1 of 14)30/12/2004 11:15:20

Theories of acids and bases

In the sodium hydroxide case, hydrogen ions from the acid are reacting
with hydroxide ions from the sodium hydroxide - in line with the Arrhenius
theory.
However, in the ammonia case, there don't appear to be any hydroxide
ions!
You can get around this by saying that the ammonia reacts with the
water it is dissolved in to produce ammonium ions and hydroxide ions:

This is a reversible reaction, and in a typical dilute ammonia solution,


about 99% of the ammonia remains as ammonia molecules.
Nevertheless, there are hydroxide ions there, and we can squeeze this
into the Arrhenius theory.
However, this same reaction also happens between ammonia gas and
hydrogen chloride gas.

In this case, there aren't any hydrogen ions or hydroxide ions in solution because there isn't any solution. The Arrhenius theory wouldn't count this
as an acid-base reaction, despite the fact that it is producing the same
product as when the two substances were in solution. That's silly!

The Bronsted-Lowry Theory of acids and bases


The theory

An acid is a proton (hydrogen ion) donor.

A base is a proton (hydrogen ion) acceptor.

http://www.chemguide.co.uk/physical/acidbaseeqia/theories.html (2 of 14)30/12/2004 11:15:20

Theories of acids and bases

The relationship between the Bronsted-Lowry theory and the


Arrhenius theory
The Bronsted-Lowry theory doesn't go against the Arrhenius theory in
any way - it just adds to it.
Hydroxide ions are still bases because they accept hydrogen ions from
acids and form water.
An acid produces hydrogen ions in solution because it reacts with the
water molecules by giving a proton to them.
When hydrogen chloride gas dissolves in water to produce hydrochloric
acid, the hydrogen chloride molecule gives a proton (a hydrogen ion) to a
water molecule. A co-ordinate (dative covalent) bond is formed between
one of the lone pairs on the oxygen and the hydrogen from the HCl.
Hydroxonium ions, H3O+, are produced.

Note: If you aren't sure about co-ordinate bonding you


should follow this link. Co-ordinate bonds will be mentioned
several times over the course of the rest of this page.
Use the BACK button on your browser to return quickly to this
page.

http://www.chemguide.co.uk/physical/acidbaseeqia/theories.html (3 of 14)30/12/2004 11:15:20

Theories of acids and bases

When an acid in solution reacts with a base, what is actually functioning


as the acid is the hydroxonium ion. For example, a proton is transferred
from a hydroxonium ion to a hydroxide ion to make water.

Showing the electrons, but leaving out the inner ones:

It is important to realise that whenever you talk about hydrogen ions in


solution, H+(aq), what you are actually talking about are hydroxonium ions.
The hydrogen chloride / ammonia problem
This is no longer a problem using the Bronsted-Lowry theory. Whether
you are talking about the reaction in solution or in the gas state,
ammonia is a base because it accepts a proton (a hydrogen ion). The
hydrogen becomes attached to the lone pair on the nitrogen of the
ammonia via a co-ordinate bond.

If it is in solution, the ammonia accepts a proton from a hydroxonium ion:

http://www.chemguide.co.uk/physical/acidbaseeqia/theories.html (4 of 14)30/12/2004 11:15:20

Theories of acids and bases

If the reaction is happening in the gas state, the ammonia accepts a


proton directly from the hydrogen chloride:

Either way, the ammonia acts as a base by accepting a hydrogen ion


from an acid.

Conjugate pairs
When hydrogen chloride dissolves in water, almost 100% of it reacts with
the water to produce hydroxonium ions and chloride ions. Hydrogen
chloride is a strong acid, and we tend to write this as a one-way reaction:

Note: I am deliberately missing state symbols off this and


the next equation in order to concentrate on the bits that
matter.
You will find more about strong and weak acids on another
page in this section.

In fact, the reaction between HCl and water is reversible, but only to a
very minor extent. In order to generalise, consider an acid HA, and think
of the reaction as being reversible.

Thinking about the forward reaction:

The HA is an acid because it is donating a proton (hydrogen ion)


to the water.

http://www.chemguide.co.uk/physical/acidbaseeqia/theories.html (5 of 14)30/12/2004 11:15:20

Theories of acids and bases

The water is a base because it is accepting a proton from the HA.

But there is also a back reaction between the hydroxonium ion and the Aion:

The H3O+ is an acid because it is donating a proton (hydrogen ion)


to the A- ion.

The A- ion is a base because it is accepting a proton from the H3O


+.

The reversible reaction contains two acids and two bases. We think of
them in pairs, called conjugate pairs.

When the acid, HA, loses a proton it forms a base, A-. When the base,
A-, accepts a proton back again, it obviously refoms the acid, HA. These
two are a conjugate pair.
Members of a conjugate pair differ from each other by the presence or
absence of the transferable hydrogen ion.
If you are thinking about HA as the acid, then A- is its conjugate base.
If you are thinking about A- as the base, then HA is its conjugate acid.
The water and the hydroxonium ion are also a conjugate pair. Thinking of
the water as a base, the hydroxonium ion is its conjugate acid because it
has the extra hydrogen ion which it can give away again.
Thinking about the hydroxonium ion as an acid, then water is its
conjugate base. The water can accept a hydrogen ion back again to
reform the hydroxonium ion.
A second example of conjugate pairs
http://www.chemguide.co.uk/physical/acidbaseeqia/theories.html (6 of 14)30/12/2004 11:15:20

Theories of acids and bases

This is the reaction between ammonia and water that we looked at earlier:

Think first about the forward reaction. Ammonia is a base because it is


accepting hydrogen ions from the water. The ammonium ion is its
conjugate acid - it can release that hydrogen ion again to reform the
ammonia.
The water is acting as an acid, and its conjugate base is the hydroxide
ion. The hydroxide ion can accept a hydrogen ion to reform the water.
Looking at it from the other side, the ammonium ion is an acid, and
ammonia is its conjugate base. The hydroxide ion is a base and water is
its conjugate acid.
Amphoteric substances
You may possibly have noticed (although probably not!) that in one of the
last two examples, water was acting as a base, whereas in the other one
it was acting as an acid.
A substance which can act as either an acid or a base is described as
being amphoteric.

The Lewis Theory of acids and bases


This theory extends well beyond the things you normally think of as acids
and bases.

http://www.chemguide.co.uk/physical/acidbaseeqia/theories.html (7 of 14)30/12/2004 11:15:20

Theories of acids and bases

The theory

An acid is an electron pair acceptor.

A base is an electron pair donor.

The relationship between the Lewis theory and the Bronsted-Lowry


theory
Lewis bases
It is easiest to see the relationship by looking at exactly what BronstedLowry bases do when they accept hydrogen ions. Three Bronsted-Lowry
bases we've looked at are hydroxide ions, ammonia and water, and they
are typical of all the rest.

http://www.chemguide.co.uk/physical/acidbaseeqia/theories.html (8 of 14)30/12/2004 11:15:20

Theories of acids and bases

The Bronsted-Lowry theory says that they are acting as bases because
they are combining with hydrogen ions. The reason they are combining
with hydrogen ions is that they have lone pairs of electrons - which is
what the Lewis theory says. The two are entirely consistent.
So how does this extend the concept of a base? At the moment it doesn't
- it just looks at it from a different angle.
But what about other similar reactions of ammonia or water, for
example? On the Lewis theory, any reaction in which the ammonia or
water used their lone pairs of electrons to form a co-ordinate bond would
be counted as them acting as a base.
Here is a reaction which you will find talked about on the page dealing
http://www.chemguide.co.uk/physical/acidbaseeqia/theories.html (9 of 14)30/12/2004 11:15:20

Theories of acids and bases

with co-ordinate bonding. Ammonia reacts with BF3 by using its lone pair
to form a co-ordinate bond with the empty orbital on the boron.

As far as the ammonia is concerned, it is behaving exactly the same as


when it reacts with a hydrogen ion - it is using its lone pair to form a coordinate bond. If you are going to describe it as a base in one case, it
makes sense to describe it as one in the other case as well.

Note: If you haven't already read the page about co-ordinate


bonding you should do so now. You will find an important
example of water acting as a Lewis base as well as this
example - although the term Lewis base isn't used on that
page.
Use the BACK button on your browser to return quickly to this
page.

http://www.chemguide.co.uk/physical/acidbaseeqia/theories.html (10 of 14)30/12/2004 11:15:20

Theories of acids and bases

Lewis acids
Lewis acids are electron pair acceptors. In the above example, the BF3 is
acting as the Lewis acid by accepting the nitrogen's lone pair. On the
Bronsted-Lowry theory, the BF3 has nothing remotely acidic about it.
This is an extension of the term acid well beyond any common use.
What about more obviously acid-base reactions - like, for example, the
reaction between ammonia and hydrogen chloride gas?

What exactly is accepting the lone pair of electrons on the nitrogen.


Textbooks often write this as if the ammonia is donating its lone pair to a
hydrogen ion - a simple proton with no electrons around it.
That is misleading! You don't usually get free hydrogen ions in chemical
systems. They are so reactive that they are always attached to
something else. There aren't any uncombined hydrogen ions in HCl.
There isn't an empty orbital anywhere on the HCl which can accept a pair
of electrons. Why, then, is the HCl a Lewis acid?
Chlorine is more electronegative than hydrogen, and that means that the
hydrogen chloride will be a polar molecule. The electrons in the
hydrogen-chlorine bond will be attracted towards the chlorine end,
leaving the hydrogen slightly positive and the chlorine slightly negative.

http://www.chemguide.co.uk/physical/acidbaseeqia/theories.html (11 of 14)30/12/2004 11:15:20

Theories of acids and bases

Note: If you aren't sure about electronegativity and bond


polarity it might be useful to follow this link.
Use the BACK button on your browser to return quickly to this
page.

The lone pair on the nitrogen of an ammonia molecule is attracted to the


slightly positive hydrogen atom in the HCl. As it approaches it, the
electrons in the hydrogen-chlorine bond are repelled still further towards
the chlorine.
Eventually, a co-ordinate bond is formed between the nitrogen and the
hydrogen, and the chlorine breaks away as a chloride ion.
This is best shown using the "curly arrow" notation commonly used in
organic reaction mechanisms.

Note: If you aren't happy about the use of curly arrows to


show movements of electron pairs, you should follow this link.
Use the BACK button on your browser to return quickly to this
page.

http://www.chemguide.co.uk/physical/acidbaseeqia/theories.html (12 of 14)30/12/2004 11:15:20

Theories of acids and bases

The whole HCl molecule is acting as a Lewis acid. It is accepting a pair


of electrons from the ammonia, and in the process it breaks up. Lewis
acids don't necessarily have to have an existing empty orbital.

A final comment on Lewis acids and bases


If you are a UK A' level student, you might occasionally come across the
terms Lewis acid and Lewis base in textbooks or other sources. All you
need to remember is:

A Lewis acid is an electron pair acceptor.

A Lewis base is an electron pair donor.

Note: Remember this by thinking of ammonia acting as a


base. Most people at this level are familiar with the reactive
lone pair on the nitrogen accepting hydrogen ions. Ammonia
is basic because of its lone pair. That means that bases must
have lone pairs to donate. Acids are the opposite.

For all general purposes, stick with the Bronsted-Lowry theory.

Where would you like to go now?


To the acid-base equilibria menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

http://www.chemguide.co.uk/physical/acidbaseeqia/theories.html (13 of 14)30/12/2004 11:15:20

Theories of acids and bases

Jim Clark 2002

http://www.chemguide.co.uk/physical/acidbaseeqia/theories.html (14 of 14)30/12/2004 11:15:20

strong and weak acids

STRONG AND WEAK ACIDS

This page explains the terms strong and weak as applied to acids. As a
part of this it defines and explains what is meant by pH, Ka and pKa.
It is important that you don't confuse the words strong and weak with the
terms concentrated and dilute.
As you will see below, the strength of an acid is related to the proportion
of it which has reacted with water to produce ions. The concentration
tells you about how much of the original acid is dissolved in the solution.
It is perfectly possible to have a concentrated solution of a weak acid, or
a dilute solution of a strong acid. Read on . . .

Strong acids
Explaining the term "strong acid"
We are going to use the Bronsted-Lowry definition of an acid.
Note: If you don't know what the Bronsted-Lowry theory of
acids is, you should read about theories of acids and bases
on another page in this section. You don't need to spend time
reading about Lewis acids and bases for the purposes of this
present page.
Use the BACK button on your browser when you are ready to
return to this page.

http://www.chemguide.co.uk/physical/acidbaseeqia/acids.html (1 of 11)30/12/2004 11:15:33

strong and weak acids

When an acid dissolves in water, a proton (hydrogen ion) is transferred


to a water molecule to produce a hydroxonium ion and a negative ion
depending on what acid you are starting from.
In the general case . . .

These reactions are all reversible, but in some cases, the acid is so good
at giving away hydrogen ions that we can think of the reaction as being
one-way. The acid is virtually 100% ionised.
For example, when hydrogen chloride dissolves in water to make
hydrochloric acid, so little of the reverse reaction happens that we can
write:

At any one time, virtually 100% of the hydrogen chloride will have
reacted to produce hydroxonium ions and chloride ions. Hydrogen
chloride is described as a strong acid.
A strong acid is one which is virtually 100% ionised in solution.
Other common strong acids include sulphuric acid and nitric acid.
You may find the equation for the ionisation written in a simplified form:

This shows the hydrogen chloride dissolved in the water splitting to give
hydrogen ions in solution and chloride ions in solution.
This version is often used in this work just to make things look easier. If
you use it, remember that the water is actually involved, and that when
you write H+(aq) what you really mean is a hydroxonium ion, H3O+.

http://www.chemguide.co.uk/physical/acidbaseeqia/acids.html (2 of 11)30/12/2004 11:15:33

strong and weak acids

Note: UK A' level students should find out what their


examiners prefer on this. You are unlikely to find this from
your syllabus, but should look at recent exam papers and
mark schemes. If you haven't got copies of your syllabus and
past papers, you should have! Follow this link to find out how
to get hold of them.

Strong acids and pH


pH is a measure of the concentration of hydrogen ions in a solution.
Strong acids like hydrochloric acid at the sort of concentrations you
normally use in the lab have a pH around 0 to 1. The lower the pH, the
higher the concentration of hydrogen ions in the solution.
Defining pH

Note: If you are asked to define pH in an exam, simply write


down the expression in black. Never try to define it in words it is a waste of time, and you are too likely to miss something
out (like mentioning that the concentration has to be in mol
dm-3). In the expression, above, the square brackets imply
that, so you don't need to mention it.

http://www.chemguide.co.uk/physical/acidbaseeqia/acids.html (3 of 11)30/12/2004 11:15:33

strong and weak acids

Working out the pH of a strong acid


Suppose you had to work out the pH of 0.1 mol dm-3 hydrochloric acid.
All you have to do is work out the concentration of the hydrogen ions in
the solution, and then use your calculator to convert it to a pH.
With strong acids this is easy.
Hydrochloric acid is a strong acid - virtually 100% ionised. Each mole of
HCl reacts with the water to give 1 mole of hydrogen ions and 1 mole of
chloride ions
That means that if the concentration of the acid is 0.1 mol dm-3, then the
concentration of hydrogen ions is also 0.1 mol dm-3.
Use your calculator to convert this into pH. My calculator wants me to
enter 0.1, and then press the "log" button. Yours might want you to do it
in a different order. You need to find out!
log10 [0.1] = -1
But pH = - log10 [0.1]
- (-1) = 1
The pH of this acid is 1.

Note: If you want more examples to look at and to try


yourself (with fully worked solutions given), you may be
interested in my chemistry calculations book. This also
includes the slightly more confusing problem of converting pH
back into hydrogen ion concentration.

http://www.chemguide.co.uk/physical/acidbaseeqia/acids.html (4 of 11)30/12/2004 11:15:33

strong and weak acids

Weak acids
Explaining the term "weak acid"
A weak acid is one which doesn't ionise fully when it is dissolved in water.
Ethanoic acid is a typical weak acid. It reacts with water to produce
hydroxonium ions and ethanoate ions, but the back reaction is more
successful than the forward one. The ions react very easily to reform the
acid and the water.

At any one time, only about 1% of the ethanoic acid molecules have
converted into ions. The rest remain as simple ethanoic acid molecules.
Most organic acids are weak. Hydrogen fluoride (dissolving in water to
produce hydrofluoric acid) is a weak inorganic acid that you may come
across elsewhere.

Note: If you are interested in exploring organic acids further,


you will find them explained elsewhere on the site. It might be
a good idea to read the rest of this page first, though.
If you want to know why hydrogen fluoride is a weak acid,
you can find out by following this link. But beware! The
explanation is very complicated and definitely not for the fainthearted!
These pages are in completely different parts of this site. If
you follow either link, use the BACK button to return to this
current page.

http://www.chemguide.co.uk/physical/acidbaseeqia/acids.html (5 of 11)30/12/2004 11:15:33

strong and weak acids

Comparing the strengths of weak acids


The position of equilibrium of the reaction between the acid and water
varies from one weak acid to another. The further to the left it lies, the
weaker the acid is.

Note: If you don't understand about position of equilibrium


follow this link before you go any further.
You are also going to need to know about equilibrium
constants, Kc for homogeneous equilibria. There is no point
in reading any more of this page unless you do!
If you follow either link, use the BACK button to return to this
current page.

The acid dissociation constant, Ka


You can get a measure of the position of an equilibrium by writing an
equilibrium constant for the reaction. The lower the value for the
constant, the more the equilibrium lies to the left.
The dissociation (ionisation) of an acid is an example of a homogeneous
reaction. Everything is present in the same phase - in this case, in
solution in water. You can therefore write a simple expression for the
equilibrium constant, Kc.
Here is the equilibrium again:

You might expect the equilibrium constant to be written as:

http://www.chemguide.co.uk/physical/acidbaseeqia/acids.html (6 of 11)30/12/2004 11:15:33

strong and weak acids

However, if you think about this carefully, there is something odd about it.
At the bottom of the expression, you have a term for the concentration of
the water in the solution. That's not a problem - except that the number is
going to be very large compared with all the other numbers.
In 1 dm3 of solution, there are going to be about 55 moles of water.

Note: 1 mole of water weighs 18 g. 1 dm3 of solution


contains approximately 1000 g of water. Divide 1000 by 18 to
get approximately 55.

If you had a weak acid with a concentration of about 1 mol dm-3, and
only about 1% of it reacted with the water, the number of moles of water
is only going to fall by about 0.01. In other words, if the acid is weak the
concentration of the water is virtually constant.
In that case, there isn't a lot of point in including it in the expression as if
it were a variable. Instead, a new equilibrium constant is defined which
leaves it out. This new equilibrium constant is called Ka.

http://www.chemguide.co.uk/physical/acidbaseeqia/acids.html (7 of 11)30/12/2004 11:15:33

strong and weak acids

Note: The term for the concentration of water hasn't just


been ignored. What has happened is that the first expression
has been rearranged to give Kc (a constant) times the
concentration of water (another constant) on the left-hand
side. The product of those is then given the name Ka.
You don't need to worry about this unless you really insist! All
you need to do is to learn the format of the expression for Ka.

You may find the Ka expression written differently if you work from the
simplified version of the equilibrium reaction:

This may be written with or without state symbols.


It is actually exactly the same as the previous expression for Ka!
Remember that although we often write H+ for hydrogen ions in solution,
what we are actually talking about are hydroxonium ions.
This second version of the Ka expression isn't as precise as the first one,
but your examiners may well accept it. Find out!

To take a specific common example, the equilibrium for the dissociation


of ethanoic acid is properly written as:

The Ka expression is:

http://www.chemguide.co.uk/physical/acidbaseeqia/acids.html (8 of 11)30/12/2004 11:15:33

strong and weak acids

If you are using the simpler version of the equilibrium . . .

. . . the Ka expression is:

Note: Because you are likely to come across both of these


versions depending on where you read about Ka, you would
be wise to get used to using either. For exam purposes,
though, use whichever your examiners seem to prefer.

The table shows some values of Ka for some simple acids:


acid

Ka (mol dm-3)

hydrofluoric acid

5.6 x 10-4

methanoic acid

1.6 x 10-4

ethanoic acid

1.7 x 10-5

hydrogen sulphide

8.9 x 10-8

These are all weak acids because the values for Ka are very small. They
are listed in order of decreasing acid strength - the Ka values get smaller

http://www.chemguide.co.uk/physical/acidbaseeqia/acids.html (9 of 11)30/12/2004 11:15:33

strong and weak acids

as you go down the table.


However, if you aren't very happy with numbers, that isn't immediately
obvious. Because the numbers are in two parts, there is too much to
think about quickly!
To avoid this, the numbers are often converted into a new, easier form,
called pKa.

An introduction to pKa
pKa bears exactly the same relationship to Ka as pH does to the
hydrogen ion concentration:

If you use your calculator on all the Ka values in the table above and
convert them into pKa values, you get:
Ka (mol dm-3)

pKa

hydrofluoric acid

5.6 x 10-4

3.3

methanoic acid

1.6 x 10-4

3.8

ethanoic acid

1.7 x 10-5

4.8

hydrogen sulphide

8.9 x 10-8

7.1

acid

Note: Notice that unlike Ka, pKa doesn't have any units.

http://www.chemguide.co.uk/physical/acidbaseeqia/acids.html (10 of 11)30/12/2004 11:15:33

strong and weak acids

Notice that the weaker the acid, the larger the value of pKa. It is now
easy to see the trend towards weaker acids as you go down the table.
Remember this:

The lower the value for pKa, the stronger the acid.

The higher the value for pKa, the weaker the acid.

Note: If you need to know about Ka and pKa, you are quite
likely to need to be able to do calculations with them. You will
probably need to be able to calculate the pH of a weak acid
from its concentration and Ka or pKa. You may need to
reverse this and calculate a value for pKa from pH and
concentration. I can't help you with these calculations on this
site, but they are all covered in detail in my chemistry
calculations book.

Where would you like to go now?


To the acid-base equilibria menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/acidbaseeqia/acids.html (11 of 11)30/12/2004 11:15:33

an introduction to chemical equilibria

AN INTRODUCTION TO CHEMICAL EQUILIBRIA

This page looks at the basic ideas underpinning the idea of a chemical
equilibrium. It talks about reversible reactions and how they behave if the
system is closed. This leads to the idea of a dynamic equilibrium, and
what the common term "position of equilibrium" means.

Reversible reactions
A reversible reaction is one which can be made to go in either direction
depending on the conditions.
If you pass steam over hot iron the steam reacts with the iron to produce
a black, magnetic oxide of iron called triiron tetroxide, Fe3O4.

The hydrogen produced in the reaction is swept away by the stream of


steam.

Under different conditions, the products of this reaction will also react
together. Hydrogen passed over hot triiron tetroxide reduces it to iron.
Steam is also produced.

This time the steam produced in the reaction is swept away by the
stream of hydrogen.
http://www.chemguide.co.uk/physical/equilibria/introduction.html (1 of 10)30/12/2004 11:15:43

an introduction to chemical equilibria

These reactions are reversible, but under the conditions normally used,
they become one-way reactions. The products aren't left in contact with
each other, so the reverse reaction can't happen.
Reversible reactions happening in a closed system
A closed system is one in which no substances are either added to the
system or lost from it. Energy can, however, be transferred in or out at
will.
In the example we've been looking at, you would have to imagine iron
being heated in steam in a closed container. Heat is being added to the
system, but none of the substances in the reaction can escape. The
system is closed.
As the triiron tetroxide and hydrogen start to be formed, they will also
react again to give the original iron and steam. So, if you analysed the
mixture after a while, what would you find?
You would find that you had established what is known as a dynamic
equilibrium. To explain what that means, we are going to use a much
simpler example . . .

Dynamic equilibria
Getting a visual feel for a dynamic equilibrium
Imagine a substance which can exist in two forms - a blue form or an
orange form - and that each form can react to give the other one. We are
going to let them react in a closed system. Neither form can escape.
http://www.chemguide.co.uk/physical/equilibria/introduction.html (2 of 10)30/12/2004 11:15:43

an introduction to chemical equilibria

Assume that the blue form turns into the orange one much faster than
the other way round. In any given time, these are the chances of the two
changes happening:

You can simulate this very easily with some coloured paper cut up into
small pieces (a different colour on each side), and a dice.
The following are the real results of a "reaction" I did myself. I started
with 16 blue squares and looked at each one in turn and decided
whether it should change colour by throwing a dice.
A blue square was turned into an orange square (the bit of paper was
turned over!) if I threw a 4, 5 or 6
An orange square was turned into a blue square only if I threw a 6 while I
was looking at that particular square.
Once I had looked at all 16 squares, I started the process all over again but obviously with a different starting pattern. The diagrams show the
results of doing this 11 times (plus the original 16 blue squares).

http://www.chemguide.co.uk/physical/equilibria/introduction.html (3 of 10)30/12/2004 11:15:43

an introduction to chemical equilibria

You can see that the "reaction" is continuing all the time. The exact
pattern of orange and blue is constantly changing. However, the overall
numbers of orange and of blue squares remain remarkably constant most commonly, 12 orange ones to 4 blue ones.

Note: To be honest, this was a lucky fluke, given the small


number of squares I was working with. If you repeated this
with much larger number of squares (say, several thousand),
you would find that your patterns settled down quite reliably
close to 75% orange and 25% blue. On the other hand, that
would be seriously tedious!
If you had the huge numbers of particles taking part in
chemical reactions, the proportions would be spot on 75% to
25%.

http://www.chemguide.co.uk/physical/equilibria/introduction.html (4 of 10)30/12/2004 11:15:43

an introduction to chemical equilibria

Explaining the term "dynamic equilibrium"


The reaction has reached equilibrium in the sense that there is no further
change in the numbers of blue and orange squares. However, the
reaction is still continuing. For every orange square that turns blue,
somewhere in the mixture it is replaced by a blue square turning orange.
This is known as a dynamic equilibrium. The word dynamic shows that
the reaction is still continuing.
You can show dynamic equilibrium in an equation for a reaction by the
use of special arrows. In the present case, you would write it as:

It is important to realise that this doesn't just mean that the reaction is
reversible. It means that you have a reversible reaction in a state of
dynamic equilibrium.
The "forward reaction" and the "back reaction"
The change from left to right in the equation (in this case from blue to
orange as it is written) is known as the forward reaction. The change
from right to left is the back reaction.
Position of equilibrium
In the example we've used, the equilibrium mixture contained more
orange squares than blue ones. Position of equilibrium is a way of
expressing this. You can say things like:

"The position of equilibrium lies towards the orange."

"The position of equilibrium lies towards the right-hand side."

If the conditions of the experiment change (by altering the relative


chances of the forward and back reactions happening), the composition
of the equilibrium mixture will also change.
For example, if changing the conditions produced more blue in the
http://www.chemguide.co.uk/physical/equilibria/introduction.html (5 of 10)30/12/2004 11:15:43

an introduction to chemical equilibria

equilibrium mixture, you would say "The position of equilibrium has


moved to the left" or "The position of equilibrium has moved towards the
blue".

Note: If you can be bothered, try the effect on the position of


equilibrium of increasing the chances of an orange square
turning blue from 1 in 6 to 2 in 6. In other words, allow it to
change if you throw either a 5 or a 6 with your dice.

Reaching equilibrium from the other side


What happens if you started the reaction with orange squares rather than
blue ones, but kept the chances of each change happening the same as
in the first example? This is the result of my "reaction".

http://www.chemguide.co.uk/physical/equilibria/introduction.html (6 of 10)30/12/2004 11:15:43

an introduction to chemical equilibria

Once again, you can see that exactly the same position of equilibrium is
being established as when we started with the blue squares. You get
exactly the same equilibrium mixture irrespective of which side of the
equation you start from - provided the conditions are the same in both
cases.

Remember: You can't get the numbers to work out exactly


using this small number of particles. Chance fluctuations are
too noticeable. Once again, if you used really large numbers
of particles, the equilibrium mixture would contain 75%
orange and 25% blue. Given the number of particles we're
working with, the "reaction" is remarkably close to that on
average.

A more formal look at dynamic equilibria


Thinking about reaction rates
This is the equation for a general reaction which has reached dynamic
equilibrium:

How did it get to that state? Let's assume that we started with A and B.
At the beginning of the reaction, the concentrations of A and B were at
their maximum. That means that the rate of the reaction was at its fastest.
As A and B react, their concentrations fall. That means that they are less
likely to collide and react, and so the rate of the forward reaction falls as
time goes on.

http://www.chemguide.co.uk/physical/equilibria/introduction.html (7 of 10)30/12/2004 11:15:43

an introduction to chemical equilibria

In the beginning, there isn't any C and D, so there can't be any reaction
between them. As time goes on, though, their concentrations in the
mixture increase and they are more likely to collide and react.
With time, the rate of the reaction between C and D increases:

Eventually, the rates of the two reactions will become equal. A and B will
be converting into C and D at exactly the same rate as C and D convert
back into A and B again.

http://www.chemguide.co.uk/physical/equilibria/introduction.html (8 of 10)30/12/2004 11:15:43

an introduction to chemical equilibria

At this point there won't be any further change in the amounts of A, B, C


and D in the mixture. As fast as something is being removed, it is being
replaced again by the reverse reaction. We have reached a position of
dynamic equilibrium.

A summary
A dynamic equilibrium occurs when you have a reversible reaction in a
closed system. Nothing can be added to the system or taken away from
it apart from energy.
At equilibrium, the quantities of everything present in the mixture remain
constant, although the reactions are still continuing. This is because the
rates of the forward and the back reactions are equal.
If you change the conditions in a way which changes the relative rates of
the forward and back reactions you will change the position of equilibrium
- in other words, change the proportions of the various substances
present in the equilibrium mixture. This is explored in detail on other
pages in this equilibrium section.

Where would you like to go now?


To the equilibrium menu . . .

http://www.chemguide.co.uk/physical/equilibria/introduction.html (9 of 10)30/12/2004 11:15:43

an introduction to chemical equilibria

To the Physical Chemistry menu . . .


To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/equilibria/introduction.html (10 of 10)30/12/2004 11:15:43

Chemical equilibria menu

Understanding Chemistry

CHEMICAL EQUILIBRIA MENU

Basic descriptive topics


Explaining chemical equilibria . . .
Explains some basic concepts: reversible reactions, closed
systems, dynamic equilibrium, and position of equilibrium.
Le Chatelier's Principle . . .
Explains the use of Le Chatelier's Principle in working out what
happens to an equilibrium if you change the conditions
(concentration, temperature, pressure or catalyst).
The Haber Process . . .
Looks in detail at how the conditions for the Haber Process for the
manufacture of ammonia are determined by the equilibrium
between hydrogen, nitrogen and ammonia.
The Contact Process . . .
Looks in detail at how the conditions for the Contact Process for
the manufacture of sulphuric acid are determined by the
equilibrium between sulphur dioxide, oxygen and sulphur trioxide.
The manufacture of ethanol from ethene . . .
Looks in detail at how the conditions for this process are
determined by the equilibrium reaction at its heart.

http://www.chemguide.co.uk/physical/equilibmenu.html (1 of 2)30/12/2004 11:15:45

Chemical equilibria menu

Mathematical bits
Kc . . .
Explains equilibrium constants expressed in terms of
concentrations.
Kp . . .
Explains equilibrium constants expressed in terms of partial
pressures. Includes an explanation of the terms partial pressure
and mole fraction. Before you read this page, you should have a
reasonable understanding of Kc.
The effect of changing conditions . . .
Looks at the way that equilibrium constants are changed (or not!)
as you change the conditions for a reaction (concentration,
temperature, pressure or catalyst), and how this ties in with Le
Chatelier's Principle. This page assumes that you are fully
confident about writing expressions for both kinds of equilibrium
constants.

Go to physical chemistry menu . . .


Go to Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/equilibmenu.html (2 of 2)30/12/2004 11:15:45

Le Chatelier's Principle

LE CHATELIER'S PRINCIPLE

This page looks at Le Chatelier's Principle and explains how to apply it to


reactions in a state of dynamic equilibrium. It covers changes to the
position of equilibrium if you change concentration, pressure or
temperature. It also explains very briefly why catalysts have no effect on
the position of equilibrium.

Important: If you aren't sure about the words dynamic


equilibrium or position of equilibrium you should read the
introductory page before you go on

It is important in understanding everything on this page to realise that Le


Chatelier's Principle is no more than a useful guide to help you work out
what happens when you change the conditions in a reaction in dynamic
equilibrium. It doesn't explain anything. I'll keep coming back to that point!

Using Le Chatelier's Principle


A statement of Le Chatelier's Principle

If a dynamic equilibrium is disturbed by changing the conditions,


the position of equilibrium moves to counteract the change.

Using Le Chatelier's Principle with a change of concentration


Suppose you have an equilibrium established between four substances
A, B, C and D.

http://www.chemguide.co.uk/physical/equilibria/lechatelier.html (1 of 9)30/12/2004 11:15:54

Le Chatelier's Principle

Note: In case you wonder, the reason for choosing this


equation rather than having just A + B on the left-hand side is
because further down this page I need an equation which has
different numbers of molecules on each side. I am going to
use that same equation throughout this page.

What would happen if you changed the conditions by increasing the


concentration of A?
According to Le Chatelier, the position of equilibrium will move in such a
way as to counteract the change. That means that the position of
equilibrium will move so that the concentration of A decreases again - by
reacting it with B and turning it into C + D. The position of equilibrium
moves to the right.

This is a useful way of converting the maximum possible amount of B


into C and D. You might use it if, for example, B was a relatively
expensive material whereas A was cheap and plentiful.
What would happen if you changed the conditions by decreasing
the concentration of A?
According to Le Chatelier, the position of equilibrium will move so that
the concentration of A increases again. That means that more C and D
will react to replace the A that has been removed. The position of
equilibrium moves to the left.

This is esssentially what happens if you remove one of the products of


http://www.chemguide.co.uk/physical/equilibria/lechatelier.html (2 of 9)30/12/2004 11:15:54

Le Chatelier's Principle

the reaction as soon as it is formed. If, for example, you removed C as


soon as it was formed, the position of equilibrium would move to the right
to replace it. If you kept on removing it, the equilibrium position would
keep on moving rightwards - turning this into a one-way reaction.

Important
This isn't in any way an explanation of why the position of equilibrium
moves in the ways described. All Le Chatelier's Principle gives you is a
quick way of working out what happens.

Note: If you know about equilibrium constants, you will find a


more detailed explanation of the effect of a change of
concentration by following this link. If you don't know anything
about equilibrium constants, you should ignore this link.
If you choose to follow it, return to this page via the BACK
button on your browser or via the equilibrium menu.

Using Le Chatelier's Principle with a change of pressure


This only applies to reactions involving gases:

What would happen if you changed the conditions by increasing the


pressure?
According to Le Chatelier, the position of equilibrium will move in such a
way as to counteract the change. That means that the position of
equilibrium will move so that the pressure is reduced again.
Pressure is caused by gas molecules hitting the sides of their container.
The more molecules you have in the container, the higher the pressure
will be. The system can reduce the pressure by reacting in such a way as
to produce fewer molecules.
http://www.chemguide.co.uk/physical/equilibria/lechatelier.html (3 of 9)30/12/2004 11:15:54

Le Chatelier's Principle

In this case, there are 3 molecules on the left-hand side of the equation,
but only 2 on the right. By forming more C and D, the system causes the
pressure to reduce.
Increasing the pressure on a gas reaction shifts the position of
equilibrium towards the side with fewer molecules.

What would happen if you changed the conditions by decreasing


the pressure?
The equilibrium will move in such a way that the pressure increases
again. It can do that by producing more molecules. In this case, the
position of equilibrium will move towards the left-hand side of the reaction.

What happens if there are the same number of molecules on both


sides of the equilibrium reaction?
In this case, increasing the pressure has no effect whatsoever on the
position of the equilibrium. Because you have the same numbers of
molecules on both sides, the equilibrium can't move in any way that will
reduce the pressure again.

Important
Again, this isn't an explanation of why the position of equilibrium moves
in the ways described. You will find a rather mathematical treatment of

http://www.chemguide.co.uk/physical/equilibria/lechatelier.html (4 of 9)30/12/2004 11:15:54

Le Chatelier's Principle

the explanation by following the link below.

Note: You will find a detailed explanation by following this


link. If you don't know anything about equilibrium constants
(particularly Kp), you should ignore this link. The same thing
applies if you don't like things to be too mathematical! If you
are a UK A' level student, you won't need this explanation.
If you choose to follow the link, return to this page via the
BACK button on your browser or via the equilibrium menu.

Using Le Chatelier's Principle with a change of temperature


For this, you need to know whether heat is given out or absorbed during
the reaction. Assume that our forward reaction is exothermic (heat is
evolved):

This shows that 250 kJ is evolved (hence the negative sign) when 1 mole
of A reacts completely with 2 moles of B. For reversible reactions, the
value is always given as if the reaction was one-way in the forward
direction.
The back reaction (the conversion of C and D into A and B) would be
endothermic by exactly the same amount.

http://www.chemguide.co.uk/physical/equilibria/lechatelier.html (5 of 9)30/12/2004 11:15:54

Le Chatelier's Principle

What would happen if you changed the conditions by increasing the


temperature?
According to Le Chatelier, the position of equilibrium will move in such a
way as to counteract the change. That means that the position of
equilibrium will move so that the temperature is reduced again.
Suppose the system is in equilibrium at 300C, and you increase the
temperature to 500C. How can the reaction counteract the change you
have made? How can it cool itself down again?
To cool down, it needs to absorb the extra heat that you have just put in.
In the case we are looking at, the back reaction absorbs heat. The
position of equilibrium therefore moves to the left. The new equilibrium
mixture contains more A and B, and less C and D.

If you were aiming to make as much C and D as possible, increasing the


temperature on a reversible reaction where the forward reaction is
exothermic isn't a good idea!
What would happen if you changed the conditions by decreasing
the temperature?
The equilibrium will move in such a way that the temperature increases
again.
Suppose the system is in equilibrium at 500C and you reduce the
temperature to 400C. The reaction will tend to heat itself up again to
return to the original temperature. It can do that by favouring the
exothermic reaction.
The position of equilibrium will move to the right. More A and B are
converted into C and D at the lower temperature.

http://www.chemguide.co.uk/physical/equilibria/lechatelier.html (6 of 9)30/12/2004 11:15:54

Le Chatelier's Principle

Summary

Increasing the temperature of a system in dynamic equilibrium


favours the endothermic reaction. The system counteracts the
change you have made by absorbing the extra heat.
Decreasing the temperature of a system in dynamic equilibrium
favours the exothermic reaction. The system counteracts the
change you have made by producing more heat.

Important
Again, this isn't in any way an explanation of why the position of
equilibrium moves in the ways described. It is only a way of helping you
to work out what happens.

Note: I am not going to attempt an explanation of this


anywhere on the site. To do it properly is far too difficult for
this level. It is possible to come up with an explanation of
sorts by looking at how the rate constants for the forward and
back reactions change relative to each other by using the
Arrhenius equation, but this isn't a standard way of doing it,
and is liable to confuse those of you going on to do a
Chemistry degree. If you aren't going to do a Chemistry
degree, you won't need to know about this anyway!

http://www.chemguide.co.uk/physical/equilibria/lechatelier.html (7 of 9)30/12/2004 11:15:54

Le Chatelier's Principle

Le Chatelier's Principle and catalysts


Catalysts have sneaked onto this page under false pretences, because
adding a catalyst makes absolutely no difference to the position of
equilibrium, and Le Chatelier's Principle doesn't apply to them.
This is because a catalyst speeds up the forward and back reaction to
the same extent. Because adding a catalyst doesn't affect the relative
rates of the two reactions, it can't affect the position of equilibrium. So
why use a catalyst?
For a dynamic equilibrium to be set up, the rates of the forward reaction
and the back reaction have to become equal. This doesn't happen
instantly. For a very slow reaction, it could take years! A catalyst speeds
up the rate at which a reaction reaches dynamic equilibrium.

Note: You might try imagining how long it would take to


establish a dynamic equilibrium if you took the visual model
on the introductory page and reduced the chances of the
colours changing by a factor of 1000 - from 3 in 6 to 3 in 6000
and from 1 in 6 to 1 in 6000.
Starting with blue squares, by the end of the time taken for
the examples on that page, you would most probably still
have entirely blue squares. Eventually, though, you would
end up with the same sort of patterns as before - containing
25% blue and 75% orange squares.

http://www.chemguide.co.uk/physical/equilibria/lechatelier.html (8 of 9)30/12/2004 11:15:54

Le Chatelier's Principle

Where would you like to go now?


To the equilibrium menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/equilibria/lechatelier.html (9 of 9)30/12/2004 11:15:54

equilibrium constants and changing conditions

EQUILIBRIUM CONSTANTS and LE


CHATELIER'S PRINCIPLE

This page looks at the relationship between equilibrium constants and Le


Chatelier's Principle. Students often get confused about how it is possible
for the position of equilibrium to change as you change the conditions of
a reaction, although the equilibrium constant may remain the same.
Be warned that this page assumes a good understanding of Le
Chatelier's Principle and how to write expressions for equilibrium
constants.

Important: If you aren't happy about the basics of


equilibrium, explore the equilibrium menu before you waste
your time on this page.
This page should only be read when you are confident about
everything else to do with equilibria.

Changing concentrations
The facts
Equilibrium constants aren't changed if you change the concentrations
of things present in the equilibrium. The only thing that changes an
equilibrium constant is a change of temperature.
The position of equilibrium is changed if you change the concentration
of something present in the mixture. According to Le Chatelier's
Principle, the position of equilibrium moves in such a way as to tend to
undo the change that you have made.
Suppose you have an equilibrium established between four substances
A, B, C and D.
http://www.chemguide.co.uk/physical/equilibria/change.html (1 of 10)30/12/2004 11:16:03

equilibrium constants and changing conditions

According to Le Chatelier's Principle, if you decrease the concentration of


C, for example, the position of equilibrium will move to the right to
increase the concentration again.

Note: The reason for choosing an equation with "2B" will


become clearer when I deal with the effect of pressure further
down the page.

Explanation in terms of the constancy of the equilibrium constant


The equilibrium constant, Kc for this reaction looks like this:

If you have moved the position of the equilibrium to the right (and so
increased the amount of C and D), why hasn't the equilibrium constant
increased?
This is actually the wrong question to ask! We need to look at it the other
way round.
Let's assume that the equilibrium constant mustn't change if you
decrease the concentration of C - because equilibrium constants are
constant at constant temperature. Why does the position of equilibrium
move as it does?
If you decrease the concentration of C, the top of the Kc expression gets
smaller. That would change the value of Kc. In order for that not to
happen, the concentrations of C and D will have to increase again, and
those of A and B must decrease. That happens until a new balance is
reached when the value of the equilibrium constant expression reverts to
what it was before.

http://www.chemguide.co.uk/physical/equilibria/change.html (2 of 10)30/12/2004 11:16:03

equilibrium constants and changing conditions

The position of equilibrium moves - not because Le Chatelier says it


must - but because of the need to keep a constant value for the
equilibrium constant.
If you decrease the concentration of C:

Changing pressure
This only applies to systems involving at least one gas.
The facts
Equilibrium constants aren't changed if you change the pressure of the
system. The only thing that changes an equilibrium constant is a change
of temperature.
The position of equilibrium may be changed if you change the
pressure. According to Le Chatelier's Principle, the position of equilibrium
moves in such a way as to tend to undo the change that you have made.
That means that if you increase the pressure, the position of equilibrium
will move in such a way as to decrease the pressure again - if that is
possible. It can do this by favouring the reaction which produces the
fewer molecules. If there are the same number of molecules on each
http://www.chemguide.co.uk/physical/equilibria/change.html (3 of 10)30/12/2004 11:16:03

equilibrium constants and changing conditions

side of the equation, then a change of pressure makes no difference to


the position of equilibrium.
Explanation
Where there are different numbers of molecules on each side of the
equation
Let's look at the same equilibrium we've used before. This one would be
affected by pressure because there are 3 molecules on the left but only 2
on the right. An increase in pressure would move the position of
equilibrium to the right.

Because this is an all-gas equilibriium, it is much easier to use Kp:

Once again, it is easy to suppose that, because the position of


equilibrium will move to the right if you increase the pressure, Kp will
increase as well. Not so!
To understand why, you need to modify the Kp expression.
Remember the relationship between partial pressure, mole fraction and
total pressure?

http://www.chemguide.co.uk/physical/equilibria/change.html (4 of 10)30/12/2004 11:16:03

equilibrium constants and changing conditions

Note: If you aren't happy with this, read the beginning of the
page about Kp before you go on.
Use the BACK button on your browser to return to this page.

Replacing all the partial pressure terms by mole fractions and total
pressure gives you this:

If you sort this out, most of the "P"s cancel out - but one is left at the
bottom of the expression.

Now, remember that Kp has got to stay constant because the


temperature is unchanged. How can that happen if you increase P?
To compensate, you would have to increase the terms on the top, xC and
xD, and decrease the terms on the bottom, xA and xB.
Increasing the terms on the top means that you have increased the mole
fractions of the molecules on the right-hand side. Decreasing the terms
on the bottom means that you have decreased the mole fractions of the
molecules on the left.
That is another way of saying that the position of equilibrium has moved
to the right - exactly what Le Chatelier's Principle predicts. The position
of equilibrium moves so that the value of Kp is kept constant.
Where there are the same numbers of molecules on each side of the
equation
http://www.chemguide.co.uk/physical/equilibria/change.html (5 of 10)30/12/2004 11:16:03

equilibrium constants and changing conditions

In this case, the position of equilibrium isn't affected by a change of


pressure. Why not?

Let's go through the same process as before:

Substituting mole fractions and total pressure:

. . . and cancelling out as far as possible:

There isn't a single "P" left in the expression. Changing the pressure
can't make any difference to the Kp expression. The position of
equilibrium doesn't need to move to keep Kp constant.

Changing temperature
The facts
Equilibrium constants are changed if you change the temperature of
the system. Kc or Kp are constant at constant temperature, but they vary
as the temperature changes.
Look at the equilibrium involving hydrogen, iodine and hydrogen iodide:

http://www.chemguide.co.uk/physical/equilibria/change.html (6 of 10)30/12/2004 11:16:03

equilibrium constants and changing conditions

The Kp expression is:

Two values for Kp are:

temperature

Kp

500 K

160

700 K

54

You can see that as the temperature increases, the value of Kp falls.

Note: You might possibly be wondering what the units of Kp


are. This particular example was chosen because in this
case, Kp doesn't have any units. It is just a number.
The units for equilibrium constants vary from case to case. It
is much easier to understand this from a book than from a lot
of maths on screen. You will find this explained in my
chemistry calculations book.

http://www.chemguide.co.uk/physical/equilibria/change.html (7 of 10)30/12/2004 11:16:03

equilibrium constants and changing conditions

This is typical of what happens with any equilibrium where the forward
reaction is exothermic. Increasing the temperature decreases the value
of the equilibrium constant.
Where the forward reaction is endothermic, increasing the temperature
increases the value of the equilibrium constant.

Note: Any explanation for this needs knowledge beyond the


scope of any UK A' level syllabus.

The position of equilibrium also changes if you change the


temperature. According to Le Chatelier's Principle, the position of
equilibrium moves in such a way as to tend to undo the change that you
have made.
If you increase the temperature, the position of equilibrium will move in
such a way as to reduce the temperature again. It will do that by
favouring the reaction which absorbs heat.
In the equilibrium we've just looked at, that will be the back reaction
because the forward reaction is exothermic.

So, according to Le Chatelier's Principle the position of equilibrium will


move to the left. Less hydrogen iodide will be formed, and the equilibrium
mixture will contain more unreacted hydrogen and iodine.
That is entirely consistent with a fall in the value of the equilibrium
constant.

http://www.chemguide.co.uk/physical/equilibria/change.html (8 of 10)30/12/2004 11:16:03

equilibrium constants and changing conditions

Adding a catalyst
The facts
Equilibrium constants aren't changed if you add (or change) a catalyst.
The only thing that changes an equilibrium constant is a change of
temperature.
The position of equilibrium is not changed if you add (or change) a
catalyst.
Explanation
A catalyst speeds up both the forward and back reactions by exactly the
same amount. Dynamic equilibrium is established when the rates of the
forward and back reactions become equal. If a catalyst speeds up both
reactions to the same extent, then they will remain equal without any
need for a shift in position of equilibrium.

http://www.chemguide.co.uk/physical/equilibria/change.html (9 of 10)30/12/2004 11:16:03

equilibrium constants and changing conditions

Note: If you know about the Arrhenius equation, it isn't too


difficult to use it to show that the ratio of the rate constants for
the forward and back reactions isn't affected by adding a
catalyst. Although the activation energies for the two
reactions change when you add a catalyst, they both change
by the same amount.
I'm not going to do this bit of algebra, because it would never
be asked in a UK A' level exam.

Where would you like to go now?


To the equilibrium menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/equilibria/change.html (10 of 10)30/12/2004 11:16:03

equilibrium constants - Kp

EQUILIBRIUM CONSTANTS: Kp

This page explains equilibrium constants expressed in terms of partial


pressures of gases, Kp. It covers an explanation of the terms mole
fraction and partial pressure, and looks at Kp for both homogeneous and
heterogeneous reactions involving gases.
The page assumes that you are already familiar with the concept of an
equilibrium constant, and that you know about Kc - an equilibrium
constant expressed in terms of concentrations

Important: If you have come directly to this page via a


search engine (including the Google site search on the Main
Menu page), you should first read the page on equilibrium
constants - Kc before you go on - unless you are already fully
confident about how to write expressions for Kc.
You will find a link back to this page at the bottom of the Kc
page.

Defining some terms


Before we can go any further, there are two terms relating to mixtures of
gases that you need to be familiar with.
Mole fraction
If you have a mixture of gases (A, B, C, etc), then the mole fraction of
gas A is worked out by dividing the number of moles of A by the total
number of moles of gas.
The mole fraction of gas A is often given the symbol xA. The mole
fraction of gas B would be xB - and so on.
http://www.chemguide.co.uk/physical/equilibria/kp.html (1 of 7)30/12/2004 11:16:15

equilibrium constants - Kp

Pretty obvious really!


For example, in a mixture of 1 mole of nitrogen and 3 moles of hydrogen,
there are a total of 4 moles of gas. The mole fraction of nitrogen is 1/4
(0.25) and of hydrogen is 3/4 (0.75).

Partial pressure
The partial pressure of one of the gases in a mixture is the pressure
which it would exert if it alone occupied the whole container.
The partial pressure of gas A is often given the symbol PA. The partial
pressure of gas B would be PB - and so on.
There are two important relationships involving partial pressures. The
first is again fairly obvious.
The total pressure of a mixture of gases is equal to the sum of the partial
pressures.

It is easy to see this visually:

Gas A is creating a pressure (its partial pressure) when its molecules hit
http://www.chemguide.co.uk/physical/equilibria/kp.html (2 of 7)30/12/2004 11:16:15

equilibrium constants - Kp

the walls of its container. Gas B does the same. When you mix them up,
they just go on doing what they were doing before. The total pressure is
due to both molecules hitting the walls - in other words, the sum of the
partial pressures.
The more important relationship is the second one:

Learn it!
That means that if you had a mixture made up of 20 moles of nitrogen,
60 moles of hydrogen and 20 moles of ammonia (a total of 100 moles of
gases) at 200 atmospheres pressure, the partial pressures would be
calculated like this:

gas

mole fraction

partial pressure

nitrogen

20/100 = 0.2

0.2 x 200 = 40 atm

hydrogen

60/100 = 0.6

0.6 x 200 = 120 atm

ammonia

20/100 = 0.2

0.2 x 200 = 40 atm

Partial pressures can be quoted in any normal pressure units. The


common ones are atmospheres or N m-2 (newtons per square metre).

Kp in homogeneous gaseous equilibria


A homogeneous equilibrium is one in which everything in the equilibrium
mixture is present in the same phase. In this case, to use Kp, everything
must be a gas.
A good example of a gaseous homogeneous equilibrium is the
http://www.chemguide.co.uk/physical/equilibria/kp.html (3 of 7)30/12/2004 11:16:15

equilibrium constants - Kp

conversion of sulphur dioxide to sulphur trioxide at the heart of the


Contact Process:

Writing an expression for Kp


We are going to start by looking at a general case with the equation:

If you allow this reaction to reach equilibrium and then measure (or work
out) the equilibrium partial pressures of everything, you can combine
these into the equilibrium constant, Kp.
Just like Kc, Kp always has the same value (provided you don't change
the temperature), irrespective of the amounts of A, B, C and D you
started with.

Kp has exactly the same format as Kc, except that partial pressures are
used instead of concentrations. The gases on the right-hand side of the
chemical equation are at the top of the expression, and those on the left
at the bottom.

Beware! People are sometimes tempted to write brackets


around the individual partial pressure terms. Don't do it! Even
if you intend to write normal round brackets, it is too easy in
an exam to write them as square brackets instead. This
makes it look as if you are confusing Kp with Kc. Examiners
don't like it, and you could be penalised.

http://www.chemguide.co.uk/physical/equilibria/kp.html (4 of 7)30/12/2004 11:16:15

equilibrium constants - Kp

The Contact Process equilibrium


You will remember that the equation for this is:

Kp is given by:

The Haber Process equilibrium


The equation for this is:

. . . and the Kp expression is:

Kc in heterogeneous equilibria
A typical example of a heterogeneous equilibrium will involve gases in
contact with solids.
Writing an expression for Kp for a heterogeneous equilibrium
Exactly as happens with Kc, you don't include any term for a solid in the
equilibrium expression.
http://www.chemguide.co.uk/physical/equilibria/kp.html (5 of 7)30/12/2004 11:16:15

equilibrium constants - Kp

The next two examples have already appeared on the Kc page.


The equilibrium produced on heating carbon with steam

Everything is exactly the same as before in the expression for Kp, except
that you leave out the solid carbon.

The equilibrium produced on heating calcium carbonate


This equilibrium is only established if the calcium carbonate is heated in
a closed system, preventing the carbon dioxide from escaping.

The only thing in this equilibrium which isn't a solid is the carbon dioxide.
That is all that is left in the equilibrium constant expression.

Calculations involving Kp
On the Kc page, I've already discussed the fact that the internet isn't a
good medium for learning how to do calculations.
If you want lots of worked examples and problems to do yourself centred
around Kp, you might be interested in my book on chemistry calculations.

http://www.chemguide.co.uk/physical/equilibria/kp.html (6 of 7)30/12/2004 11:16:15

equilibrium constants - Kp

Note: If you are interested in my chemistry calculations book


you might like to follow this link.

Where would you like to go now?


To look at Kp . . .
To the equilibrium menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/equilibria/kp.html (7 of 7)30/12/2004 11:16:15

equilibrium constants - Kc

EQUILIBRIUM CONSTANTS: Kc

This page explains what is meant by an equilibrium constant, introducing


equilibrium constants expressed in terms of concentrations, Kc. It
assumes that you are familiar with the concept of a dynamic equilibrium,
and know what is meant by the terms "homogeneous" and
"heterogeneous" as applied to chemical reactions.

Important: If you aren't sure about dynamic equilibria it is


important that you follow this link before you go on.
If you aren't sure what homogeneous and heterogeneous
mean, you would find it useful to follow this link and read the
beginning of the page that you will find (actually on catalysis).
Use the BACK button on your browser to return to this page.

We need to look at two different types of equilibria (homogeneous and


heterogeneous) separately, because the equilibrium constants are
defined differently.

A homogeneous equilibrium has everything present in the same


phase. The usual examples include reactions where everything is
a gas, or everything is present in the same solution.
A heterogeneous equilibrium has things present in more than
one phase. The usual examples include reactions involving solids
and gases, or solids and liquids.

Kc in homogeneous equilibria
This is the more straightforward case. It applies where everything in the
http://www.chemguide.co.uk/physical/equilibria/kc.html (1 of 8)30/12/2004 11:16:24

equilibrium constants - Kc

equilibrium mixture is present as a gas, or everything is present in the


same solution.
A good example of a gaseous homogeneous equilibrium is the
conversion of sulphur dioxide to sulphur trioxide at the heart of the
Contact Process:

A commonly used liquid example is the esterification reaction between


an organic acid and an alcohol - for example:

Writing an expression for Kc


We are going to look at a general case with the equation:

No state symbols have been given, but they will be all (g), or all (l), or all
(aq) if the reaction was between substances in solution in water.
If you allow this reaction to reach equilibrium and then measure the
equilibrium concentrations of everything, you can combine these
concentrations into an expression known as an equilibrium constant.
The equilibrium constant always has the same value (provided you don't
change the temperature), irrespective of the amounts of A, B, C and D
you started with. It is also unaffected by a change in pressure or whether
or not you are using a catalyst.

http://www.chemguide.co.uk/physical/equilibria/kc.html (2 of 8)30/12/2004 11:16:24

equilibrium constants - Kc

Compare this with the chemical equation for the equilibrium. The
convention is that the substances on the right-hand side of the equation
are written at the top of the Kc expression, and those on the left-hand
side at the bottom.
The indices (the powers that you have to raise the concentrations to - for
example, squared or cubed or whatever) are just the numbers that
appear in the equation.

Note: If you have come across orders of reaction, don't


confuse this with the powers that appear in the rate equation
for a reaction. Those powers (the order of the reaction with
respect to each of the reactants) are experimentally
determined. They don't have any direct connection with the
numbers that appear in the equation
You may come across attempts to derive the expression for
Kc by writing rate equations for the forward and back
reactions. Except in a very limited number of very simple
examples, this can't be done! These attempts make the
fundamental mistake of obtaining the rate equation from the
chemical equation. That's WRONG! Deriving an expression
for Kc is impossible at this level of chemistry.
It isn't relevant to this page, but if you want to find out more
about orders of reaction, you might like to follow this link at
some time in the future.

http://www.chemguide.co.uk/physical/equilibria/kc.html (3 of 8)30/12/2004 11:16:24

equilibrium constants - Kc

Some specific examples


The esterification reaction equilibrium
A typical equation might be:

There is only one molecule of everything shown in the equation. That


means that all the powers in the equilibrium constant expression are "1".
You don't need to write those into the Kc expression.

As long as you keep the temperature the same, whatever proportions of


acid and alcohol you mix together, once equilibrium is reached, Kc
always has the same value. At room temperature, this value is
approximately 4 for this reaction.
The equilibrium in the hydrolysis of esters
This is the reverse of the last reaction:

The Kc expression is:

http://www.chemguide.co.uk/physical/equilibria/kc.html (4 of 8)30/12/2004 11:16:24

equilibrium constants - Kc

If you compare this with the previous example, you will see that all that
has happened is that the expression has turned upside-down. Its value at
room temperature will be approximately 1/4 (0.25).
It is really important to write down the equilibrium reaction whenever you
talk about an equilibrium constant. That is the only way that you can be
sure that you have got the expression the right way up - with the righthand substances on the top and the left-hand ones at the bottom.
The Contact Process equilibrium
You will remember that the equation for this is:

This time the Kc expression will include some visible powers:

Although everything is present as a gas, you still measure concentrations


in mol dm-3. There is another equilibrium constant called Kp which is
more frequently used for gases. You will find a link to that at the bottom
of the page.
The Haber Process equilibrium
The equation for this is:

. . . and the Kc expression is:

http://www.chemguide.co.uk/physical/equilibria/kc.html (5 of 8)30/12/2004 11:16:24

equilibrium constants - Kc

Kc in heterogeneous equilibria
Typical examples of a heterogeneous equilibrium include:
The equilibrium established if steam is in contact with red hot carbon.
Here we have gases in contact with a solid.

If you shake copper with silver nitrate solution, you get this equilibrium
involving solids and aqueous ions:

Writing an expression for Kc for a heterogeneous equilibrium


The important difference this time is that you don't include any term for a
solid in the equilibrium expression.
Taking another look at the two examples above, and adding a third one:
The equilibrium produced on heating carbon with steam

Everything is exactly the same as before in the equilibrium constant


expression, except that you leave out the solid carbon.

http://www.chemguide.co.uk/physical/equilibria/kc.html (6 of 8)30/12/2004 11:16:24

equilibrium constants - Kc

The equilibrium produced between copper and silver ions

Both the copper on the left-hand side and the silver on the right are
solids. Both are left out of the equilibrium constant expression.

The equilibrium produced on heating calcium carbonate


This equilibrium is only established if the calcium carbonate is heated in
a closed system, preventing the carbon dioxide from escaping.

The only thing in this equilibrium which isn't a solid is the carbon dioxide.
That is all that is left in the equilibrium constant expression.

Calculations involving Kc
There are all sorts of calculations you might be expected to do which are
centred around equilibrium constants. You might be expected to
calculate a value for Kc including its units (which vary from case to case).
Alternatively you might have to calculate equilibrium concentrations from
a given value of Kc and given starting concentrations.
This is simply too huge a topic to be able to deal with satisfactorily on the
internet. It isn't the best medium for learning how to do chemistry
http://www.chemguide.co.uk/physical/equilibria/kc.html (7 of 8)30/12/2004 11:16:24

equilibrium constants - Kc

calculations. It is much easier to do this from a carefully structured book


giving you lots of worked examples and lots of problems to try yourself.
If you have found this site useful, you might like to have a look at my
book on chemistry calculations. It covers equilibrium constant
calculations starting with the most trivial cases, and gradually getting
harder - up to the moderately difficult examples which may be asked in a
UK A' level examination.

Note: If you are interested in my chemistry calculations book


you might like to follow this link.

Where would you like to go now?


To look at Kp . . .
To the equilibrium menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/equilibria/kc.html (8 of 8)30/12/2004 11:16:24

types of catalysis

TYPES OF CATALYSIS

This page looks at the the different types of catalyst (heterogeneous and
homogeneous) with examples of each kind, and explanations of how
they work. You will also find a description of one example of
autocatalysis - a reaction which is catalysed by one of its products.

Note: This page doesn't deal with the effect of catalysts on


rates of reaction. If you don't already know about that, you
might like to follow this link first. Return to this page via the
BACK button on your browser.

Types of catalytic reactions


Catalysts can be divided into two main types - heterogeneous and
homogeneous. In a heterogeneous reaction, the catalyst is in a different
phase from the reactants. In a homogeneous reaction, the catalyst is in
the same phase as the reactants.
What is a phase?
If you look at a mixture and can see a boundary between two of the
components, those substances are in different phases. A mixture
containing a solid and a liquid consists of two phases. A mixture of
various chemicals in a single solution consists of only one phase,
because you can't see any boundary between them.

http://www.chemguide.co.uk/physical/catalysis/introduction.html (1 of 14)30/12/2004 11:16:42

types of catalysis

You might wonder why phase differs from the term physical state (solid,
liquid or gas). It includes solids, liquids and gases, but is actually a bit
more general. It can also apply to two liquids (oil and water, for example)
which don't dissolve in each other. You could see the boundary between
the two liquids.

If you want to be fussy about things, the diagrams actually show more
phases than are labelled. Each, for example, also has the glass beaker
as a solid phase. All probably have a gas above the liquid - that's another
phase. We don't count these extra phases because they aren't a part of
the reaction.

Heterogeneous catalysis
This involves the use of a catalyst in a different phase from the reactants.

http://www.chemguide.co.uk/physical/catalysis/introduction.html (2 of 14)30/12/2004 11:16:42

types of catalysis

Typical examples involve a solid catalyst with the reactants as either


liquids or gases.

Note: It is important that you remember the difference


between the two terms heterogeneous and homogeneous.
hetero implies different (as in heterosexual). Heterogeneous
catalysis has the catalyst in a different phase from the
reactants.
homo implies the same (as in homosexual). Homogeneous
catalysis has the catalyst in the same phase as the reactants.

How the heterogeneous catalyst works (in general terms)


Most examples of heterogeneous catalysis go through the same stages:
One or more of the reactants are adsorbed on to the surface of the
catalyst at active sites.
Adsorption is where something sticks to a surface. It isn't the same
as absorption where one substance is taken up within the
structure of another. Be careful!
An active site is a part of the surface which is particularly good at
adsorbing things and helping them to react.
There is some sort of interaction between the surface of the catalyst and
the reactant molecules which makes them more reactive.
This might involve an actual reaction with the surface, or some
weakening of the bonds in the attached molecules.
The reaction happens.
At this stage, both of the reactant molecules might be attached to
the surface, or one might be attached and hit by the other one
moving freely in the gas or liquid.
http://www.chemguide.co.uk/physical/catalysis/introduction.html (3 of 14)30/12/2004 11:16:42

types of catalysis

The product molecules are desorbed.


Desorption simply means that the product molecules break away.
This leaves the active site available for a new set of molecules to
attach to and react.
A good catalyst needs to adsorb the reactant molecules strongly enough
for them to react, but not so strongly that the product molecules stick
more or less permanently to the surface.
Silver, for example, isn't a good catalyst because it doesn't form strong
enough attachments with reactant molecules. Tungsten, on the other
hand, isn't a good catalyst because it adsorbs too strongly.
Metals like platinum and nickel make good catalysts because they
adsorb strongly enough to hold and activate the reactants, but not so
strongly that the products can't break away.

Examples of heterogeneous catalysis


The hydrogenation of a carbon-carbon double bond
The simplest example of this is the reaction between ethene and
hydrogen in the presence of a nickel catalyst.

In practice, this is a pointless reaction, because you are converting the


extremely useful ethene into the relatively useless ethane. However, the
same reaction will happen with any compound containing a carboncarbon double bond.
One important industrial use is in the hydrogenation of vegetable oils to
make margarine, which also involves reacting a carbon-carbon double
bond in the vegetable oil with hydrogen in the presence of a nickel
catalyst.
Ethene molecules are adsorbed on the surface of the nickel. The double
http://www.chemguide.co.uk/physical/catalysis/introduction.html (4 of 14)30/12/2004 11:16:42

types of catalysis

bond between the carbon atoms breaks and the electrons are used to
bond it to the nickel surface.

Hydrogen molecules are also adsorbed on to the surface of the nickel.


When this happens, the hydrogen molecules are broken into atoms.
These can move around on the surface of the nickel.

If a hydrogen atom diffuses close to one of the bonded carbons, the bond
between the carbon and the nickel is replaced by one between the
carbon and hydrogen.

That end of the original ethene now breaks free of the surface, and
eventually the same thing will happen at the other end.

http://www.chemguide.co.uk/physical/catalysis/introduction.html (5 of 14)30/12/2004 11:16:42

types of catalysis

As before, one of the hydrogen atoms forms a bond with the carbon, and
that end also breaks free. There is now space on the surface of the
nickel for new reactant molecules to go through the whole process again.

Note: Several metals, including nickel, have the ability to


absorb hydrogen into their structure as well as adsorb it on to
the surface. In these cases, the hydrogen molecules are also
converted into atoms which can diffuse through the metal
structure.
This happens with nickel if the hydrogen is under high
pressures, but I haven't been able to find any information
about whether it is also absorbed under the lower pressures
usually used for these hydrogenation reactions. I have
therefore stuck with the usual explanation in terms of
adsorption.
If anyone has any firm information about this, could they
contact me via the address on the about this site page.

http://www.chemguide.co.uk/physical/catalysis/introduction.html (6 of 14)30/12/2004 11:16:42

types of catalysis

Catalytic converters
Catalytic converters change poisonous molecules like carbon monoxide
and various nitrogen oxides in car exhausts into more harmless
molecules like carbon dioxide and nitrogen. They use expensive metals
like platinum, palladium and rhodium as the heterogeneous catalyst.
The metals are deposited as thin layers onto a ceramic honeycomb. This
maximises the surface area and keeps the amount of metal used to a
minimum.
Taking the reaction between carbon monoxide and nitrogen monoxide as
typical:

In the same sort of way as the previous example, the carbon monoxide
and nitrogen monoxide will be adsorbed on the surface of the catalyst,
where they react. The carbon dioxide and nitrogen are then desorbed.

The use of vanadium(V) oxide in the Contact Process


During the Contact Process for manufacturing sulphuric acid, sulphur
dioxide has to be converted into sulphur trioxide. This is done by passing
sulphur dioxide and oxygen over a solid vanadium(V) oxide catalyst.

Note: The equation is written with the half in it to make the


explanation below tidier. You may well be familiar with the
equation written as twice that shown, but the present version
is perfectly acceptable. It is also shown as a one-way rather
than a reversible reaction to avoid complicating things.

http://www.chemguide.co.uk/physical/catalysis/introduction.html (7 of 14)30/12/2004 11:16:42

types of catalysis

This example is slightly different from the previous ones because the
gases actually react with the surface of the catalyst, temporarily changing
it. It is a good example of the ability of transition metals and their
compounds to act as catalysts because of their ability to change their
oxidation state.

Note: If you aren't sure about oxidation states, it might be


useful to follow this link before you go on. Use the BACK
button on your browser to return to this page.

The sulphur dioxide is oxidised to sulphur trioxide by the vanadium(V)


oxide. In the process, the vanadium(V) oxide is reduced to vanadium(IV)
oxide.

The vanadium(IV) oxide is then re-oxidised by the oxygen.

This is a good example of the way that a catalyst can be changed during
the course of a reaction. At the end of the reaction, though, it will be
chemically the same as it started.

Note: If you want more detail about the Contact Process,


you will find a full description of the conditions used and the
reasons for them by following this link.

http://www.chemguide.co.uk/physical/catalysis/introduction.html (8 of 14)30/12/2004 11:16:42

types of catalysis

Homogeneous catalysis
This has the catalyst in the same phase as the reactants. Typically
everything will be present as a gas or contained in a single liquid phase.
The examples contain one of each of these . . .
Examples of homogeneous catalysis
The reaction between persulphate ions and iodide ions
This is a solution reaction that you may well only meet in the context of
catalysis, but it is a lovely example!
Persulphate ions (peroxodisulphate ions), S2O82-, are very powerful
oxidising agents. Iodide ions are very easily oxidised to iodine. And yet
the reaction between them in solution in water is very slow.
If you look at the equation, it is easy to see why that is:

The reaction needs a collision between two negative ions. Repulsion is


going to get seriously in the way of that!
The catalysed reaction avoids that problem completely. The catalyst can
be either iron(II) or iron(III) ions which are added to the same solution.
This is another good example of the use of transition metal compounds
as catalysts because of their ability to change oxidation state.
For the sake of argument, we'll take the catalyst to be iron(II) ions. As
you will see shortly, it doesn't actually matter whether you use iron(II) or
iron(III) ions.
The persulphate ions oxidise the iron(II) ions to iron(III) ions. In the
process the persulphate ions are reduced to sulphate ions.

http://www.chemguide.co.uk/physical/catalysis/introduction.html (9 of 14)30/12/2004 11:16:42

types of catalysis

The iron(III) ions are strong enough oxidising agents to oxidise iodide
ions to iodine. In the process, they are reduced back to iron(II) ions again.

Both of these individual stages in the overall reaction involve collision


between positive and negative ions. This will be much more likely to be
successful than collision between two negative ions in the uncatalysed
reaction.
What happens if you use iron(III) ions as the catalyst instead of iron(II)
ions? The reactions simply happen in a different order.

The destruction of atmospheric ozone


This is a good example of homogeneous catalysis where everything is
present as a gas.
Ozone, O3, is constantly being formed and broken up again in the high
atmosphere by the action of ultraviolet light. Ordinary oxygen molecules
absorb ultraviolet light and break into individual oxygen atoms. These
have unpaired electrons, and are known as free radicals. They are very
reactive.

The oxygen radicals can then combine with ordinary oxygen molecules
to make ozone.

Ozone can also be split up again into ordinary oxygen and an oxygen
radical by absorbing ultraviolet light.

http://www.chemguide.co.uk/physical/catalysis/introduction.html (10 of 14)30/12/2004 11:16:42

types of catalysis

This formation and breaking up of ozone is going on all the time. Taken
together, these reactions stop a lot of harmful ultraviolet radiation
penetrating the atmosphere to reach the surface of the Earth.
The catalytic reaction we are interested in destroys the ozone and so
stops it absorbing UV in this way.
Chlorofluorocarbons (CFCs) like CF2Cl2, for example, were used
extensively in aerosols and as refrigerants. Their slow breakdown in the
atmosphere produces chlorine atoms - chlorine free radicals. These
catalyse the destruction of the ozone.
This happens in two stages. In the first, the ozone is broken up and a
new free radical is produced.

The chlorine radical catalyst is regenerated by a second reaction


involving an oxygen radical produced from one of the reactions we've
looked at previously.

Because the chlorine radical keeps on being regenerated, each one can
destroy thousands of ozone molecules.

Autocatalysis
The oxidation of ethanedioic acid by manganate(VII) ions
In autocatalysis, the reaction is catalysed by one of its products. One of
the simplest examples of this is in the oxidation of a solution of
ethanedioic acid (oxalic acid) by an acidified solution of potassium
manganate(VII) (potassium permanganate).

http://www.chemguide.co.uk/physical/catalysis/introduction.html (11 of 14)30/12/2004 11:16:42

types of catalysis

The reaction is very slow at room temperature. It is used as a titration to


find the concentration of potassium manganate(VII) solution and is
usually carried out at a temperature of about 60C. Even so, it is quite
slow to start with.
The reaction is catalysed by manganese(II) ions. There obviously aren't
any of those present before the reaction starts, and so it starts off
extremely slowly at room temperature. However, if you look at the
equation, you will find manganese(II) ions amongst the products. More
and more catalyst is produced as the reaction proceeds and so the
reaction speeds up.
You can measure this effect by plotting the concentration of one of the
reactants as time goes on. You get a graph quite unlike the normal rate
curve for a reaction.
Most reactions give a rate curve which looks like this:

Concentrations are high at the beginning and so the reaction is fast shown by a rapid fall in the reactant concentration. As things get used
up, the reaction slows down and eventually stops as one or more of the
reactants are completely used up.
An example of autocatalysis gives a curve like this:

http://www.chemguide.co.uk/physical/catalysis/introduction.html (12 of 14)30/12/2004 11:16:42

types of catalysis

You can see the slow (uncatalysed) reaction at the beginning. As catalyst
begins to be formed in the mixture, the reaction speeds up - getting
faster and faster as more and more catalyst is formed. Eventually, of
course, the rate falls again as things get used up.
Warning!
Don't assume that a rate curve which looks like this necessarily shows an
example of autocatalysis. There are other effects which might produce a
similar graph.
For example, if the reaction involved a solid reacting with a liquid, there
might be some sort of surface coating on the solid which the liquid has to
penetrate before the expected reaction can happen.
A more common possibility is that you have a strongly exothermic
reaction and aren't controlling the temperature properly. The heat
evolved during the reaction speeds the reaction up.

Where would you like to go now?


To the catalysis menu . . .
To the Physical Chemistry menu . . .
http://www.chemguide.co.uk/physical/catalysis/introduction.html (13 of 14)30/12/2004 11:16:42

types of catalysis

To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/catalysis/introduction.html (14 of 14)30/12/2004 11:16:42

The effect of catalysts on rates of reaction

THE EFFECT OF CATALYSTS ON REACTION


RATES

This page describes and explains the way that adding a catalyst affects
the rate of a reaction. It assumes that you are already familiar with basic
ideas about the collision theory of reaction rates, and with the MaxwellBoltzmann distribution of molecular energies in a gas.

Note: If you haven't already read the page about collision


theory, you should do so before you go on.
Use the BACK button on your browser to return to this page,
or come back via the rates of reaction menu.

Note that this is only a preliminary look at catalysis as far as it affects


rates of reaction. If you are looking for more detail, there is a separate
section dealing with catalysts which you can access via a link at the
bottom of the page.

The facts
What are catalysts?
A catalyst is a substance which speeds up a reaction, but is chemically
unchanged at the end of the reaction. When the reaction has finished,
you would have exactly the same mass of catalyst as you had at the
beginning.
Some examples
Some common examples which you may need for other parts of your
syllabus include:
http://www.chemguide.co.uk/physical/basicrates/catalyst.html (1 of 6)30/12/2004 11:16:47

The effect of catalysts on rates of reaction

reaction

catalyst

Decomposition of hydrogen peroxide

manganese(IV) oxide, MnO2

Nitration of benzene

concentrated sulphuric acid

Manufacture of ammonia by the


Haber Process

iron

Conversion of SO2 into SO3 during


the Contact Process to make
sulphuric acid

vanadium(V) oxide, V2O5

Hydrogenation of a C=C double


bond

nickel

Note: You can find details of these and other catalytic


reactions by exploring the menu for the main section on
catalysis. You will find a link at the bottom of this page.

The explanation
The key importance of activation energy
Collisions only result in a reaction if the particles collide with enough
energy to get the reaction started. This minimum energy required is
called the activation energy for the reaction.

http://www.chemguide.co.uk/physical/basicrates/catalyst.html (2 of 6)30/12/2004 11:16:47

The effect of catalysts on rates of reaction

Note: What follows assumes you have a reasonable idea


about activation energy and its relationship with the MaxwellBoltzmann distribution. This is covered on the introductory
page about collision theory.
If you aren't confident about this, follow this link, and use the
BACK button on your browser to return to this page.

You can mark the position of activation energy on a Maxwell-Boltzmann


distribution to get a diagram like this:

Only those particles represented by the area to the right of the activation
energy will react when they collide. The great majority don't have enough
energy, and will simply bounce apart.
Catalysts and activation energy
To increase the rate of a reaction you need to increase the number of
http://www.chemguide.co.uk/physical/basicrates/catalyst.html (3 of 6)30/12/2004 11:16:47

The effect of catalysts on rates of reaction

successful collisions. One possible way of doing this is to provide an


alternative way for the reaction to happen which has a lower activation
energy.
In other words, to move the activation energy on the graph like this:

Adding a catalyst has exactly this effect on activation energy. A catalyst


provides an alternative route for the reaction. That alternative route has a
lower activation energy. Showing this on an energy profile:

http://www.chemguide.co.uk/physical/basicrates/catalyst.html (4 of 6)30/12/2004 11:16:47

The effect of catalysts on rates of reaction

A word of caution!
Be very careful if you are asked about this in an exam. The correct form
of words is
"A catalyst provides an alternative route for the reaction with a
lower activation energy."
It does not "lower the activation energy of the reaction". There is a subtle
difference between the two statements that is easily illustrated with a
simple analogy.
Suppose you have a mountain between two valleys so that the only way
for people to get from one valley to the other is over the mountain. Only
the most active people will manage to get from one valley to the other.
Now suppose a tunnel is cut through the mountain. Many more people
will now manage to get from one valley to the other by this easier route.
You could say that the tunnel route has a lower activation energy than
going over the mountain.
But you haven't lowered the mountain! The tunnel has provided an
alternative route but hasn't lowered the original one. The original
mountain is still there, and some people will still choose to climb it.
In the chemistry case, if particles collide with enough energy they can still
react in exactly the same way as if the catalyst wasn't there. It is simply
http://www.chemguide.co.uk/physical/basicrates/catalyst.html (5 of 6)30/12/2004 11:16:47

The effect of catalysts on rates of reaction

that the majority of particles will react via the easier catalysed route.

Where would you like to go now?


To the more detailed catalysis menu . . .
To the rates of reaction menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/basicrates/catalyst.html (6 of 6)30/12/2004 11:16:47

An introduction to the collision theory in rates of reaction

THE COLLISION THEORY OF REACTION RATES

This page describes the collision theory of reaction rates. It concentrates


on the key things which decide whether a particular collision will result in
a reaction - in particular, the energy of the collision, and whether or not
the molecules hit each other the right way around (the orientation of the
collision).
The individual factors which affect the rate of a reaction (temperature,
concentration, and so on) are discussed on separate pages. You can get
at these via the rates of reaction menu - there is a link at the bottom of
the page.
We are going to look in detail at reactions which involve a collision
between two species.

Species: This is a useful term which covers any sort of


particle you like - molecule, ion, or free radical.

Reactions where a single species falls apart in some way are slightly
simpler because you won't be involved in worrying about the orientation
of collisions. Reactions involving collisions between more than two
species are going to be extremely uncommon (see below).

Reactions involving collisions between two species


It is pretty obvious that if you have a situation involving two species they
can only react together if they come into contact with each other. They
first have to collide, and then they may react.
Why "may react"? It isn't enough for the two species to collide - they
have to collide the right way around, and they have to collide with
enough energy for bonds to break.

http://www.chemguide.co.uk/physical/basicrates/introduction.html (1 of 8)30/12/2004 11:16:55

An introduction to the collision theory in rates of reaction

(The chances of all this happening if your reaction needed a collision


involving more than 2 particles are remote. All three (or more) particles
would have to arrive at exactly the same point in space at the same time,
with everything lined up exactly right, and having enough energy to react.
That's not likely to happen very often!)
The orientation of collision
Consider a simple reaction involving a collision between two molecules ethene, CH2=CH2, and hydrogen chloride, HCl, for example. These react
to give chloroethane.

As a result of the collision between the two molecules, the double bond
between the two carbons is converted into a single bond. A hydrogen
atom gets attached to one of the carbons and a chlorine atom to the
other.

Note: The mechanism for this reaction is dealt with on a


separate page. This might help you to understand why the
orientation of the two molecules is so important.
If you want to read a bit more about this, follow this link and
use the BACK button on your browser to return to this page.

The reaction can only happen if the hydrogen end of the H-Cl bond
approaches the carbon-carbon double bond. Any other collision between
the two molecules doesn't work. The two simply bounce off each other.

http://www.chemguide.co.uk/physical/basicrates/introduction.html (2 of 8)30/12/2004 11:16:55

An introduction to the collision theory in rates of reaction

Of the collisions shown in the diagram, only collision 1 may possibly lead
on to a reaction.
If you haven't read the page about the mechanism of the reaction, you
may wonder why collision 2 won't work as well. The double bond has a
high concentration of negative charge around it due to the electrons in
the bonds. The approaching chlorine atom is also slightly negative
because it is more electronegative than hydrogen. The repulsion simply
causes the molecules to bounce off each other.

http://www.chemguide.co.uk/physical/basicrates/introduction.html (3 of 8)30/12/2004 11:16:55

An introduction to the collision theory in rates of reaction

Note: If you aren't sure about electronegativity , you might


like to follow this link.
Use the BACK button on your browser to return to this page.

In any collision involving unsymmetrical species, you would expect that


the way they hit each other will be important in deciding whether or not a
reaction happens.

The energy of the collision


Activation Energy
Even if the species are orientated properly, you still won't get a reaction
unless the particles collide with a certain minimum energy called the
activation energy of the reaction.
Activation energy is the minimum energy required before a reaction can
occur. You can show this on an energy profile for the reaction. For a
simple over-all exothermic reaction, the energy profile looks like this:

http://www.chemguide.co.uk/physical/basicrates/introduction.html (4 of 8)30/12/2004 11:16:55

An introduction to the collision theory in rates of reaction

Note: The only difference if the reaction was endothermic


would be the relative positions of the reactants and products
lines. For an endothermic change, the products would have a
higher energy than the reactants, and so the green arrow
would be pointing upwards. It makes no difference to the
discussion about the activation energy.

If the particles collide with less energy than the activation energy, nothing
important happens. They bounce apart. You can think of the activation
energy as a barrier to the reaction. Only those collisions which have
energies equal to or greater than the activation energy result in a
reaction.
Any chemical reaction results in the breaking of some bonds (needing
energy) and the making of new ones (releasing energy). Obviously some
bonds have to be broken before new ones can be made. Activation
energy is involved in breaking some of the original bonds.
Where collisions are relatively gentle, there isn't enough energy available
to start the bond-breaking process, and so the particles don't react.

The Maxwell-Boltzmann Distribution


Because of the key role of activation energy in deciding whether a
collision will result in a reaction, it would obviously be useful to know
what sort of proportion of the particles present have high enough
energies to react when they collide.
In any system, the particles present will have a very wide range of
energies. For gases, this can be shown on a graph called the MaxwellBoltzmann Distribution which is a plot of the number of particles having
each particular energy.

http://www.chemguide.co.uk/physical/basicrates/introduction.html (5 of 8)30/12/2004 11:16:55

An introduction to the collision theory in rates of reaction

Note: The graph only applies to gases, but the conclusions


that we can draw from it can also be applied to reactions
involving liquids.

The area under the curve is a measure of the total number of particles
present.

Note: The reason for this lies in some maths beyond the
scope of an A'level chemistry course. It is important that you
remember that the area under the curve gives a count of the
number of particles even if you don't understand why!

http://www.chemguide.co.uk/physical/basicrates/introduction.html (6 of 8)30/12/2004 11:16:55

An introduction to the collision theory in rates of reaction

The Maxwell-Boltzmann Distribution and activation energy


Remember that for a reaction to happen, particles must collide with
energies equal to or greater than the activation energy for the reaction.
We can mark the activation energy on the Maxwell-Boltzmann
distribution:

Notice that the large majority of the particles don't have enough energy to
react when they collide. To enable them to react we either have to
change the shape of the curve, or move the activation energy further to
the left. This is described on other pages.

http://www.chemguide.co.uk/physical/basicrates/introduction.html (7 of 8)30/12/2004 11:16:55

An introduction to the collision theory in rates of reaction

Note: You can change the shape of the curve by changing


the temperature of the reaction. You can change the position
of the activation energy by adding a catalyst to the reaction.
You could either go straight to these pages if you are
interested, or access them later via the rates of reaction
menu (link at the bottom of the page).

Note: If you have come to this page as a UK GCSE student


(or a student on a similar introductory chemistry course
elsewhere) and want some more help, you may be interested
in my GCSE Chemistry book. This link will take you to a page
describing it.

Where would you like to go now?


To the rates of reaction menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/basicrates/introduction.html (8 of 8)30/12/2004 11:16:55

The effect of temperature on rates of reaction

THE EFFECT OF TEMPERATURE ON REACTION


RATES

This page describes and explains the way that changing the temperature
affects the rate of a reaction. It assumes that you are already familiar
with basic ideas about the collision theory, and with the MaxwellBoltzmann distribution of molecular energies in a gas.

Note: If you haven't already read the page about collision


theory, you should do so before you go on.
Use the BACK button on your browser to return to this page,
or come back via the rates of reaction menu.

The facts
What happens?
As you increase the temperature the rate of reaction increases. As a
rough approximation, for many reactions happening at around room
temperature, the rate of reaction doubles for every 10C rise in
temperature.
You have to be careful not to take this too literally. It doesn't apply to all
reactions. Even where it is approximately true, it may be that the rate
doubles every 9C or 11C or whatever. The number of degrees needed
to double the rate will also change gradually as the temperature
increases.

http://www.chemguide.co.uk/physical/basicrates/temperature.html (1 of 6)30/12/2004 11:17:01

The effect of temperature on rates of reaction

Note: You will find the effect of temperature on rate explored


in a slightly more mathematical way on a separate page.

Examples
Some reactions are virtually instantaneous - for example, a precipitation
reaction involving the coming together of ions in solution to make an
insoluble solid, or the reaction between hydrogen ions from an acid and
hydroxide ions from an alkali in solution. So heating one of these won't
make any noticeable difference to the rate of the reaction.
Almost any other reaction you care to name will happen faster if you heat
it - either in the lab, or in industry.

The explanation
Increasing the collision frequency
Particles can only react when they collide. If you heat a substance, the
particles move faster and so collide more frequently. That will speed up
the rate of reaction.
That seems a fairly straightforward explanation until you look at the
numbers!
It turns out that the frequency of two-particle collisions in gases is
proportional to the square root of the kelvin temperature. If you increase
the temperature from 293 K to 303 K (20C to 30C), you will increase
the collision frequency by a factor of:

That's an increase of 1.7% for a 10 rise. The rate of reaction will


probably have doubled for that increase in temperature - in other words,
http://www.chemguide.co.uk/physical/basicrates/temperature.html (2 of 6)30/12/2004 11:17:01

The effect of temperature on rates of reaction

an increase of about 100%. The effect of increasing collision frequency


on the rate of the reaction is very minor. The important effect is quite
different . . .

The key importance of activation energy


Collisions only result in a reaction if the particles collide with enough
energy to get the reaction started. This minimum energy required is
called the activation energy for the reaction.

Note: What follows assumes you have a reasonable idea


about activation energy and its relationship with the MaxwellBoltzmann distribution. This is covered on the introductory
page about collision theory.
If you aren't confident about this, follow this link, and use the
BACK button on your browser to return to this page.

You can mark the position of activation energy on a Maxwell-Boltzmann


distribution to get a diagram like this:

http://www.chemguide.co.uk/physical/basicrates/temperature.html (3 of 6)30/12/2004 11:17:01

The effect of temperature on rates of reaction

Only those particles represented by the area to the right of the activation
energy will react when they collide. The great majority don't have enough
energy, and will simply bounce apart.
To speed up the reaction, you need to increase the number of the very
energetic particles - those with energies equal to or greater than the
activation energy. Increasing the temperature has exactly that effect - it
changes the shape of the graph.
In the next diagram, the graph labelled T is at the original temperature.
The graph labelled T+t is at a higher temperature.

http://www.chemguide.co.uk/physical/basicrates/temperature.html (4 of 6)30/12/2004 11:17:01

The effect of temperature on rates of reaction

If you now mark the position of the activation energy, you can see that
although the curve hasn't moved very much overall, there has been such
a large increase in the number of the very energetic particles that many
more now collide with enough energy to react.

Remember that the area under a curve gives a count of the number of
particles. On the last diagram, the area under the higher temperature
curve to the right of the activation energy looks to have at least doubled therefore at least doubling the rate of the reaction.
http://www.chemguide.co.uk/physical/basicrates/temperature.html (5 of 6)30/12/2004 11:17:01

The effect of temperature on rates of reaction

Summary
Increasing the temperature increases reaction rates because of the
disproportionately large increase in the number of high energy collisions.
It is only these collisions (possessing at least the activation energy for
the reaction) which result in a reaction.

Where would you like to go now?


To the rates of reaction menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/basicrates/temperature.html (6 of 6)30/12/2004 11:17:01

rate constants and the arrhenius equation

RATE CONSTANTS AND THE ARRHENIUS


EQUATION

This page looks at the way that rate constants vary with temperature and
activation energy as shown by the Arrhenius equation.

Note: If you aren't sure what a rate constant is, you should
read the page about orders of reaction before you go on. This
present page is at the hard end of the rates of reaction work
on this site. If you aren't reasonably confident about the basic
rates of reaction work, explore the rates of reaction menu first.

The Arrhenius equation


Rate constants and rate equations
You will remember that the rate equation for a reaction between two
substances A and B looks like this:

http://www.chemguide.co.uk/physical/basicrates/arrhenius.html (1 of 7)30/12/2004 11:17:08

rate constants and the arrhenius equation

Note: If you don't remember this, you must read the page
about orders of reaction before you go on. Use the BACK
button on your browser to return to this page.

The rate equation shows the effect of changing the concentrations of the
reactants on the rate of the reaction. What about all the other things (like
temperature and catalysts, for example) which also change rates of
reaction? Where do these fit into this equation?
These are all included in the so-called rate constant - which is only
actually constant if all you are changing is the concentration of the
reactants. If you change the temperature or the catalyst, for example, the
rate constant changes.
This is shown mathematically in the Arrhenius equation.
The Arrhenius equation

What the various symbols mean


Starting with the easy ones . . .
Temperature, T
To fit into the equation, this has to be meaured in kelvin.

http://www.chemguide.co.uk/physical/basicrates/arrhenius.html (2 of 7)30/12/2004 11:17:08

rate constants and the arrhenius equation

The gas constant, R


This is a constant which comes from an equation, pV=nRT, which
relates the pressure, volume and temperature of a particular
number of moles of gas. It turns up in all sorts of unlikely places!
Activation energy, EA
This is the minimum energy needed for the reaction to occur. To fit
this into the equation, it has to be expressed in joules per mole not in kJ mol-1.

Note: If you aren't sure about activation energy, you should


read the introductory page on rates of reaction before you go
on. Use the BACK button on your browser to return to this
page.

And then the rather trickier ones . . .


e
This has a value of 2.71828 . . . and is a mathematical number, a
bit like pi. You don't need to worry exactly what it means, although
if you have to do calculations with the Arrhenius equation, you may
have to find it on your calculator. You should find an ex button probably on the same key as "ln".
The expression, e-(EA / RT)
For reasons that are beyond the scope of any course at this level,
this expression counts the fraction of the molecules present in a
gas which have energies equal to or in excess of activation energy
at a particular temperature. You will find a simple calculation
associated with this further down the page.
The frequency factor, A
You may also find this called the pre-exponential factor or the
http://www.chemguide.co.uk/physical/basicrates/arrhenius.html (3 of 7)30/12/2004 11:17:08

rate constants and the arrhenius equation

steric factor.
A is a term which includes factors like the frequency of collisions
and their orientation. It varies slightly with temperature, although
not much. It is often taken as constant across small temperature
ranges.
By this time you've probably forgotten what the original Arrhenius
equation looked like! Here it is again:

You may also come across it in a different form created by a


mathematical operation on the standard one:

"ln" is a form of logarithm. Don't worry about what it means. If you need
to use this equation, just find the "ln" button on your calculator.

Using the Arrhenius equation


The effect of a change of temperature
You can use the Arrhenius equation to show the effect of a change of
temperature on the rate constant - and therefore on the rate of the
reaction. If the rate constant doubles, for example, so also will the rate of
the reaction. Look back at the rate equation at the top of this page if you
aren't sure why that is.
What happens if you increase the temperature by 10C from, say, 20C
to 30C (293 K to 303 K)?
The frequency factor, A, in the equation is approximately constant for
such a small temperature change. We need to look at how e-(EA / RT)
changes - the fraction of molecules with energies equal to or in excess of
http://www.chemguide.co.uk/physical/basicrates/arrhenius.html (4 of 7)30/12/2004 11:17:08

rate constants and the arrhenius equation

the activation energy.


Let's assume an activation energy of 50 kJ mol-1. In the equation, we
have to write that as 50000 J mol-1. The value of the gas constant, R, is
8.31 J K-1 mol-1.
At 20C (293 K) the value of the fraction is:

By raising the temperature just a little bit (to 303 K), this increases:

You can see that the fraction of the molecules able to react has almost
doubled by increasing the temperature by 10C. That causes the rate of
reaction to almost double. This is the value in the rule-of-thumb often
used in simple rate of reaction work.

Note: This approximation (about the rate of a reaction


doubling for a 10 degree rise in temperature) only works for
reactions with activation energies of about 50 kJ mol-1 fairly
close to room temperature. If you can be bothered, use the
equation to find out what happens if you increase the
temperature from, say 1000 K to 1010 K. Work out the
expression -(EA / RT) and then use the ex button on your
calculator to finish the job.
The rate constant goes on increasing as the temperature
goes up, but the rate of increase falls off quite rapidly at
higher temperatures.

http://www.chemguide.co.uk/physical/basicrates/arrhenius.html (5 of 7)30/12/2004 11:17:08

rate constants and the arrhenius equation

The effect of a catalyst


A catalyst will provide a route for the reaction with a lower activation
energy. Suppose in the presence of a catalyst that the activation energy
falls to 25 kJ mol-1. Redoing the calculation at 293 K:

If you compare that with the corresponding value where the activation
energy was 50 kJ mol-1, you will see that there has been a massive
increase in the fraction of the molecules which are able to react. There
are almost 30000 times more molecules which can react in the presence
of the catalyst compared to having no catalyst (using our assumptions
about the activation energies).
It's no wonder catalysts speed up reactions!

Other calculations involving the Arrhenius equation


If you have values for the rate of reaction or for the rate constant at
different temperatures, you can use these to work out the activation
energy of the reaction. Only one UK A' level Exam Board expects you to
be able to do these calculations. They are included in my chemistry
calculations book, and I can't repeat the material on this site.

http://www.chemguide.co.uk/physical/basicrates/arrhenius.html (6 of 7)30/12/2004 11:17:08

rate constants and the arrhenius equation

Note: There is no way of making this sufficiently different


from what is in the book to avoid being in breach of contract
with my publishers if I included it on this site.
If you are interested in my chemistry calculations book you
might like to follow this link.

Where would you like to go now?


To the rates of reaction menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/basicrates/arrhenius.html (7 of 7)30/12/2004 11:17:08

orders of reaction and rate equations

ORDERS OF REACTION AND RATE EQUATIONS

Changing the concentration of substances taking part in a reaction


usually changes the rate of the reaction. A rate equation shows this
effect mathematically. Orders of reaction are a part of the rate equation.
This page introduces and explains the various terms you will need to
know about.

Note: If you aren't sure about why changing concentration


affects rates of reaction you might like to follow this link and
come back to this page afterwards - either via the rates of
reaction menu or by using the BACK button on your browser.

Rate equations
Measuring a rate of reaction
There are several simple ways of measuring a reaction rate. For
example, if a gas was being given off during a reaction, you could take
some measurements and work out the volume being given off per
second at any particular time during the reaction.
A rate of 2 cm3 s-1 is obviously twice as fast as one of 1 cm3 s-1.

Note: Read cm3 s-1 as "cubic centimetres per second".

http://www.chemguide.co.uk/physical/basicrates/orders.html (1 of 8)30/12/2004 11:17:14

orders of reaction and rate equations

However, for this more formal and mathematical look at rates of reaction,
the rate is usually measured by looking at how fast the concentration of
one of the reactants is falling at any one time.
For example, suppose you had a reaction between two substances A
and B. Assume that at least one of them is in a form where it is sensible
to measure its concentration - for example, in solution or as a gas.

For this reaction you could measure the rate of the reaction by finding out
how fast the concentration of, say, A was falling per second.
You might, for example, find that at the beginning of the reaction, its
concentration was falling at a rate of 0.0040 mol dm-3 s-1.

Note: Read mol dm-3 s-1 as "moles per cubic decimetre (or
litre) per second".

This means that every second the concentration of A was falling by


0.0040 moles per cubic decimetre. This rate will decrease during the
reaction as A gets used up.
Summary
For the purposes of rate equations and orders of reaction, the rate of a
reaction is measured in terms of how fast the concentration of one of the
reactants is falling. Its units are mol dm-3 s-1.

Orders of reaction
I'm not going to define what order of reaction means straight away - I'm
going to sneak up on it!
Orders of reaction are always found by doing experiments. You can't
deduce anything about the order of a reaction just by looking at the
http://www.chemguide.co.uk/physical/basicrates/orders.html (2 of 8)30/12/2004 11:17:14

orders of reaction and rate equations

equation for the reaction.


So let's suppose that you have done some experiments to find out what
happens to the rate of a reaction as the concentration of one of the
reactants, A, changes. Some of the simple things that you might find are:
One possibility: The rate of reaction is proportional to the
concentration of A
That means that if you double the concentration of A, the rate doubles as
well. If you increase the concentration of A by a factor of 4, the rate goes
up 4 times as well.
You can express this using symbols as:

Writing a formula in square brackets is a standard way of showing a


concentration measured in moles per cubic decimetre (litre).
You can also write this by getting rid of the proportionality sign and
introducing a constant, k.

Another possibility: The rate of reaction is proportional to the


square of the concentration of A
This means that if you doubled the concentration of A, the rate would go
http://www.chemguide.co.uk/physical/basicrates/orders.html (3 of 8)30/12/2004 11:17:14

orders of reaction and rate equations

up 4 times (22). If you tripled the concentration of A, the rate would


increase 9 times (32). In symbol terms:

Generalising this
By doing experiments involving a reaction between A and B, you would
find that the rate of the reaction was related to the concentrations of A
and B in this way:

This is called the rate equation for the reaction.


The concentrations of A and B have to be raised to some power to show
how they affect the rate of the reaction. These powers are called the
orders of reaction with respect to A and B.
For UK A' level purposes, the orders of reaction you are likely to meet will
be 0, 1 or 2. But other values are possible including fractional ones like
1.53, for example.
If the order of reaction with respect to A is 0 (zero), this means that the
concentration of A doesn't affect the rate of reaction. Mathematically, any
number raised to the power of zero (x0) is equal to 1. That means that
that particular term disappears from the rate equation.

http://www.chemguide.co.uk/physical/basicrates/orders.html (4 of 8)30/12/2004 11:17:14

orders of reaction and rate equations

The overall order of the reaction is found by adding up the individual


orders. For example, if the reaction is first order with respect to both A
and B (a = 1 and b = 1), the overall order is 2. We call this an overall
second order reaction.
Some examples
Each of these examples involves a reaction between A and B, and each
rate equation comes from doing some experiments to find out how the
concentrations of A and B affect the rate of reaction.
Example 1:

In this case, the order of reaction with respect to both A and B is 1. The
overall order of reaction is 2 - found by adding up the individual orders.

Note: Where the order is 1 with respect to one of the


reactants, the "1" isn't written into the equation. [A] means [A]
1.

Example 2:

This reaction is zero order with respect to A because the concentration of


A doesn't affect the rate of the reaction. The order with respect to B is 2 it's a second order reaction with respect to B. The reaction is also second
order overall (because 0 + 2 = 2).
Example 3:

This reaction is first order with respect to A and zero order with respect
to B, because the concentration of B doesn't affect the rate of the

http://www.chemguide.co.uk/physical/basicrates/orders.html (5 of 8)30/12/2004 11:17:14

orders of reaction and rate equations

reaction. The reaction is first order overall (because 1 + 0 = 1).


What if you have some other number of reactants?
It doesn't matter how many reactants there are. The concentration of
each reactant will occur in the rate equation, raised to some power.
Those powers are the individual orders of reaction. The overall order of
the reaction is found by adding them all up.

The rate constant


Surprisingly, the rate constant isn't actually a true constant! It varies, for
example, if you change the temperature of the reaction, add a catalyst, or
change the catalyst.
The rate constant is constant for a given reaction only if all you are
changing is the concentration of the reactants. You will find more about
the effect of temperature and catalysts on the rate constant on another
page.

Note: If you want to follow up this further look at rate


constants you might like to follow this link. Alternatively, you
could visit it later via the rates of reaction menu.

Calculations involving orders of reaction


You will almost certainly have to be able to calculate orders of reaction
and rate constants from given data or from your own experiments.
There are all sorts of ways of doing these sums, and it is important that
you practice the methods that your syllabus wants. Check your syllabus
and past exam papers to see what sort of examples you need to be able
to work out.

http://www.chemguide.co.uk/physical/basicrates/orders.html (6 of 8)30/12/2004 11:17:14

orders of reaction and rate equations

Note: For UK A'level students, if you haven't got copies of


your syllabus and past papers follow this link to find out how
to get hold of them.

Many text books make these sums look really difficult. In fact for A' level
purposes, the calculations are usually fairly trivial. You will find them
explained in detail in my chemistry calculations book.

Note: There are several reasons why there are very few
calculations on this site. It is much easier to learn to do sums
from a carefully organised book than from a website; I would
be in breach of my contract with my publishers if I included
material similar to what is in the book; and I need to sell a few
books to generate some income!
If you are interested in my chemistry calculations book you
might like to follow this link.

Where would you like to go now?


To a simple look at how orders of reaction are related to reaction
mechanisms . .
To the rates of reaction menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

http://www.chemguide.co.uk/physical/basicrates/orders.html (7 of 8)30/12/2004 11:17:14

orders of reaction and rate equations

Jim Clark 2002

http://www.chemguide.co.uk/physical/basicrates/orders.html (8 of 8)30/12/2004 11:17:14

The effect of concentration on rates of reaction

THE EFFECT OF CONCENTRATION ON


REACTION RATES

This page describes and explains the way that changing the
concentration of a solution affects the rate of a reaction. Be aware that
this is an introductory page only. If you are interested in orders of
reaction, you will find separate pages dealing with these. You can access
these via the rates of reaction menu (link at the bottom of the page).

The facts
What happens?
For many reactions involving liquids or gases, increasing the
concentration of the reactants increases the rate of reaction. In a few
cases, increasing the concentration of one of the reactants may have little
noticeable effect of the rate. These cases are discussed and explained
further down this page.
Don't assume that if you double the concentration of one of the reactants
that you will double the rate of the reaction. It may happen like that, but
the relationship may well be more complicated.

Note: The mathematical relationship between concentration


and rate of reaction is dealt with on the page about orders of
reaction. If you are interested, you can use this link or read
about it later via the rate of reaction menu (link at the bottom
of the page).

http://www.chemguide.co.uk/physical/basicrates/concentration.html (1 of 7)30/12/2004 11:17:22

The effect of concentration on rates of reaction

Some examples
The examples on this page all involve solutions. Changing the
concentration of a gas is achieved by changing its pressure. This is
covered on a separate page.

Note: If you want to explore the effect of changing pressure


on the rate of a reaction, you could use this link. Alternatively,
use the link to the rates of reaction menu at the bottom of this
page.

Zinc and hydrochloric acid


In the lab, zinc granules react fairly slowly with dilute hydrochloric acid,
but much faster if the acid is concentrated.

The catalytic decomposition of hydrogen peroxide


Solid manganese(IV) oxide is often used as a catalyst in this reaction.
Oxygen is given off much faster if the hydrogen peroxide is concentrated
than if it is dilute.

The reaction between sodium thiosulphate solution and


hydrochloric acid
This is a reaction which is often used to explore the relationship between
concentration and rate of reaction in introductory courses (like GCSE).
When a dilute acid is added to sodium thiosulphate solution, a pale
yellow precipitate of sulphur is formed.

As the sodium thiosulphate solution is diluted more and more, the


precipitate takes longer and longer to form.
http://www.chemguide.co.uk/physical/basicrates/concentration.html (2 of 7)30/12/2004 11:17:22

The effect of concentration on rates of reaction

The explanation
Cases where changing the concentration affects the rate of the
reaction
This is the common case, and is easily explained.
Collisions involving two particles
The same argument applies whether the reaction involves collision
between two different particles or two of the same particle.
In order for any reaction to happen, those particles must first collide. This
is true whether both particles are in solution, or whether one is in solution
and the other a solid. If the concentration is higher, the chances of
collision are greater.

http://www.chemguide.co.uk/physical/basicrates/concentration.html (3 of 7)30/12/2004 11:17:22

The effect of concentration on rates of reaction

Reactions involving only one particle


If a reaction only involves a single particle splitting up in some way, then
the number of collisions is irrelevant. What matters now is how many of
the particles have enough energy to react at any one time.

Note: If you aren't sure about this, then read the page about
collision theory and activation energy before you go on. Use
the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/physical/basicrates/concentration.html (4 of 7)30/12/2004 11:17:22

The effect of concentration on rates of reaction

Suppose that at any one time 1 in a million particles have enough energy
to equal or exceed the activation energy. If you had 100 million particles,
100 of them would react. If you had 200 million particles in the same
volume, 200 of them would now react. The rate of reaction has doubled
by doubling the concentration.

Cases where changing the concentration doesn't affect the rate of


the reaction
At first glance this seems very surprising!
Where a catalyst is already working as fast as it can
Suppose you are using a small amount of a solid catalyst in a reaction,
and a high enough concentration of reactant in solution so that the
catalyst surface was totally cluttered up with reacting particles.
Increasing the concentration of the solution even more can't have any
effect because the catalyst is already working at its maximum capacity.
In certain multi-step reactions
This is the more important effect from an A' level point of view. Suppose
you have a reaction which happens in a series of small steps. These
steps are likely to have widely different rates - some fast, some slow.
For example, suppose two reactants A and B react together in these two
stages:

The overall rate of the reaction is going to be governed by how fast A


splits up to make X and Y. This is described as the rate determining
step of the reaction.

http://www.chemguide.co.uk/physical/basicrates/concentration.html (5 of 7)30/12/2004 11:17:22

The effect of concentration on rates of reaction

If you increase the concentration of A, you will increase the chances of


this step happening for reasons we've looked at above.
If you increase the concentration of B, that will undoubtedly speed up the
second step, but that makes hardly any difference to the overall rate. You
can picture the second step as happening so fast already that as soon as
any X is formed, it is immediately pounced on by B. That second reaction
is already "waiting around" for the first one to happen.

Note: The overall rate of reaction isn't entirely independent


of the concentration of B. If you lowered its concentration
enough, you will eventually reduce the rate of the second
reaction to the point where it is similar to the rate of the first.
Both concentrations will matter if the concentration of B is low
enough.
However, for ordinary concentrations, you can say that (to a
good approximation) the overall rate of reaction is unaffected
by the concentration of B.

The best specific examples of reactions of this type comes from organic
chemistry. These involve the reaction between a tertiary halogenoalkane
(alkyl halide) and a number of possible substances - including hydroxide
ions. These are examples of nucleophilic substitution using a mechanism
known as SN1.

Note: If you are interested in exploring nucleophilic


substitution reactions further, you could follow this link.
Otherwise, you can find more about how the relationship
between concentration and rate of reaction is affected by
reaction mechanisms by exploring the topics at the bottom of
the rates of reaction menu (link below).

http://www.chemguide.co.uk/physical/basicrates/concentration.html (6 of 7)30/12/2004 11:17:22

The effect of concentration on rates of reaction

Where would you like to go now?


To the rates of reaction menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/basicrates/concentration.html (7 of 7)30/12/2004 11:17:22

The effect of pressure on rates of reaction

THE EFFECT OF PRESSURE ON REACTION


RATES

This page describes and explains the way that changing the pressure of
a gas changes the rate of a reaction.

The facts
What happens?
Increasing the pressure on a reaction involving reacting gases increases
the rate of reaction. Changing the pressure on a reaction which involves
only solids or liquids has no effect on the rate.
An example
In the manufacture of ammonia by the Haber Process, the rate of
reaction between the hydrogen and the nitrogen is increased by the use
of very high pressures.

In fact, the main reason for using high pressures is to improve the
percentage of ammonia in the equilibrium mixture, but there is a useful
effect on rate of reaction as well.

Note: If you want to explore equilibria you will find the topic
covered in a separate section of the site.

http://www.chemguide.co.uk/physical/basicrates/pressure.html (1 of 4)30/12/2004 11:17:27

The effect of pressure on rates of reaction

The explanation
The relationship between pressure and concentration
Increasing the pressure of a gas is exactly the same as increasing its
concentration. If you have a given mass of gas, the way you increase its
pressure is to squeeze it into a smaller volume. If you have the same
mass in a smaller volume, then its concentration is higher.
You can also show this relationship mathematically if you have come
across the ideal gas equation:

Rearranging this gives:

Because "RT" is constant as long as the temperature is constant, this


shows that the pressure is directly proportional to the concentration. If
you double one, you will also double the other.

http://www.chemguide.co.uk/physical/basicrates/pressure.html (2 of 4)30/12/2004 11:17:27

The effect of pressure on rates of reaction

Note: If you should be able to do calculations involving the


ideal gas equation, but aren't very happy about them, you
might be interested in my chemistry calculations book.

The effect of increasing the pressure on the rate of reaction


Collisions involving two particles
The same argument applies whether the reaction involves collision
between two different particles or two of the same particle.
In order for any reaction to happen, those particles must first collide. This
is true whether both particles are in the gas state, or whether one is a
gas and the other a solid. If the pressure is higher, the chances of
collision are greater.

http://www.chemguide.co.uk/physical/basicrates/pressure.html (3 of 4)30/12/2004 11:17:27

The effect of pressure on rates of reaction

Reactions involving only one particle


If a reaction only involves a single particle splitting up in some way, then
the number of collisions is irrelevant. What matters now is how many of
the particles have enough energy to react at any one time.

Note: If you aren't sure about this, then read the page about
collision theory and activation energy before you go on. Use
the BACK button on your browser to return to this page.

Suppose that at any one time 1 in a million particles have enough energy
to equal or exceed the activation energy. If you had 100 million particles,
100 of them would react. If you had 200 million particles in the same
volume, 200 of them would now react. The rate of reaction has doubled
by doubling the pressure.

Where would you like to go now?


To the rates of reaction menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/basicrates/pressure.html (4 of 4)30/12/2004 11:17:27

Understanding Chemistry - Physical Chemistry Menu

Understanding Chemistry

PHYSICAL CHEMISTRY MENU

Rates of Reaction
Basic ideas about rates of reaction . . .
Covers collision theory, and describes and explains the effects of
surface area, concentration, pressure, temperature and catalysts
on reaction rates. Also takes an introductory look at the rate
equation for a reaction - particularly order of reaction and the rate
constant.
Catalysis . . .
A detailed look at various types of catalytic action with a wide
range of examples taken from A level syllabuses.

Equilibria
Chemical equilibria . . .
An introduction to the ideas surrounding chemical equilibria
including the concept of a dynamic equilibrium, Le Chatelier's
Principle, and a detailed look at the equilibrium aspects of the
Haber Process, the Contact Process and the manufacture of
ethanol from ethene. Plus an introductory look at equilibrium
constants, Kc and Kp.
Redox equilibria . . .

http://www.chemguide.co.uk/physmenu.html (1 of 2)30/12/2004 11:17:29

Understanding Chemistry - Physical Chemistry Menu

An introduction to the measurement and use of standard electrode


potentials (redox potentials). Includes a discussion of the
electrochemical series and the use of redox potentials in predicting
the feasibility of redox reactions.
Acid-base equilibria . . .
Including: acid-base theories; strong and weak acids and bases;
an introduction to pH, Kw and pKw, Ka and pKa, Kb and pKb; buffer
solutions; pH curves and indicators.
Phase equilibria . . .
An introduction to saturated vapour pressure, Raoult's Law, and to
several types of phase diagrams - including phase diagrams for
pure substances, solutions of non-volatile solutes, eutectic
mixtures, systems of two miscible liquids (including fractional
distillation), and systems of two immiscible liquids (including steam
distillation). I am working on this at the moment, so not everything
is available yet.

Go to Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/physmenu.html (2 of 2)30/12/2004 11:17:29

Rates of reaction menu

Understanding Chemistry

RATES OF REACTION MENU

Basic descriptive topics


The collision theory . . .
Discusses the collision theory of reaction rates, including the
importance of activation energy and the Maxwell-Boltzmann
distribution.
The effect of surface area on rate of reaction . . .
Describes and explains the effect of surface area on the rate of a
reaction between a solid and a liquid or a gas.
The effect of concentration on rate of reaction . . .
Describes and explains the effect of changing the concentration on
the rate of a reaction involving liquids or gases. If you are
interested in rate equations, orders of reaction and rate constants,
these are explained separately - see below.
The effect of pressure on rate of reaction . . .
Describes and explains the effect of changing the pressure on the
rate of a reaction involving gases.
The effect of temperature on rate of reaction . . .
Describes and explains the effect of changing the temperature on
the rate of a reaction, including the importance of activation energy
and the Maxwell-Boltzmann distribution. The Arrhenius equation is
http://www.chemguide.co.uk/physical/basicratesmenu.html (1 of 3)30/12/2004 11:17:33

Rates of reaction menu

dealt with separately - see below.


The effect of a catalyst on rate of reaction . . .
Describes and explains the effect of adding a catalyst to a
reaction. This is an introductory look only. You will find a separate
section dealing with catalysts in detail - link via the physical
chemistry menu (see below).

Rate equations and orders of reaction


Orders of reaction and rate equations . . .
An introductory look at orders of reaction, rate equations and the
rate constant.
The relationship between order and mechanisms . . .
Looks at some simple cases to show how orders of reaction can
sometimes give useful information about the mechanism of a
reaction. Also discusses the difference between the sometimes
confusing terms "order " and "molecularity " of a reaction.
The Arrhenius equation . . .
Discusses the mathematical relationship between the rate
constant and temperature and activation energy.

Go to physical chemistry menu . . .


Go to Main Menu . . .

http://www.chemguide.co.uk/physical/basicratesmenu.html (2 of 3)30/12/2004 11:17:33

Rates of reaction menu

Jim Clark 2002

http://www.chemguide.co.uk/physical/basicratesmenu.html (3 of 3)30/12/2004 11:17:33

The effect of surface area on rates of reaction

THE EFFECT OF SURFACE AREA ON REACTION


RATES

This page describes and explains the effect of changing the surface area
of a solid on the rate of a reaction it is involved in. This applies to
reactions involving a solid and a gas, or a solid and a liquid. It includes
cases where the solid is acting as a catalyst.

The facts
What happens?
The more finely divided the solid is, the faster the reaction happens. A
powdered solid will normally produce a faster reaction than if the same
mass is present as a single lump. The powdered solid has a greater
surface area than the single lump.

Note: Why normally? What exceptions can there be?


Imagine a case of a very fine powder reacting with a gas. If
the powder was in one big heap, the gas may not be able to
penetrate it. That means that its effective surface area is
much the same as (or even less than) it would be if it were
present in a single lump.
A small heap of fine magnesium powder tends to burn rather
more slowly than a strip of magnesium ribbon, for example.

http://www.chemguide.co.uk/physical/basicrates/surfacearea.html (1 of 4)30/12/2004 11:17:43

The effect of surface area on rates of reaction

Some examples
Calcium carbonate and hydrochloric acid
In the lab, powdered calcium carbonate reacts much faster with dilute
hydrochloric acid than if the same mass was present as lumps of marble
or limestone.

The catalytic decomposition of hydrogen peroxide


This is another familiar lab reaction. Solid manganese(IV) oxide is often
used as the catalyst. Oxygen is given off much faster if the catalyst is
present as a powder than as the same mass of granules.

Catalytic converters
Catalytic converters use metals like platinum, palladium and rhodium to
convert poisonous compounds in vehicle exhausts into less harmful
things. For example, a reaction which removes both carbon monoxide
and an oxide of nitrogen is:

Because the exhaust gases are only in contact with the catalyst for a
very short time, the reactions have to be very fast. The extremely
expensive metals used as the catalyst are coated as a very thin layer
onto a ceramic honeycomb structure to maximise the surface area.

The explanation
You are only going to get a reaction if the particles in the gas or liquid
collide with the particles in the solid. Increasing the surface area of the
solid increases the chances of collision taking place.

http://www.chemguide.co.uk/physical/basicrates/surfacearea.html (2 of 4)30/12/2004 11:17:43

The effect of surface area on rates of reaction

Imagine a reaction between magnesium metal and a dilute acid like


hydrochloric acid. The reaction involves collision between magnesium
atoms and hydrogen ions.

Increasing the number of collisions per second increases the rate of


reaction.

Where would you like to go now?


To the rates of reaction menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .
http://www.chemguide.co.uk/physical/basicrates/surfacearea.html (3 of 4)30/12/2004 11:17:43

The effect of surface area on rates of reaction

Jim Clark 2002

http://www.chemguide.co.uk/physical/basicrates/surfacearea.html (4 of 4)30/12/2004 11:17:43

Catalysis menu

Understanding Chemistry

CATALYSIS MENU

An introduction to types of catalysis . . .


Explains the difference between heterogeneous and
homogeneous catalysis with examples of each. (Heterogeneous:
hydrogenation of carbon-carbon double bonds, catalytic
converters, vanadium(V) oxide in the Contact Process.
Homogeneous: the reaction between persulphate ions and iodide
ions, ozone destruction.) Also deals with an example of
autocatalysis.
Examples of catalysis in the inorganic chemical industry . . .
Catalysts in the Contact Process, the Haber Process, and the
conversion of ammonia into nitric acid.
Examples of catalysis in the petrochemical industry . . .
Examples of catalysts used in catalytic cracking, isomerisation and
reforming.
Examples of acid catalysis in organic chemistry . . .
Examples of the use of acids as catalysts in organic chemistry.
Includes the hydration of carbon-carbon double bonds to make
alcohols, esterification and its reverse (the acid catalysed
hydrolysis of esters), and the nitration of benzene.
Examples of other catalytic reactions in organic chemistry . . .
Includes the formation of epoxyethane from ethene, Friedel-Crafts
http://www.chemguide.co.uk/physical/catalysismenu.html (1 of 2)30/12/2004 11:17:45

Catalysis menu

reactions, and the halogenation of benzene.

Go to physical chemistry menu . . .


Go to Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/catalysismenu.html (2 of 2)30/12/2004 11:17:45

catalysts in inorganic industrial processes

CATALYSTS IN INORGANIC INDUSTRIAL


PROCESSES

This page takes a brief look at the catalysts used in the Contact Process
to manufacture sulphuric acid, in the Haber Process to manufacture
ammonia, and in the conversion of ammonia into nitric acid.
If you want full details of the Contact Process or the Haber Process, you
will find links below. Full details of the manufacture of nitric acid are given
on this page because they don't appear elsewhere on this site.

The Contact Process for the manufacture of sulphuric acid


At the heart of the Contact Process is a reaction which converts sulphur
dioxide into sulphur trioxide. Sulphur dioxide gas is passed together with
air (as a source of oxygen) over a solid vanadium(V) oxide catalyst. This
is therefore an example of heterogeneous catalysis.

The fact that this is a reversible reaction makes no difference to the


operation of the catalyst. It speeds up both the forward reaction and the
back reaction by the same amount.

Note: If you don't understand the term heterogeneous


catalysis, or want the mechanism for this reaction, follow this
link to the introductory page on catalysis.
If you want full details of the Contact Process, including the
reasons for all the conditions, you will find it by following this
link.

http://www.chemguide.co.uk/physical/catalysis/inorganic.html (1 of 3)30/12/2004 11:17:53

catalysts in inorganic industrial processes

The Haber Process for the manufacture of ammonia


The Haber Process combines hydrogen and nitrogen to make ammonia
using an iron catalyst. This is another reversible reaction, and another
example of heterogeneous catalysis.

Note: If you want full details of the Haber Process, including


the reasons for all the conditions, you will find it by following
this link.

The manufacture of nitric acid from ammonia


This is yet another example of heterogeneous catalysis.
This process involves oxidation of the ammonia from the Haber Process
by oxygen in the air in the presence of a platinum-rhodium catalyst.
Large sheets of metal gauze are used in order to reduce expense and to
maximise the surface area of the catalyst. Although in principle the
sheets would last for ever because the metals are acting as a catalyst, in
practice they do deteriorate over time and have to be replaced.
The sheets of gauze are held at a temperature of about 900C. The
reaction is very exothermic, and once it starts the temperature is
maintained by the heat evolved.
The ammonia is oxidised to nitrogen monoxide gas.

This is cooled. At ordinary temperatures and in the presence of excess


air, it is oxidised further to nitrogen dioxide.

http://www.chemguide.co.uk/physical/catalysis/inorganic.html (2 of 3)30/12/2004 11:17:53

catalysts in inorganic industrial processes

The nitrogen dioxide (still in the presence of excess air) is absorbed in


water where it reacts to give a concentrated solution of nitric acid.

Where would you like to go now?


To the catalysis menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/catalysis/inorganic.html (3 of 3)30/12/2004 11:17:53

The Contact Process for the manufacture of sulphuric acid

THE CONTACT PROCESS

This page describes the Contact Process for the manufacture of


sulphuric acid, and then goes on to explain the reasons for the conditions
used in the process. It looks at the effect of proportions, temperature,
pressure and catalyst on the composition of the equilibrium mixture, the
rate of the reaction and the economics of the process.

Important: If you aren't sure about using Le Chatelier's


Principle or about the effect of changing conditions on rates
of reaction you should explore these links before you go on.
When you are reading this page, if you find that you aren't
understanding the effect of changing one of the conditions on
the position of equilibrium or on the rate of the reaction, come
back and follow up these links.

A brief summary of the Contact Process


The Contact Process:

makes sulphur dioxide;


convers the sulphur dioxide into sulphur trioxide (the reversible
reaction at the heart of the process);
converts the sulphur trioxide into concentrated sulphuric acid.

Making the sulphur dioxide

http://www.chemguide.co.uk/physical/equilibria/contact.html (1 of 7)30/12/2004 11:18:18

The Contact Process for the manufacture of sulphuric acid

This can either be made by burning sulphur in an excess of air:

. . . or by heating sulphide ores like pyrite in an excess of air:

In either case, an excess of air is used so that the sulphur dioxide


produced is already mixed with oxygen for the next stage.
Converting the sulphur dioxide into sulphur trioxide
This is a reversible reaction, and the formation of the sulphur trioxide is
exothermic.

A flow scheme for this part of the process looks like this:

http://www.chemguide.co.uk/physical/equilibria/contact.html (2 of 7)30/12/2004 11:18:18

The Contact Process for the manufacture of sulphuric acid

The reasons for all these conditions will be explored in detail further
down the page.
Converting the sulphur trioxide into sulphuric acid
This can't be done by simply adding water to the sulphur trioxide - the
reaction is so uncontrollable that it creates a fog of sulphuric acid.
Instead, the sulphur trioxide is first dissolved in concentrated sulphuric
acid:

The product is known as fuming sulphuric acid or oleum.


This can then be reacted safely with water to produce concentrated
http://www.chemguide.co.uk/physical/equilibria/contact.html (3 of 7)30/12/2004 11:18:18

The Contact Process for the manufacture of sulphuric acid

sulphuric acid - twice as much as you originally used to make the fuming
sulphuric acid.

Explaining the conditions


The proportions of sulphur dioxide and oxygen
The mixture of sulphur dioxide and oxygen going into the reactor is in
equal proportions by volume.
Avogadro's Law says that equal volumes of gases at the same
temperature and pressure contain equal numbers of molecules. That
means that the gases are going into the reactor in the ratio of 1 molecule
of sulphur dioxide to 1 of oxygen.
That is an excess of oxygen relative to the proportions demanded by the
equation.

According to Le Chatelier's Principle, Increasing the concentration of


oxygen in the mixture causes the position of equilibrium to shift towards
the right. Since the oxygen comes from the air, this is a very cheap way
of increasing the conversion of sulphur dioxide into sulphur trioxide.
Why not use an even higher proportion of oxygen? This is easy to see if
you take an extreme case. Suppose you have a million molecules of
oxygen to every molecule of sulphur dioxide.
The equilibrium is going to be tipped very strongly towards sulphur
trioxide - virtually every molecule of sulphur dioxide will be converted into
sulphur trioxide. Great! But you aren't going to produce much sulphur
trioxide every day. The vast majority of what you are passing over the
catalyst is oxygen which has nothing to react with.
By increasing the proportion of oxygen you can increase the percentage
of the sulphur dioxide converted, but at the same time decrease the total

http://www.chemguide.co.uk/physical/equilibria/contact.html (4 of 7)30/12/2004 11:18:18

The Contact Process for the manufacture of sulphuric acid

amount of sulphur trioxide made each day. The 1 : 1 mixture turns out to
give you the best possible overall yield of sulphur trioxide.

The temperature
Equilibrium considerations
You need to shift the position of the equilibrium as far as possible to the
right in order to produce the maximum possible amount of sulphur
trioxide in the equilibrium mixture.
The forward reaction (the production of sulphur trioxide) is exothermic.

According to Le Chatelier's Principle, this will be favoured if you lower the


temperature. The system will respond by moving the position of
equilibrium to counteract this - in other words by producing more heat.
In order to get as much sulphur trioxide as possible in the equilibrium
mixture, you need as low a temperature as possible. However, 400 - 450
C isn't a low temperature!
Rate considerations
The lower the temperature you use, the slower the reaction becomes. A
manufacturer is trying to produce as much sulphur trioxide as possible
per day. It makes no sense to try to achieve an equilibrium mixture which
contains a very high proportion of sulphur trioxide if it takes several years
for the reaction to reach that equilibrium.
You need the gases to reach equilibrium within the very short time that
they will be in contact with the catalyst in the reactor.
The compromise
400 - 450C is a compromise temperature producing a fairly high
proportion of sulphur trioxide in the equilibrium mixture, but in a very
short time.

http://www.chemguide.co.uk/physical/equilibria/contact.html (5 of 7)30/12/2004 11:18:18

The Contact Process for the manufacture of sulphuric acid

The pressure
Equilibrium considerations

Notice that there are 3 molecules on the left-hand side of the equation,
but only 2 on the right.
According to Le Chatelier's Principle, if you increase the pressure the
system will respond by favouring the reaction which produces fewer
molecules. That will cause the pressure to fall again.
In order to get as much sulphur trioxide as possible in the equilibrium
mixture, you need as high a pressure as possible. High pressures also
increase the rate of the reaction. However, the reaction is done at
pressures close to atmospheric pressure!
Economic considerations
Even at these relatively low pressures, there is a 99.5% conversion of
sulphur dioxide into sulphur trioxide. The very small improvement that
you could achieve by increasing the pressure isn't worth the expense of
producing those high pressures.

The catalyst
Equilibrium considerations
The catalyst has no effect whatsoever on the position of the equilibrium.
Adding a catalyst doesn't produce any greater percentage of sulphur
trioxide in the equilibrium mixture. Its only function is to speed up the
reaction.
Rate considerations
In the absence of a catalyst the reaction is so slow that virtually no
http://www.chemguide.co.uk/physical/equilibria/contact.html (6 of 7)30/12/2004 11:18:18

The Contact Process for the manufacture of sulphuric acid

reaction happens in any sensible time. The catalyst ensures that the
reaction is fast enough for a dynamic equilibrium to be set up within the
very short time that the gases are actually in the reactor.

Note: If you are interested in the mechanism for the catalytic


reaction you will find it on an introductory page on types of
catalysts.
Use the BACK button on your browser if you want to return to
this page.

Where would you like to go now?


To the equilibrium menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/equilibria/contact.html (7 of 7)30/12/2004 11:18:18

The Haber Process for the manufacture of ammonia

THE HABER PROCESS

This page describes the Haber Process for the manufacture of ammonia
from nitrogen and hydrogen, and then goes on to explain the reasons for
the conditions used in the process. It looks at the effect of temperature,
pressure and catalyst on the composition of the equilibrium mixture, the
rate of the reaction and the economics of the process.

Important: If you aren't sure about using Le Chatelier's


Principle or about the effect of changing conditions on rates
of reaction you should explore these links before you go on.
When you are reading this page, if you find that you aren't
understanding the effect of changing one of the conditions on
the position of equilibrium or on the rate of the reaction, come
back and follow up these links.

A brief summary of the Haber Process


The Haber Process combines nitrogen from the air with hydrogen
derived mainly from natural gas (methane) into ammonia. The reaction is
reversible and the production of ammonia is exothermic.

A flow scheme for the Haber Process looks like this:

http://www.chemguide.co.uk/physical/equilibria/haber.html (1 of 6)30/12/2004 11:18:26

The Haber Process for the manufacture of ammonia

Some notes on the conditions


The catalyst
The catalyst is actually slightly more complicated than pure iron. It has
potassium hydroxide added to it as a promoter - a substance that
increases its efficiency.
The pressure
The pressure varies from one manufacturing plant to another, but is
always high. You can't go far wrong in an exam quoting 200
atmospheres.
Recycling
At each pass of the gases through the reactor, only about 15% of the
nitrogen and hydrogen converts to ammonia. (This figure also varies
from plant to plant.) By continual recycling of the unreacted nitrogen and
hydrogen, the overall conversion is about 98%.

Explaining the conditions


http://www.chemguide.co.uk/physical/equilibria/haber.html (2 of 6)30/12/2004 11:18:26

The Haber Process for the manufacture of ammonia

The proportions of nitrogen and hydrogen


The mixture of nitrogen and hydrogen going into the reactor is in the ratio
of 1 volume of nitrogen to 3 volumes of hydrogen.
Avogadro's Law says that equal volumes of gases at the same
temperature and pressure contain equal numbers of molecules. That
means that the gases are going into the reactor in the ratio of 1 molecule
of nitrogen to 3 of hydrogen.
That is the proportion demanded by the equation.
In some reactions you might choose to use an excess of one of the
reactants. You would do this if it is particularly important to use up as
much as possible of the other reactant - if, for example, it was much
more expensive. That doesn't apply in this case.
There is always a down-side to using anything other than the equation
proportions. If you have an excess of one reactant there will be
molecules passing through the reactor which can't possibly react
because there isn't anything for them to react with. This wastes reactor
space - particularly space on the surface of the catalyst.

The temperature
Equilibrium considerations
You need to shift the position of the equilibrium as far as possible to the
right in order to produce the maximum possible amount of ammonia in
the equilibrium mixture.
The forward reaction (the production of ammonia) is exothermic.

According to Le Chatelier's Principle, this will be favoured if you lower the


temperature. The system will respond by moving the position of
equilibrium to counteract this - in other words by producing more heat.

http://www.chemguide.co.uk/physical/equilibria/haber.html (3 of 6)30/12/2004 11:18:26

The Haber Process for the manufacture of ammonia

In order to get as much ammonia as possible in the equilibrium mixture,


you need as low a temperature as possible. However, 400 - 450C isn't a
low temperature!
Rate considerations
The lower the temperature you use, the slower the reaction becomes. A
manufacturer is trying to produce as much ammonia as possible per day.
It makes no sense to try to achieve an equilibrium mixture which contains
a very high proportion of ammonia if it takes several years for the
reaction to reach that equilibrium.
You need the gases to reach equilibrium within the very short time that
they will be in contact with the catalyst in the reactor.
The compromise
400 - 450C is a compromise temperature producing a reasonably high
proportion of ammonia in the equilibrium mixture (even if it is only 15%),
but in a very short time.

The pressure
Equilibrium considerations

Notice that there are 4 molecules on the left-hand side of the equation,
but only 2 on the right.
According to Le Chatelier's Principle, if you increase the pressure the
system will respond by favouring the reaction which produces fewer
molecules. That will cause the pressure to fall again.
In order to get as much ammonia as possible in the equilibrium mixture,
you need as high a pressure as possible. 200 atmospheres is a high
pressure, but not amazingly high.
Rate considerations
http://www.chemguide.co.uk/physical/equilibria/haber.html (4 of 6)30/12/2004 11:18:26

The Haber Process for the manufacture of ammonia

Increasing the pressure brings the molecules closer together. In this


particular instance, it will increase their chances of hitting and sticking to
the surface of the catalyst where they can react. The higher the pressure
the better in terms of the rate of a gas reaction.
Economic considerations
Very high pressures are very expensive to produce on two counts.
You have to build extremely strong pipes and containment vessels to
withstand the very high pressure. That increases your capital costs when
the plant is built.
High pressures cost a lot to produce and maintain. That means that the
running costs of your plant are very high.
The compromise
200 atmospheres is a compromise pressure chosen on economic
grounds. If the pressure used is too high, the cost of generating it
exceeds the price you can get for the extra ammonia produced.

The catalyst
Equilibrium considerations
The catalyst has no effect whatsoever on the position of the equilibrium.
Adding a catalyst doesn't produce any greater percentage of ammonia in
the equilibrium mixture. Its only function is to speed up the reaction.
Rate considerations
In the absence of a catalyst the reaction is so slow that virtually no
reaction happens in any sensible time. The catalyst ensures that the
reaction is fast enough for a dynamic equilibrium to be set up within the
very short time that the gases are actually in the reactor.

http://www.chemguide.co.uk/physical/equilibria/haber.html (5 of 6)30/12/2004 11:18:26

The Haber Process for the manufacture of ammonia

Separating the ammonia


When the gases leave the reactor they are hot and at a very high
pressure. Ammonia is easily liquefied under pressure as long as it isn't
too hot, and so the temperature of the mixture is lowered enough for the
ammonia to turn to a liquid. The nitrogen and hydrogen remain as gases
even under these high pressures, and can be recycled.

Where would you like to go now?


To the equilibrium menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/equilibria/haber.html (6 of 6)30/12/2004 11:18:26

catalysts in the petrochemical industry

CATALYSTS IN THE PETROCHEMICAL


INDUSTRY

This page looks briefly at some of the basic processes in the


petrochemical industry (cracking, isomerisation and reforming) as
examples of important catalytic reactions.

Catalytic cracking
Cracking is the name given to breaking up large hydrocarbon molecules
into smaller and more useful bits. This is achieved by using high
pressures and temperatures without a catalyst, or lower temperatures and
pressures in the presence of a catalyst.
The source of the large hydrocarbon molecules is often the naphtha
fraction or the gas oil fraction from the fractional distillation of crude oil
(petroleum). These fractions are obtained from the distillation process as
liquids, but are re-vaporised before cracking.
The hydrocarbons are mixed with a very fine catalyst powder. These days
the catalysts are zeolites (complex aluminosilicates) - these are more
efficient than the older mixtures of aluminium oxide and silicon dioxide.
The whole mixture is blown rather like a liquid through a reaction chamber
at a temperature of about 500C. Because the mixture behaves like a
liquid, this is known as fluid catalytic cracking (or fluidised catalytic
cracking).
Although the mixture of gas and fine solid behaves as a liquid, this is
nevertheless an example of heterogeneous catalysis - the catalyst is in a
different phase from the reactants.

http://www.chemguide.co.uk/physical/catalysis/petrochem.html (1 of 7)30/12/2004 11:18:46

catalysts in the petrochemical industry

Note: If you don't understand the term heterogeneous


catalysis, follow this link to the introductory page on catalysis.
Use the BACK button on your browser to return quickly to this
page.

The catalyst is recovered afterwards, and the cracked mixture is


separated by cooling and further fractional distillation.
There isn't any single unique reaction happening in the cracker. The
hydrocarbon molecules are broken up in a fairly random way to produce
mixtures of smaller hydrocarbons, some of which have carbon-carbon
double bonds. One possible reaction involving the hydrocarbon C15H32
might be:

Or, showing more clearly what happens to the various atoms and bonds:

This is only one way in which this particular molecule might break up. The
ethene and propene are important materials for making plastics or
http://www.chemguide.co.uk/physical/catalysis/petrochem.html (2 of 7)30/12/2004 11:18:46

catalysts in the petrochemical industry

producing other organic chemicals. The octane is one of the molecules


found in petrol (gasoline).

Isomerisation
Hydrocarbons used in petrol (gasoline) are given an octane rating which
relates to how effectively they perform in the engine. A hydrocarbon with a
high octane rating burns more smoothly than one with a low octane rating.
Molecules with "straight chains" have a tendency to pre-ignition. When the
petrol / air mixture is compressed they tend to explode, and then explode
a second time when the spark is passed through them. This double
explosion produces knocking in the engine.

Note: A straight chain hydrocarbon isn't literally straight! It


just means that all the carbon atoms are joined up one after
another in a row. If you made a model it would be extremely
bendy!

Octane ratings are based on a scale on which heptane is given a rating of


0, and 2,2,4-trimethylpentane (an isomer of octane) a rating of 100.

http://www.chemguide.co.uk/physical/catalysis/petrochem.html (3 of 7)30/12/2004 11:18:46

catalysts in the petrochemical industry

Note: If you don't understand what is meant by the term


structural isomer you really ought to follow this link before you
go on.
If you aren't happy about naming organic compounds, you
might like to follow up this link at some time in the future. It
isn't particularly important for the purposes of understanding
the current page.
Use the BACK button on your browser to return to this page.

In order to raise the octane rating of the molecules found in petrol


(gasoline) and so make the petrol burn better in modern engines, the oil
industry rearranges straight chain molecules into their isomers with
branched chains.
One process uses a platinum catalyst on a zeolite base at a temperature
of about 250C and a pressure of 13 - 30 atmospheres. It is used
particularly to change straight chains containing 5 or 6 carbon atoms into
their branched isomers.
For example:

Reforming
Reforming is another process used to improve the octane rating of
hydrocarbons to be used in petrol, but is also a useful source of aromatic
compounds for the chemical industry. Aromatic compounds are ones
based on a benzene ring.

http://www.chemguide.co.uk/physical/catalysis/petrochem.html (4 of 7)30/12/2004 11:18:46

catalysts in the petrochemical industry

Note: If you aren't familiar with the structure of benzene you


might like to have a quick look at this link before you go on.
There's no need to get too bogged down in it though for the
purposes of the current page.
Use the BACK button on your browser to return to this page.

Reforming uses a platinum catalyst suspended on aluminium oxide


together with various promoters to make the catalyst more efficient. The
original molecules are passed as vapours over the solid catalyst at a
temperature of about 500C.
Isomerisation reactions occur (as above) but, in addition, chain molecules
get converted into rings with the loss of hydrogen. Hexane, for example,
gets converted into benzene, and heptane into methylbenzene.

http://www.chemguide.co.uk/physical/catalysis/petrochem.html (5 of 7)30/12/2004 11:18:46

catalysts in the petrochemical industry

Note: The Kekul structure has been used for benzene


because it is easier to understand if you aren't familiar with
benzene chemistry. You could perfectly well use the standard
symbol of a circle inside a hexagon instead.
The original hexane and heptane molecules have been drawn
bent into ring shapes so that you can easily see how they are
changed during reforming.

http://www.chemguide.co.uk/physical/catalysis/petrochem.html (6 of 7)30/12/2004 11:18:46

catalysts in the petrochemical industry

Where would you like to go now?


To the catalysis menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/catalysis/petrochem.html (7 of 7)30/12/2004 11:18:46

acid catalysis in organic chemistry

ACID CATALYSIS IN ORGANIC CHEMISTRY

This page looks at the use of acid catalysts in some organic reactions. It
covers the nitration of benzene, the hydration of ethene to manufacture
ethanol, and the reactions both to produce esters and to hydrolyse them
under acidic conditions. You will find links to the full mechanisms for
each of these reactions if you want them.

The nitration of benzene


Benzene is treated with a mixture of concentrated nitric acid and
concentrated sulphuric acid at a temperature not exceeding 50C. As the
temperature increases there is a greater chance of getting more than one
nitro group, -NO2, substituted onto the ring.
Nitrobenzene is formed.

or:

The concentrated sulphuric acid is acting as a catalyst. Because


everything is present in the same liquid phase, this is a good example of
homogeneous catalysis.

http://www.chemguide.co.uk/physical/catalysis/acidcat.html (1 of 5)30/12/2004 11:19:05

acid catalysis in organic chemistry

Note: If you don't understand the term homogeneous


catalysis, follow this link to the introductory page on catalysis.
If you want the mechanism for the nitration of benzene, you
will find it on another part of this site.
Use the BACK button on your browser if you wish to return to
this page.

The hydration of ethene to make ethanol


Ethene is mixed with steam and passed over a catalyst consisting of
solid silicon dioxide coated with phosphoric(V) acid. The temperature
used is 300C and the pressure is about 60 to 70 atmospheres.
Because the catalyst is in a different phase from the reactants, this is an
example of heterogeneous catalysis.

This is a reversible reaction and only about 5% of the ethene reacts on


each pass over the catalyst. When the reaction mixture is cooled, the
ethanol and any excess steam condense, and the gaseous ethene can
be recycled through the process.
A conversion rate of about 95% is achieved by continual recycling in this
way.

http://www.chemguide.co.uk/physical/catalysis/acidcat.html (2 of 5)30/12/2004 11:19:05

acid catalysis in organic chemistry

Note: If you are interested in the reasons for the conditions


used in this reaction, you will find them in the equilibrium
section of this site by following this link.
If you are interested in the mechanism for the hydration of
ethene, follow this link.
Use the BACK button on your browser if you wish to return to
this page.

Making esters - the esterification reaction


Esters are what is formed when an organic acid reacts with an alcohol in
the presence of concentrated sulphuric acid as the catalyst. Everything is
present in a single liquid phase, and so this is an example of
homogeneous catalysis.
For example, ethanoic acid reacts with ethanol to produce ethyl
ethanoate.

Note: If you aren't happy about the names of esters, you will
find them discussed if you follow this link.
Use the BACK button on your browser to return to this page
afterwards.

The ethyl ethanoate has the lowest boiling point of anything in the
mixture, and so is distilled off as soon as it is formed. This helps to
reduce the reverse reaction.

http://www.chemguide.co.uk/physical/catalysis/acidcat.html (3 of 5)30/12/2004 11:19:05

acid catalysis in organic chemistry

Note: If you are interested in the mechanism for the


esterification reaction, follow this link.
Use the BACK button on your browser if you wish to return to
this page.

The acid catalysed hydrolysis of esters


In principle, this is the reverse of the esterification reaction but, in
practice, it has to be done slightly differently. The ester is heated under
reflux with a dilute acid such as dilute hydrochloric acid or dilute sulphuric
acid.

Heating under reflux: This involves heating the mixture in a


flask with a condenser placed vertically in the neck. Any
escaping vapours condense and run back into the flask.

The equation for the reaction is simply the esterification equation written
backwards.

The dilute acid used as the catalyst also provides the water for the
reaction. You need a large excess of water in order to increase the
chances of the forward reaction happening and the ester hydrolysing.
You would normally hydrolyse esters quite differently by heating them
with sodium hydroxide solution (alkaline hydrolysis). This isn't an
example of a catalytic reaction because the hydroxide ions are used up
during the reaction.

http://www.chemguide.co.uk/physical/catalysis/acidcat.html (4 of 5)30/12/2004 11:19:05

acid catalysis in organic chemistry

The main advantage of doing it like this is that it is a one-way reaction.


The ester can be completely hydrolysed rather than only partially if the
reaction is reversible.

Note: If you are interested in the mechanism for the acid


catalysed hydrolyis, follow this link.
Use the BACK button on your browser if you wish to return to
this page.

Where would you like to go now?


To the catalysis menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/catalysis/acidcat.html (5 of 5)30/12/2004 11:19:05

The manufacture of ethanol from ethene

THE MANUFACTURE OF ETHANOL

This page describes the manufacture of ethanol by the direct hydration of


ethene, and then goes on to explain the reasons for the conditions used
in the process. It looks at the effect of proportions, temperature, pressure
and catalyst on the composition of the equilibrium mixture and the rate of
the reaction.

Important: If you aren't sure about using Le Chatelier's


Principle or about the effect of changing conditions on rates
of reaction you should explore these links before you go on.
When you are reading this page, if you find that you aren't
understanding the effect of changing one of the conditions on
the position of equilibrium or on the rate of the reaction, come
back and follow up these links.

A brief summary of the manufacture of ethanol


Ethanol is manufactured by reacting ethene with steam. The reaction is
reversible, and the formation of the ethanol is exothermic.

Only 5% of the ethene is converted into ethanol at each pass through the
reactor. By removing the ethanol from the equilibrium mixture and
recycling the ethene, it is possible to achieve an overall 95% conversion.
A flow scheme for the reaction looks like this:

http://www.chemguide.co.uk/physical/equilibria/ethanol.html (1 of 6)30/12/2004 11:19:20

The manufacture of ethanol from ethene

Note: This is a bit of a simplification! When the gases from


the reactor are cooled, then excess steam will condense as
well as the ethanol. The ethanol will have to be separated
from the water by fractional distillation.
All the sources I have looked at gloss over this, so I don't
have any details. I assume it is a normal fractional distillation
of an ethanol-water mixture.

Explaining the conditions


The proportions of ethene and steam
The equation shows that the ethene and steam react 1 : 1. In order to get
this ratio, you would have to use equal volumes of the two gases.

Because water is cheap, it would seem sensible to use an excess of

http://www.chemguide.co.uk/physical/equilibria/ethanol.html (2 of 6)30/12/2004 11:19:20

The manufacture of ethanol from ethene

steam in order to move the position of equilibrium to the right according


to Le Chatelier's Principle. In practice, an excess of ethene is used.
This is very surprising at first sight. Even if the reaction was one-way,
you couldn't possibly convert all the ethene into ethanol. There isn't
enough steam to react with it.
The reason for this oddity lies with the nature of the catalyst. The catalyst
is phosphoric(V) acid coated onto a solid silicon dioxide support. If you
use too much steam, it dilutes the catalyst and can even wash it off the
support, making it useless.

The temperature
Equilibrium considerations
You need to shift the position of the equilibrium as far as possible to the
right in order to produce the maximum possible amount of ethanol in the
equilibrium mixture.
The forward reaction (the production of ethanol) is exothermic.

According to Le Chatelier's Principle, this will be favoured if you lower the


temperature. The system will respond by moving the position of
equilibrium to counteract this - in other words by producing more heat.
In order to get as much ethanol as possible in the equilibrium mixture,
you need as low a temperature as possible. However, 300C isn't
particularly low.
Rate considerations
The lower the temperature you use, the slower the reaction becomes. A
manufacturer is trying to produce as much ethanol as possible per day. It
makes no sense to try to achieve an equilibrium mixture which contains a
very high proportion of ethanol if it takes several years for the reaction to
reach that equilibrium.

http://www.chemguide.co.uk/physical/equilibria/ethanol.html (3 of 6)30/12/2004 11:19:20

The manufacture of ethanol from ethene

You need the gases to reach equilibrium within the very short time that
they will be in contact with the catalyst in the reactor.
The compromise
300C is a compromise temperature producing an acceptable proportion
of ethanol in the equilibrium mixture, but in a very short time. Under
these conditions, about 5% of the ethene reacts to give ethanol at each
pass over the catalyst.

The pressure
Equilibrium considerations

Notice that there are 2 molecules on the left-hand side of the equation,
but only 1 on the right.
According to Le Chatelier's Principle, if you increase the pressure the
system will respond by favouring the reaction which produces fewer
molecules. That will cause the pressure to fall again.
In order to get as much ethanol as possible in the equilibrium mixture,
you need as high a pressure as possible. High pressures also increase
the rate of the reaction. However, the pressure used isn't all that high.
Problems with high pressures
There are two quite separate problems in this case:

High pressures are expensive. It costs more to build the original


plant because you need extremely strong pipes and containment
vessels. It also needs a lot of energy to produce the high
pressures. That can make the ethanol uneconomic to produce.
At high pressures, the ethene polyerises to make poly(ethene).
Apart from wasting ethene, this could also clog up the plant.

http://www.chemguide.co.uk/physical/equilibria/ethanol.html (4 of 6)30/12/2004 11:19:20

The manufacture of ethanol from ethene

Note: If you are interested in the mechanism for the


polymerisation of ethene, you might like to follow this link although it isn't relevant to the current topic.
Use the BACK button on your browser to return to this page.

The catalyst
Equilibrium considerations
The catalyst has no effect whatsoever on the position of the equilibrium.
Adding a catalyst doesn't produce any greater percentage of ethanol in
the equilibrium mixture. Its only function is to speed up the reaction.
Rate considerations
In the absence of a catalyst the reaction is so slow that virtually no
reaction happens in any sensible time. The catalyst ensures that the
reaction is fast enough for a dynamic equilibrium to be set up within the
very short time that the gases are actually in the reactor.

Note: If you are interested in the mechanism for the


hydration of ethene and the role of the catalyst in it, you will
find it in a section on catalysis by following this link.
Use the BACK button on your browser if you want to return to
this page.

http://www.chemguide.co.uk/physical/equilibria/ethanol.html (5 of 6)30/12/2004 11:19:20

The manufacture of ethanol from ethene

Where would you like to go now?


To the equilibrium menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/equilibria/ethanol.html (6 of 6)30/12/2004 11:19:20

free radical addition in the polymerisation of ethene

THE POLYMERISATION OF ETHENE

A Free Radical Addition Reaction


This page gives you the facts and a simple, uncluttered mechanism for
the polymerisation of ethene by a free radical addition reaction. If you
want the mechanism explained to you in detail, there is a link at the
bottom of the page.
The facts
An addition reaction is one in which two or more molecules join together
to give a single product. During the polymerisation of ethene, thousands
of ethene molecules join together to make poly(ethene) - commonly
called polythene.

The number of molecules joining up is very variable, but is in the region


of 2000 to 20000.
Conditions
Temperature: about 200C
Pressure:
about 2000 atmospheres
a small amount of oxygen as an
Initiator:
impurity
Note: The oxygen is sometimes described as a catalyst for
the reaction. That's not strictly true. A catalyst can be
recovered unchanged at the end of a reaction, but in this
case the oxygen is used up. It gets incorporated into the
polymer molecules - as you will see shortly.

http://www.chemguide.co.uk/mechanisms/freerad/polym.html (1 of 3)30/12/2004 11:19:25

free radical addition in the polymerisation of ethene

The mechanism
The over-all process is known as free radical addition.
Chain initiation
The chain is initiated by free radicals, Ra , produced by reaction
between some of the ethene and the oxygen initiator.
Chain propagation
Each time a free radical hits an ethene molecule a new longer free
radical is formed.

etc
Chain termination
Eventually two free radicals hit each other producing a final molecule.
The process stops here because no new free radicals are formed.

Because chain termination is a random process, poly(ethene) will be


made up of chains of all sorts of different lengths.

Where would you like to go now?


Help! Talk me through this mechanism . . .
To menu of free radical reactions. . .
To menu of other types of mechanism. . .

http://www.chemguide.co.uk/mechanisms/freerad/polym.html (2 of 3)30/12/2004 11:19:25

free radical addition in the polymerisation of ethene

To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/freerad/polym.html (3 of 3)30/12/2004 11:19:25

Explaining free radical addition in the polymerisation of ethene

EXPLAINING THE POLYMERISATION OF


ETHENE

A Free Radical Addition Reaction


This page guides you through the mechanism for the polymerisation of
ethene by a free radical addition reaction.
We are going to talk through this mechanism in a very detailed way so
that you get a feel for what is going on. You couldn't possibly do the
same thing in an exam. At the bottom of the page, you will find the
condensed down version corresponding to the sort of answer you would
produce in an exam.
You will remember that during the polymerisation of ethene, thousands of
ethene molecules join together to make poly(ethene) - commonly called
polythene. The reaction is done at high pressures in the presence of a
trace of oxygen as an initiator.

The function of the oxygen - chain initiation


The oxygen reacts with some of the ethene to give an organic peroxide.
Organic peroxides are very reactive molecules containing oxygenoxygen single bonds which are quite weak and which break easily to give
free radicals.
You can short-cut the process by adding other organic peroxides directly
to the ethene instead of using oxygen if you want to.
You don't need to worry about the exact formulae of the free radicals
which start the reaction off - they vary depending on their source. For
simplicity we give them a general formula: Ra
Chain propagation

http://www.chemguide.co.uk/mechanisms/freerad/polymtt.html (1 of 5)30/12/2004 11:19:30

Explaining free radical addition in the polymerisation of ethene

In an ethene molecule, CH2=CH2, the two


pairs of electrons which make up the double
bond aren't the same. One pair is held
securely on the line between the two carbon
nuclei in a bond called a sigma bond. The
other pair is more loosely held in an orbital
above and below the plane of the molecule known as a pi bond.

Note: It would be helpful - but not essential - if you read


about the structure of ethene before you went on. If the
diagram above is unfamiliar to you, then you certainly ought
to read this background material.
Use the BACK button on your browser to return to this page.

Imagine what happens if a free radical approaches the pi bond in ethene.

Note: Don't worry that we've gone back to a simpler


diagram. As long as you realise that the pair of electrons
shown between the two carbon atoms is in a pi bond - and
therefore vulnerable - that's all that really matters for this
mechanism.

http://www.chemguide.co.uk/mechanisms/freerad/polymtt.html (2 of 5)30/12/2004 11:19:30

Explaining free radical addition in the polymerisation of ethene

The sigma bond between the carbon atoms isn't affected by any of this.
The free radical, Ra , uses one of the electrons in the pi bond to help to
form a new bond between itself and the left hand carbon atom. The other
electron returns to the right hand carbon.
You can show this using "curly arrow" notation if you want to:

Note: If you aren't sure about about curly arrow notation you
can follow this link.
Use the BACK button on your browser to return to this page.

This is energetically worth doing because the new bond between the
radical and the carbon is stronger than the pi bond which is broken. You
would get more energy out when the new bond is made than was used to
break the old one. The more energy that is given out, the more stable the
system becomes.
What we've now got is a bigger free radical - lengthened by CH2CH2.
That can react with another ethene molecule in the same way:

So now the radical is even bigger. That can react with another ethene and so on and so on. The polymer chain gets longer and longer.
Chain termination
The chain doesn't, however, grow indefinitely. Sooner or later two free
radicals will collide together.

http://www.chemguide.co.uk/mechanisms/freerad/polymtt.html (3 of 5)30/12/2004 11:19:30

Explaining free radical addition in the polymerisation of ethene

That immediately stops the growth of two chains and produces one of the
final molecules in the poly(ethene). It is important to realise that the poly
(ethene) is going to be a mixture of molecules of different sizes, made in
this sort of random way.
Simplifying all this for exam purposes
The over-all process is known as free radical addition.
Chain initiation
The chain is initiated by free radicals, Ra , produced by reaction
between some of the ethene and the oxygen initiator.
Chain propagation
Each time a free radical hits an ethene molecule a new longer free
radical is formed.

etc
Chain termination
Eventually two free radicals hit each other producing a final molecule.
The process stops here because no new free radicals are formed.

Because chain termination is a random process, poly(ethene) will be


made up of chains of all sorts of different lengths.

Where would you like to go now?

http://www.chemguide.co.uk/mechanisms/freerad/polymtt.html (4 of 5)30/12/2004 11:19:30

Explaining free radical addition in the polymerisation of ethene

To menu of free radical reactions. . .


To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/freerad/polymtt.html (5 of 5)30/12/2004 11:19:30

mechanism for the hydration of ethene

THE MECHANISM FOR THE ACID CATALYSED


HYDRATION OF ETHENE

This page describes the mechanism for the hydration of ethene to make
ethanol using phosphoric(V) acid as the catalyst.

The hydration of ethene to make ethanol


A reminder of the facts
Ethene is mixed with steam and passed over a catalyst consisting of
solid silicon dioxide coated with phosphoric(V) acid. The temperature
used is 300C and the pressure is about 60 to 70 atmospheres.

Note: If you are interested in the reasons for the conditions


used in this reaction, you will find them in the equilibrium
section of this site by following this link.
Use the BACK button on your browser to return to this page.

The mechanism for the hydration of ethene


This assumes that you know about the electrophilic addition reactions of
ethene, and about the use of curly arrows in organic reaction
mechanisms. If you aren't happy about either of these follow the link
below before you go any further.

http://www.chemguide.co.uk/physical/catalysis/hydrate.html (1 of 6)30/12/2004 11:19:39

mechanism for the hydration of ethene

Note: Follow this link for the electrophilic addition reactions


of ethene. You will find a link explaining the use of curly
arrows in mechanisms on that page if you need it.
These pages are in a completely different part of the site. The
quickest way to return here is to use the BACK button on
your browser, or the history file or the Go menu if you get
seriously waylaid!

All the steps in the mechanism below are shown as one-way reactions
because it makes the mechanism look less confusing. It doesn't affect
the argument, but in fact all the steps are reversible.
Step 1
All of the hydrogen atoms in the phosphoric(V) acid are fairly positively
charged because they are attached to a very electronegative oxygen
atom.

Note: If you aren't sure about electronegativity, you might


like to follow this link.
Use the BACK button on your browser to return quickly to this
page.

http://www.chemguide.co.uk/physical/catalysis/hydrate.html (2 of 6)30/12/2004 11:19:39

mechanism for the hydration of ethene

One of these hydrogens is strongly attracted to the carbon-carbon double


bond. The pi part of the bond breaks and the electrons in it move down to
make a new bond with the hydrogen atom. That forces the electrons in
the hydrogen-oxygen bond down entirely onto the oxygen.

Note: It is easy to see why the oxygen carries a negative


charge. It has gained full control over the electron pair in the
original bond - so has acquired an extra electron which
originally belonged to the hydrogen.
The carbon atom has a positive charge because one of the
electrons in the pi bond originally belonged to it. It loses that
electron when the pi bond breaks.

Step 2
The carbocation (carbonium ion) formed reacts with one of the lone pairs
on a water molecule. A carbocation is one which carries a positive
charge on a carbon atom.

http://www.chemguide.co.uk/physical/catalysis/hydrate.html (3 of 6)30/12/2004 11:19:39

mechanism for the hydration of ethene

Note: The easiest way of remembering that the oxygen has


to carry a positive charge is that you are reacting a positive
ion with a neutral molecule. That means that there must be a
positive charge on the product somewhere.
The only way that an oxygen atom can be joined to three
things at the same time is if it carries a positive charge. A
positive charge gives the oxygen the same electronic
structure as a nitrogen atom which can, of course, form three
bonds.

Step 3
Finally, one of the hydrogens on the oxygen is removed by reaction with
the dihydrogenphosphate(V) ion, H2PO4-, formed in the first step.

The phosphoric(V) acid catalyst has been regenerated.


The reversibility of this reaction
As mentioned before, this reaction is entirely reversible - and so each
step is reversible. In the reverse direction, this is a dehydration of ethanol
to make ethene. If you are interested, you will find the mechanism written
in the reverse direction on another page.

http://www.chemguide.co.uk/physical/catalysis/hydrate.html (4 of 6)30/12/2004 11:19:39

mechanism for the hydration of ethene

Note: This link will lead you to the reverse of this hydration
process, the dehydration of ethanol.
The catalyst used on that page is concentrated sulphuric acid
rather than phosphoric(V) acid, because that is the acid often
used in the lab. You could write either of these mechanisms
(in either direction) perfectly well using the structure of either
of these acids.
Use the BACK button on your browser if you want to return to
this page afterwards.

Where would you like to go now?


To the catalysis menu . . .
To the Physical Chemistry menu . . .
To the organic mechanisms menu . . .
To Main Menu . . .

http://www.chemguide.co.uk/physical/catalysis/hydrate.html (5 of 6)30/12/2004 11:19:39

mechanism for the hydration of ethene

You might also be interested in:


properties and reactions of alkenes . . .
Covers all the physical and chemical properties of alkenes required
by UK A level syllabuses.
properties and reactions of alcohols . . .
A similar survey of alcohol chemistry.

Jim Clark 2002 (modified 2004)

http://www.chemguide.co.uk/physical/catalysis/hydrate.html (6 of 6)30/12/2004 11:19:39

Alcohols Menu

Understanding Chemistry

ALCOHOLS MENU

Background . . .
An introduction to the alcohols and their physical properties.
The manufacture of alcohols . . .
The manufacture of alcohols from alkenes, and the production of
ethanol by fermentation.
Dehydrating alcohols to make alkenes . . .
Various lab methods of converting alcohols into alkenes.
The reaction between alcohols and sodium . . .
Details of the reaction, and some simple properties of the alkoxide
ions formed.
Replacing the OH group by halogen atoms . . .
The reactions between alcohols and phosphorus halides (such as
phosphorus(V) chloride) or sulphur dichloride oxide (thionyl
chloride), and mixtures such as sodium bromide and concentrated
sulphuric acid. Includes testing for the OH group using PCl5.
The oxidation of alcohols . . .
The oxidation of alcohols using acidified potassium dichromate(VI)
solution, and its use in making aldehydes, ketones and carboxylic

http://www.chemguide.co.uk/organicprops/alcoholmenu.html (1 of 2)30/12/2004 11:19:44

Alcohols Menu

acids. Using the reaction as a test for primary, secondary and


tertiary alcohols.
Making esters from alcohols . . .
Mainly the reaction between alcohols and carboxylic acids to
produce esters, together at a brief look at making esters from the
reactions between alcohols and acyl chlorides or acid anhydrides.
The triiodomethane (iodoform) reaction . . .
The use of this reaction to detect the presence of the CH3CH(OH)
group in alcohols.
Uses of alcohols . . .
Concentrates on ethanol and methanol, but with some mention of
propan-2-ol.

Go to menu of other organic compounds . . .


Go to Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/alcoholmenu.html (2 of 2)30/12/2004 11:19:44

an introduction to alcohols

INTRODUCING ALCOHOLS

This page explains what alcohols are, and what the difference is between
primary, secondary and tertiary alcohols. It looks in some detail at their
simple physical properties such as solubility and boiling points. Details of
the chemical reactions of alcohols are described on separate pages.

What are alcohols?


Examples
Alcohols are compounds in which one or more hydrogen atoms in an
alkane have been replaced by an -OH group. For the purposes of UK A
level, we will only look at compounds containing one -OH group.
For example:

Note: If you aren't confident about naming organic


compounds, then you really ought to follow this link before
you go on.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/alcohols/background.html (1 of 10)30/12/2004 11:20:02

an introduction to alcohols

The different kinds of alcohols


Alcohols fall into different classes depending on how the -OH group is
positioned on the chain of carbon atoms. There are some chemical
differences between the various types.
Primary alcohols
In a primary (1) alcohol, the carbon which carries the -OH group is only
attached to one alkyl group.

Note: An alkyl group is a group such as methyl, CH3, or


ethyl, CH3CH2. These are groups containing chains of
carbon atoms which may be branched. Alkyl groups are
given the general symbol R.

Some examples of primary alcohols include:

Notice that it doesn't matter how complicated the attached alkyl group is.
In each case there is only one linkage to an alkyl group from the CH2
group holding the -OH group.
There is an exception to this. Methanol, CH3OH, is counted as a primary
alcohol even though there are no alkyl groups attached to the carbon
with the -OH group on it.
Secondary alcohols
In a secondary (2) alcohol, the carbon with the -OH group attached is
http://www.chemguide.co.uk/organicprops/alcohols/background.html (2 of 10)30/12/2004 11:20:02

an introduction to alcohols

joined directly to two alkyl groups, which may be the same or different.
Examples:

Tertiary alcohols
In a tertiary (3) alcohol, the carbon atom holding the -OH group is
attached directly to three alkyl groups, which may be any combination of
same or different.
Examples:

Physical properties of alcohols


Boiling Points
The chart shows the boiling points of some simple primary alcohols with
up to 4 carbon atoms.
They are:

http://www.chemguide.co.uk/organicprops/alcohols/background.html (3 of 10)30/12/2004 11:20:02

an introduction to alcohols

They are compared with the equivalent alkane (methane to butane) with
the same number of carbon atoms.

Notice that:

The boiling point of an alcohol is always much higher than that of


the alkane with the same number of carbon atoms.
The boiling points of the alcohols increase as the number of
carbon atoms increases.

The patterns in boiling point reflect the patterns in intermolecular


attractions.

Note: If you aren't happy about intermolecular forces


(including van der Waals dispersion forces and hydrogen
bonds) then you really ought to follow this link before you go
on. The next bit won't make much sense to you if you aren't
familiar with the various sorts of intermolecular forces.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/alcohols/background.html (4 of 10)30/12/2004 11:20:02

an introduction to alcohols

Hydrogen bonding
Hydrogen bonding occurs between molecules where you have a
hydrogen atom attached to one of the very electronegative elements fluorine, oxygen or nitrogen.
In the case of alcohols, there are hydrogen bonds set up between the
slightly positive hydrogen atoms and lone pairs on oxygens in other
molecules.

The hydrogen atoms are slightly positive because the bonding electrons
are pulled away from them towards the very electronegative oxygen
atoms.

Note: If you want to be fussy, the diagram is slightly


misleading in that it suggests that all of the lone pairs on the
oxygen atoms are forming hydrogen bonds. In an alcohol that
can't happen. Taking the alcohol as a whole, there are only
half as many slightly positive hydrogen atoms as there are
lone pairs. At any one time, half of the lone pairs in the total
liquid alcohol won't have hydrogen bonds from them because
there aren't enough slightly positive hydrogens to go around.
In the diagram, to show the 3-dimensional arrangement, the
wedge-shaped lines show bonds coming out of the screen or
paper towards you. The dotted bonds (other than the
hydrogen bonds) show bonds going back into the screen or
paper away from you.

http://www.chemguide.co.uk/organicprops/alcohols/background.html (5 of 10)30/12/2004 11:20:02

an introduction to alcohols

In alkanes, the only intermolecular forces are van der Waals dispersion
forces. Hydrogen bonds are much stronger than these and therefore it
takes more energy to separate alcohol molecules than it does to
separate alkane molecules.
That's the main reason that the boiling points are higher.

The effect of van der Waals forces . . .


. . . on the boiling points of the alcohols:
Hydrogen bonding isn't the only intermolecular force in alcohols. There
are also van der Waals dispersion forces and dipole-dipole interactions.
The hydrogen bonding and the dipole-dipole interactions will be much the
same for all the alcohols, but the dispersion forces will increase as the
alcohols get bigger.
These attractions get stronger as the molecules get longer and have
more electrons. That increases the sizes of the temporary dipoles that
are set up.
This is why the boiling points increase as the number of carbon atoms in
the chains increases. It takes more energy to overcome the dispersion
forces, and so the boiling points rise.
. . . on the comparison between alkanes and alcohols:
Even if there wasn't any hydrogen bonding or dipole-dipole interactions,
the boiling point of the alcohol would be higher than the corresponding
alkane with the same number of carbon atoms.
Compare ethane and ethanol:

http://www.chemguide.co.uk/organicprops/alcohols/background.html (6 of 10)30/12/2004 11:20:02

an introduction to alcohols

Ethanol is a longer molecule, and the oxygen brings with it an extra 8


electrons. Both of these will increase the size of the van der Waals
dispersion forces and so the boiling point.
If you were doing a really fair comparison to show the effect of the
hydrogen bonding on boiling point it would be better to compare ethanol
with propane rather than ethane. The length would then be much the
same, and the number of electrons is exactly the same.

Solubility of alcohols in water


The small alcohols are completely soluble in water. Whatever proportions
you mix them in, you will get a single solution.
However, solubility falls as the length of the hydrocarbon chain in the
alcohol increases. Once you get to four carbons and beyond, the fall in
solubility is noticeable, and you may well end up with two layers in your
test tube.
The solubility of the small alcohols in water
Consider ethanol as a typical small alcohol. In both pure water and pure
ethanol the main intermolecular attractions are hydrogen bonds.

In order to mix the two, you would have to break the hydrogen bonds
between the water molecules and the hydrogen bonds between the
ethanol molecules. It needs energy to do both of these things.
However, when the molecules are mixed, new hydrogen bonds are made
http://www.chemguide.co.uk/organicprops/alcohols/background.html (7 of 10)30/12/2004 11:20:02

an introduction to alcohols

between water molecules and ethanol molecules.

The energy released when these new hydrogen bonds are made more or
less compensates for that needed to break the original ones.
In addition, there is an increase in the disorder of the system - an
increase in entropy. That is another factor in deciding whether things
happen or not.

Note: If you haven't come across entropy before, don't worry


about it. I mention it because the energy released when the
new bonds are made isn't quite enough to compensate for
breaking the old ones, meaning that the mixing process is
endothermic. If it weren't for the increase in entropy, the
solution wouldn't be formed.
To really understand this, you need to have studied entropy
and free energy. If you should know about this, but aren't
happy about the calculations involved, you might like to have
a look at chapter 11 of my chemistry calculations book.

http://www.chemguide.co.uk/organicprops/alcohols/background.html (8 of 10)30/12/2004 11:20:02

an introduction to alcohols

The lower solubility of bigger alcohols


Imagine what happens when you have got, say, 5 carbon atoms in each
alcohol molecule.

The hydrocarbon chains are forcing their way between water molecules
and so breaking hydrogen bonds between those water molecules.
The -OH end of the alcohol molecules can form new hydrogen bonds
with water molecules, but the hydrocarbon "tail" doesn't form hydrogen
bonds
That means that quite a lot of the original hydrogen bonds being broken
aren't replaced by new ones.
All you get in place of those original hydrogen bonds are van der Waals
dispersion forces between the water and the hydrocarbon "tails". These
attractions are much weaker. That means that you don't get enough
energy back to compensate for the hydrogen bonds being broken. Even
allowing for the increase in disorder, the process becomes less feasible.
As the length of the alcohol increases, this situation just gets worse, and
so the solubility falls.

Where would you like to go now?

http://www.chemguide.co.uk/organicprops/alcohols/background.html (9 of 10)30/12/2004 11:20:02

an introduction to alcohols

To the alcohols menu . . .


To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/alcohols/background.html (10 of 10)30/12/2004 11:20:02

manufacture of alcohols

THE MANUFACTURE OF ALCOHOLS

This page looks at the manufacture of alcohols by the direct hydration of


alkenes, concentrating mainly on the hydration of ethene to make
ethanol. It then compares that method with making ethanol by
fermentation.

Manufacturing alcohols from alkenes


The manufacture of ethanol from ethene
Ethanol is manufactured by reacting ethene with steam. The catalyst
used is solid silicon dioxide coated with phosphoric(V) acid. The reaction
is reversible.

Only 5% of the ethene is converted into ethanol at each pass through the
reactor. By removing the ethanol from the equilibrium mixture and
recycling the ethene, it is possible to achieve an overall 95% conversion.
A flow scheme for the reaction looks like this:

http://www.chemguide.co.uk/organicprops/alcohols/manufacture.html (1 of 7)30/12/2004 11:20:12

manufacture of alcohols

Note: This is a bit of a simplification! When the gases from


the reactor are cooled, then excess steam will condense as
well as the ethanol. The ethanol will have to be separated
from the water by fractional distillation.
All the sources I have looked at gloss over this, so I don't
have any details. I assume it is a normal fractional distillation
of an ethanol-water mixture.
If you are interested in the reasons for the conditions used in
this reaction, you will find them in the equilibrium section of
this site by following this link.
If you are interested in the mechanism for the hydration of
ethene, follow this link to the catalysis section.
Because these pages are (somewhat illogically!) in different
parts of the site, use the BACK button (or HISTORY file or
GO menu) on your browser to return to this page later.

http://www.chemguide.co.uk/organicprops/alcohols/manufacture.html (2 of 7)30/12/2004 11:20:12

manufacture of alcohols

The manufacture of other alcohols from alkenes


Some - but not all - other alcohols can be made by similar reactions. The
catalyst used and the reaction conditions will vary from alcohol to
alcohol. The only set of conditions you are going to need for UK A level
purposes are those given above for manufacturing ethanol.
The reason that there is a problem with some alcohols is well illustrated
with trying to make an alcohol from propene, CH3CH=CH2.
In principle, there are two different alcohols which might be formed:

You might expect to get either propan-1-ol or propan-2-ol depending on


which way around the water adds to the double bond. In practice what
you get is propan-2-ol.
If you add a molecule H-X across a carbon-carbon double bond, the
hydrogen nearly always gets attached to the carbon with the most
hydrogens on it already - in this case the CH2 rather than the CH.

http://www.chemguide.co.uk/organicprops/alcohols/manufacture.html (3 of 7)30/12/2004 11:20:12

manufacture of alcohols

Note: The reason for this is dealt with in detail in the


mechanism section of this site on a page about addition to
unsymmetrical alkenes.
If you choose to follow this link, use the BACK button on your
browser to return to this page.

The effect of this is that there are bound to be some alcohols which it is
impossible to make by reacting alkenes with steam because the addition
would be the wrong way around.

Making ethanol by fermentation


This method only applies to ethanol. You can't make any other alcohol
this way.
The process
The starting material for the process varies widely, but will normally be
some form of starchy plant material such as maize (US: corn), wheat,
barley or potatoes.
Starch is a complex carbohydrate, and other carbohydrates can also be
used - for example, in the lab sucrose (sugar) is normally used to
produce ethanol. Industrially, this wouldn't make sense. It would be silly
to refine sugar if all you were going to use it for was fermentation. There
is no reason why you shouldn't start from the original sugar cane, though.
The first step is to break complex carbohydrates into simpler ones.
For example, if you were starting from starch in grains like wheat or
barley, the grain is heated with hot water to extract the starch and then
warmed with malt. Malt is germinated barley which contains enzymes
which break the starch into a simpler carbohydrate called maltose,
C12H22O11.
http://www.chemguide.co.uk/organicprops/alcohols/manufacture.html (4 of 7)30/12/2004 11:20:12

manufacture of alcohols

Maltose has the same molecular formula as sucrose but contains two
glucose units joined together, whereas sucrose contains one glucose
and one fructose unit.
Yeast is then added and the mixture is kept warm (say 35C) for perhaps
several days until fermentation is complete. Air is kept out of the mixture
to prevent oxidation of the ethanol produced to ethanoic acid (vinegar).
Enzymes in the yeast first convert carbohydrates like maltose or sucrose
into even simpler ones like glucose and fructose, both C6H12O6, and
then convert these in turn into ethanol and carbon dioxide.
You can show these changes as simple chemical equations, but the
biochemistry of the reactions is much, much more complicated than this
suggests.

Yeast is killed by ethanol concentrations in excess of about 15%, and


that limits the purity of the ethanol that can be produced. The ethanol is
separated from the mixture by fractional distillation to give 96% pure
ethanol.
For theoretical reasons, it is impossible to remove the last 4% of water by
fractional distillation.

Note: If you are interested in this, you will have to read


about it in a physical chemistry textbook under "azeotropic
mixtures". Currently I haven't dealt with these on this site,
and am unlikely to do so until at least 2005. I will add a link if
or when I get around to it!

http://www.chemguide.co.uk/organicprops/alcohols/manufacture.html (5 of 7)30/12/2004 11:20:12

manufacture of alcohols

A comparison of fermentation with the direct hydration of ethene


Fermentation

Hydration of ethene

Type of process

A batch process.
Everything is put into a
container and then left
until fermentation is
complete. That batch
is then cleared out and
a new reaction set up.
This is inefficient.

A continuous flow
process. A stream of
reactants is passed
continuously over a
catalyst. This is a
more efficient way of
doing things.

Rate of reaction

Very slow.

Very rapid.

Quality of
product

Produces very impure


ethanol which needs
further processing

Produces much purer


ethanol.

Reaction
conditions

Uses gentle
temperatures and
atmospheric pressure.

Uses high
temperatures and
pressures, needing
lots of energy input.

Use of resources

Uses renewable
resources based on
plant material.

Uses finite resources


based on crude oil.

Where would you like to go now?


To the alcohols menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

http://www.chemguide.co.uk/organicprops/alcohols/manufacture.html (6 of 7)30/12/2004 11:20:12

manufacture of alcohols

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/alcohols/manufacture.html (7 of 7)30/12/2004 11:20:12

dehydration of alcohols

THE DEHYDRATION OF ALCOHOLS

This page (a simple duplicate of a page in the section on alkenes!) looks


at the dehydration of alcohols in the lab to make alkenes - for example,
dehydrating ethanol to make ethene.

Dehydration of alcohols using aluminium oxide as catalyst


The dehydration of ethanol to give ethene
This is a simple way of making gaseous alkenes like ethene. If ethanol
vapour is passed over heated aluminium oxide powder, the ethanol is
essentially cracked to give ethene and water vapour.

To make a few test tubes of ethene, you can use this apparatus:

It wouldn't be too difficult to imagine scaling this up by boiling some


ethanol in a flask and passing the vapour over aluminium oxide heated in
a long tube.

http://www.chemguide.co.uk/organicprops/alcohols/dehydration.html (1 of 4)30/12/2004 11:20:21

dehydration of alcohols

Dehydration of alcohols using an acid catalyst


The acid catalysts normally used are either concentrated sulphuric acid
or concentrated phosphoric(V) acid, H3PO4.
Concentrated sulphuric acid produces messy results. Not only is it an
acid, but it is also a strong oxidising agent. It oxidises some of the
alcohol to carbon dioxide and at the same time is reduced itself to
sulphur dioxide. Both of these gases have to be removed from the alkene.
It also reacts with the alcohol to produce a mass of carbon. There are
other side reactions as well, but these aren't required by any current UK
A level syllabus.

The dehydration of ethanol to give ethene


Ethanol is heated with an excess of concentrated sulphuric acid at a
temperature of 170C. The gases produced are passed through sodium
hydroxide solution to remove the carbon dioxide and sulphur dioxide
produced from side reactions.
The ethene is collected over water.

WARNING! This is potentially an extremely dangerous


preparation because of the close proximity of the very hot
concentrated sulphuric acid and the sodium hydroxide
solution. I knew of one chemistry teacher who put several
students into hospital by getting it wrong! That was many
years ago before safety was taken quite so seriously as it is
now.

The concentrated sulphuric acid is a catalyst. Write it over the arrow


rather than in the equation.

http://www.chemguide.co.uk/organicprops/alcohols/dehydration.html (2 of 4)30/12/2004 11:20:21

dehydration of alcohols

Note: You will find the mechanism for the dehydration of


alcohols in the mechanism section of this site. You will also
find a discussion of how to cope with questions about the
dehydration of more complicated alcohols if you follow a link
at the bottom of that page.
Use the BACK button (or the HISTORY file or GO menu) on
your browser if you want to return to this page.

The dehydration of cyclohexanol to give cyclohexene


This is a preparation commonly used at this level to illustrate the
formation and purification of a liquid product. The fact that the carbon
atoms happen to be joined in a ring makes no difference whatever to the
chemistry of the reaction.
Cyclohexanol is heated with concentrated phosphoric(V) acid and the
liquid cyclohexene distils off and can be collected and purified.
Phosphoric(V) acid tends to be used in place of sulphuric acid because it
is safer and produces a less messy reaction.

Where would you like to go now?


To the alcohols menu . . .

http://www.chemguide.co.uk/organicprops/alcohols/dehydration.html (3 of 4)30/12/2004 11:20:21

dehydration of alcohols

To the menu of other organic compounds . . .


To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/alcohols/dehydration.html (4 of 4)30/12/2004 11:20:21

alcohols and sodium

REACTING ALCOHOLS WITH SODIUM

This page describes the reaction between alcohols and metallic sodium,
and takes a very brief look at the properties of the alkoxide which is
formed. We will look at the reaction between sodium and ethanol as being
typical, but you could substitute any other alcohol you wanted to - the
reaction would be the same.

The reaction between sodium and ethanol


Details of the reaction
If a small piece of sodium is dropped into some ethanol, it reacts steadily
to give off bubbles of hydrogen gas and leaves a colourless solution of
sodium ethoxide, CH3CH2ONa. Sodium ethoxide is known as an
alkoxide.
If the solution is evaporated carefully to dryness, the sodium ethoxide is
left as a white solid.

Although at first sight you might think this was something new and
complicated, in fact it is exactly the same (apart from being a more gentle
reaction) as the reaction between sodium and water - something you
have probably known about for years.
Compare the two:

We normally, of course, write the sodium hydroxide formed as NaOH


rather than HONa - but that's the only difference.

http://www.chemguide.co.uk/organicprops/alcohols/sodium.html (1 of 5)30/12/2004 11:20:29

alcohols and sodium

Sodium ethoxide is just like sodium hydroxide, except that the hydrogen
has been replaced by an ethyl group. Sodium hydroxide contains OHions; sodium ethoxide contains CH3CH2O- ions.

Note: The reason that the ethoxide formula is written with the
oxygen on the right unlike the hydroxide ion is simply a matter
of clarity. If you write it the other way around, it doesn't
immediately look as if it comes from ethanol. You will find the
same thing happens when you write formulae for organic salts
like sodium ethanoate, for example.

Using the reaction


There are two simple uses for this reaction:
To dispose of small amounts of sodium safely
If you spill some sodium on the bench, or have a small amount left over
from a reaction, you can't just chuck it in the sink. It tends to react
explosively with the water - and comes flying back out at you again!
It reacts much more gently with ethanol. Ethanol is therefore used to
dissolve small quantities of waste sodium. The solution formed can be
washed away without problems (provided you remember that sodium
ethoxide is strongly alkaline - see below).
To test for the -OH group in alcohols
Because of the dangers involved in handling sodium, this is not the best
test for an alcohol at this level. Because sodium reacts violently with acids
to produce a salt and hydrogen, you would first have to be sure that the
liquid you were testing was neutral.
You would also have to be confident that there was no trace of water
present because sodium reacts with the -OH group in water even better
than with the one in an alcohol.
With those provisos, if you add a tiny piece of sodium to a neutral liquid
free of water and get bubbles of hydrogen produced, then the liquid is an
http://www.chemguide.co.uk/organicprops/alcohols/sodium.html (2 of 5)30/12/2004 11:20:29

alcohols and sodium

alcohol.

Some simple reactions of alkoxide ions


This is going beyond the demands of UK A level, but you might come
across the first example as a part of a bit of practical work. The second
example is to reinforce the similarity between sodium ethoxide and
sodium hydroxide.
Once again we will take the ethoxide ions in sodium ethoxide as typical.
Essentially, ethoxide (and other alkoxide) ions behave like hydroxide ions.
Ethoxide ions are strongly basic
If you add water to sodium ethoxide, it dissolves to give a colourless
solution with a high pH - typically pH 14. The solution is strongly alkaline.
The reason is that the ethoxide ions remove hydrogen ions from water
molecules to produce hydroxide ions. It is these which produce the high
pH.

Ethoxide ions are good nucleophiles


A nucleophile is something which carries a negative or partial negative
charge which it uses to attack positive centres in other molecules or ions.
Hydroxide ions are good nucleophiles, and you may have come across
the reaction between a halogenoalkane (also called a haloalkane or alkyl
halide) and sodium hydroxide solution. The hydroxide ions replace the
halogen atom.

In this case, an alcohol is formed.

http://www.chemguide.co.uk/organicprops/alcohols/sodium.html (3 of 5)30/12/2004 11:20:29

alcohols and sodium

Note: You will find the mechanism for this reaction in the
mechanism section of this site.
Use the BACK button on your browser to return to this page.

The ethoxide ion behaves in exactly the same way. If you knew the
mechanism for the hydroxide ion reaction you could work out exactly what
happens in the reaction between a halogenoalkane and ethoxide ion.
Compare this equation with the last one.

The only difference is that where there was a hydrogen atom at the righthand end of the product molecule, you now have an alkyl group.
Two alkyl (or other hydrocarbon) groups bridged by an oxygen atom is
called an ether. This particular one is 1-ethoxypropane or ethyl propyl
ether. This reaction is a good way of making ethers in the lab.

Note: There are all sorts of ways of naming ethers. For UK A


level purposes, the problem doesn't arise - you almost
certainly won't have to name them.
If you have looked at the chemistry of halogenoalkanes, you
may be aware that there is a competition between substitution
and elimination when they react with hydroxide ions. Exactly
the same competition occurs in their reactions with ethoxide
ions. Just as the hydroxide ion can act as either a base or a
nucleophile, exactly the same is true of alkoxide ions like the
ethoxide ion.
You could read about the reactions of halogenoalkanes with
hydroxide ions and work out for yourself what is going to

http://www.chemguide.co.uk/organicprops/alcohols/sodium.html (4 of 5)30/12/2004 11:20:29

alcohols and sodium

happen in the possible elimination reaction if you used sodium


ethoxide rather than sodium hydroxide. Because this is getting
well beyond UK A level, I haven't given any detail for this
anywhere on the site. The whole point about understanding
chemistry (and especially mechanisms!) is that you can work
things out for yourself when you need to!

Where would you like to go now?


To the alcohols menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/alcohols/sodium.html (5 of 5)30/12/2004 11:20:29

halogenoalkanes (haloalkanes) and hydroxide ions

REACTIONS BETWEEN HALOGENOALKANES


AND HYDROXIDE IONS

This page looks at the reactions between halogenoalkanes (haloalkanes


or alkyl halides) and hydroxide ions from sodium or potassium hydroxide
solution. It covers both substitution and elimination reactions.
To a large extent, this page simply brings together information from a
number of other pages on the site. If you want information about the
mechanisms for these reactions you will find them elsewhere. This page
has links to all the other pages that you will need.

Substitution or elimination?
There are two different sorts of reaction that you can get depending on
the conditions used and the type of halogenoalkane. Primary, secondary
and tertiary halogenoalkanes behave differently in this respect.

Note: If you aren't sure what primary, secondary and tertiary


halogenoalkanes are, you should read the beginning of the
introduction to halogenoalkanes by following this link before
you go on.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/haloalkanes/hydroxide.html (1 of 5)30/12/2004 11:20:34

halogenoalkanes (haloalkanes) and hydroxide ions

Substitution reactions
In a substitution reaction, the halogen atom is replaced by an -OH group
to give an alcohol.
For example:

Or, as an ionic equation:

In the example, 2-bromopropane is converted into propan-2-ol.


The halogenoalkane is heated under reflux with a solution of sodium or
potassium hydroxide. Heating under reflux means heating with a
condenser placed vertically in the flask to prevent loss of volatile
substances from the mixture.
The solvent is usually a 50/50 mixture of ethanol and water, because
everything will dissolve in that. The halogenoalkane is insoluble in water.
If you used water alone as the solvent, the halogenoalkane and the
sodium hydroxide solution wouldn't mix and the reaction could only
happen where the two layers met.

Note: If you want the mechanisms for substitution reactions


follow this link. Use the BACK button (or the GO menu or
HISTORY file) on your browser to return to this page later if
you want to.

http://www.chemguide.co.uk/organicprops/haloalkanes/hydroxide.html (2 of 5)30/12/2004 11:20:34

halogenoalkanes (haloalkanes) and hydroxide ions

Elimination reactions
Halogenoalkanes also undergo elimination reactions in the presence of
sodium or potassium hydroxide.

The 2-bromopropane has reacted to give an alkene - propene.


Notice that a hydrogen atom has been removed from one of the end
carbon atoms together with the bromine from the centre one. In all simple
elimination reactions the things being removed are on adjacent carbon
atoms, and a double bond is set up between those carbons.
The halogenoalkane is heated under reflux with a concentrated solution
of sodium or potassium hydroxide in ethanol. Propene is formed and,
because this is a gas, it passes through the condenser and can be
collected.

Note: If you want to read about the mechanisms for


elimination reactions follow this link. Use the BACK button (or
the GO menu or HISTORY file) on your browser to return to
this page later if you want to.

What decides whether you get substitution or elimination?


The reagents you are using are the same for both substitution or
elimination - the halogenoalkane and either sodium or potassium
hydroxide solution. In all cases, you will get a mixture of both reactions
happening - some substitution and some elimination. What you get most
of depends on a number of factors.
The type of halogenoalkane

http://www.chemguide.co.uk/organicprops/haloalkanes/hydroxide.html (3 of 5)30/12/2004 11:20:34

halogenoalkanes (haloalkanes) and hydroxide ions

This is the most important factor.


type of halogenoalkane

substitution or elimination?

primary

mainly substitution

secondary

both substitution and elimination

tertiary

mainly elimination

For example, whatever you do with tertiary halogenoalkanes, you will


tend to get mainly the elimination reaction, whereas with primary ones
you will tend to get mainly substitution. However, you can influence
things to some extent by changing the conditions.
The solvent
The proportion of water to ethanol in the solvent matters.

Water encourages substitution.

Ethanol encourages elimination.

The temperature
Higher temperatures encourage elimination.
Concentration of the sodium or potassium hydroxide solution
Higher concentrations favour elimination.
In summary
For a given halogenoalkane, to favour elimination rather than
substitution, use:

higher temperatures

a concentrated solution of sodium or potassium hydroxide

pure ethanol as the solvent

http://www.chemguide.co.uk/organicprops/haloalkanes/hydroxide.html (4 of 5)30/12/2004 11:20:34

halogenoalkanes (haloalkanes) and hydroxide ions

To favour substitution rather than elimination, use:

lower temperatures

more dilute solutions of sodium or potassium hydroxide

more water in the solvent mixture

Note: The explanations for these effects are well beyond the
demands of UK A level syllabuses. Some things you just
have to know!

Where would you like to go now?


To the halogenoalkanes menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/haloalkanes/hydroxide.html (5 of 5)30/12/2004 11:20:34

an introduction to halogenoalkanes (haloalkanes)

INTRODUCING HALOGENOALKANES
(haloalkanes or alkyl halides)

Halogenoalkanes are also known as haloalkanes or alkyl halides. This


page explains what they are and discusses their physical properties. It also
takes an initial look at their chemical reactivity. Details of the chemical
reactions of halogenoalkanes are described on separate pages.

What are halogenoalkanes?


Examples
Halogenoalkanes are compounds in which one or more hydrogen atoms in
an alkane have been replaced by halogen atoms (fluorine, chlorine,
bromine or iodine). For the purposes of UK A level, we will only look at
compounds containing one halogen atom.
For example:

Note: If you aren't confident about naming organic


compounds, then you really ought to follow this link before you
go on.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/haloalkanes/background.html (1 of 10)30/12/2004 11:20:42

an introduction to halogenoalkanes (haloalkanes)

The different kinds of halogenoalkanes


Halogenoalkanes fall into different classes depending on how the halogen
atom is positioned on the chain of carbon atoms. There are some chemical
differences between the various types.
Primary halogenoalkanes
In a primary (1) halogenoalkane, the carbon which carries the halogen
atom is only attached to one other alkyl group.

Note: An alkyl group is a group such as methyl, CH3, or ethyl,


CH3CH2. These are groups containing chains of carbon atoms
which may be branched. Alkyl groups are given the general
symbol R.

Some examples of primary halogenoalkanes include:

Notice that it doesn't matter how complicated the attached alkyl group is. In
each case there is only one linkage to an alkyl group from the CH2 group
holding the halogen.
There is an exception to this. CH3Br and the other methyl halides are often
counted as primary halogenoalkanes even though there are no alkyl
groups attached to the carbon with the halogen on it.
Secondary halogenoalkanes
In a secondary (2) halogenoalkane, the carbon with the halogen attached
is joined directly to two other alkyl groups, which may be the same or
different.
Examples:

http://www.chemguide.co.uk/organicprops/haloalkanes/background.html (2 of 10)30/12/2004 11:20:42

an introduction to halogenoalkanes (haloalkanes)

Tertiary halogenoalkanes
In a tertiary (3) halogenoalkane, the carbon atom holding the halogen is
attached directly to three alkyl groups, which may be any combination of
same or different.
Examples:

Physical properties of halogenoalkanes


Boiling Points
The chart shows the boiling points of some simple halogenoalkanes.

Notice that three of these have boiling points below room temperature
http://www.chemguide.co.uk/organicprops/haloalkanes/background.html (3 of 10)30/12/2004 11:20:42

an introduction to halogenoalkanes (haloalkanes)

(taken as being about 20C). These will be gases at room temperature. All
the others you are likely to come across are liquids.
Remember:

the only methyl halide which is a liquid is iodomethane;

chloroethane is a gas.

The patterns in boiling point reflect the patterns in intermolecular


attractions.

Note: If you aren't happy about intermolecular forces


(particularly van der Waals dispersion forces and dipole-dipole
interactions) then you really ought to follow this link before you
go on.
Use the BACK button on your browser to return to this page.

van der Waals dispersion forces


These attractions get stronger as the molecules get longer and have more
electrons. That increases the sizes of the temporary dipoles that are set up.
This is why the boiling points increase as the number of carbon atoms in
the chains increases. Look at the chart for a particular type of halide (a
chloride, for example). Dispersion forces get stronger as you go from 1 to 2
to 3 carbons in the chain. It takes more energy to overcome them, and so
the boiling points rise.
The increase in boiling point as you go from a chloride to a bromide to an
iodide (for a given number of carbon atoms) is also because of the
increase in number of electrons leading to larger dispersion forces. There
are lots more electrons in, for example, iodomethane than there are in
chloromethane - count them!

http://www.chemguide.co.uk/organicprops/haloalkanes/background.html (4 of 10)30/12/2004 11:20:42

an introduction to halogenoalkanes (haloalkanes)

van der Waals dipole-dipole attractions


The carbon-halogen bonds (apart from the carbon-iodine bond) are polar,
because the electron pair is pulled closer to the halogen atom than the
carbon. This is because (apart from iodine) the halogens are more
electronegative than carbon.
The electronegativity values are:
C 2.5

4.0

Cl 3.0
Br 2.8
I

2.5

Note: If you don't understand about electronegativity and polar


bonds then you should follow this link before you go on.
Use the BACK button on your browser to return to this page.

This means that in addition to the dispersion forces there will be forces due
to the attractions between the permanent dipoles (except in the iodide
case).
The size of those dipole-dipole attractions will fall as the bonds get less
polar (as you go from chloride to bromide to iodide, for example).
Nevertheless, the boiling points rise! This shows that the effect of the
permanent dipole-dipole attractions is much less important than that of the
temporary dipoles which cause the dispersion forces.
The large increase in number of electrons by the time you get to the iodide
completely outweighs the loss of any permanent dipoles in the molecules.

Boiling points of some isomers


http://www.chemguide.co.uk/organicprops/haloalkanes/background.html (5 of 10)30/12/2004 11:20:42

an introduction to halogenoalkanes (haloalkanes)

The examples show that the boiling points fall as the isomers go from a
primary to a secondary to a tertiary halogenoalkane. This is a simple result
of the fall in the effectiveness of the dispersion forces.
The temporary dipoles are greatest for the longest molecule. The
attractions are also stronger if the molecules can lie closely together. The
tertiary halogenoalkane is very short and fat, and won't have much close
contact with its neighbours.

Solubility of halogenoalkanes
Solubility in water
The halogenoalkanes are at best only very slightly soluble in water.
In order for a halogenoalkane to dissolve in water you have to break
attractions between the halogenoalkane molecules (van der Waals
dispersion and dipole-dipole interactions) and break the hydrogen bonds
between water molecules. Both of these cost energy.

Note: If you aren't sure about hydrogen bonds, then you


should follow this link before you go on.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/haloalkanes/background.html (6 of 10)30/12/2004 11:20:42

an introduction to halogenoalkanes (haloalkanes)

Energy is released when new attractions are set up between the


halogenoalkane and the water molecules. These will only be dispersion
forces and dipole-dipole interactions. These aren't as strong as the original
hydrogen bonds in the water, and so not as much energy is released as
was used to separate the water molecules.
The energetics of the change are sufficiently "unprofitable" that very little
dissolves.

Note: To be accurate, you also have to consider entropy


changes when things dissolve. If you don't yet know about
entropy, don't worry about it!

Solubility in organic solvents


Halogenoalkanes tend to dissolve in organic solvents because the new
intermolecular attractions have much the same strength as the ones being
broken in the separate halogenoalkane and solvent.

Chemical reactivity of halogenoalkanes


The importance of bond strengths
The pattern in strengths of the four carbon-halogen bonds are:

http://www.chemguide.co.uk/organicprops/haloalkanes/background.html (7 of 10)30/12/2004 11:20:42

an introduction to halogenoalkanes (haloalkanes)

Notice that bond strength falls as you go from C-F to C-I, and notice how
much stronger the carbon-fluorine bond is than the rest.

Note: You will find almost as many different values for bond
strengths (or bond enthalpies or bond energies) as there are
different sources! Don't worry about this - the pattern is always
the same. This is why you have got a chart here rather than
actual numbers.

In order for anything to react with the halogenoalkanes, the carbonhalogen bond has got to be broken. Because that gets easier as you go
from fluoride to chloride to bromide to iodide, the compounds get more
reactive in that order.
Iodoalkanes are the most reactive and fluoroalkanes are the least. In fact,
fluoroalkanes are so unreactive that we shall pretty well ignore them
completely from now on in this section!

The influence of bond polarity


Of the four halogens, fluorine is the most electronegative and iodine the
least. That means that the electron pair in the carbon-fluorine bond will be
dragged most towards the halogen end.
Looking at the methyl halides as simple examples:

The electronegativities of carbon and iodine are equal and so there will be
no separation of charge on the bond.

http://www.chemguide.co.uk/organicprops/haloalkanes/background.html (8 of 10)30/12/2004 11:20:42

an introduction to halogenoalkanes (haloalkanes)

Note: In the diagram, " " (read as "delta") means "slightly" - so


+ means "slightly positive". The larger the symbols, the more
positive or negative the atoms are intended to be.
If this isn't reasonably familiar to you, then you ought to have
read the page about electronegativity and polar bonds
mentioned above!
Use the BACK button on your browser to return to this page.

One of the important set of reactions of halogenoalkanes involves


replacing the halogen by something else - substitution reactions. These
reactions involve either:

the carbon-halogen bond breaking to give positive and negative


ions. The ion with the positively charged carbon atom then reacts
with something either fully or slightly negatively charged.
something either fully or negatively charged attracted to the slightly
positive carbon atom and pushing off the halogen atom.

You might have thought that either of these would be more effective in the
case of the carbon-fluorine bond with the quite large amounts of positive
and negative charge already present. But that's not so - quite the opposite
is true!
The thing that governs the reactivity is the strength of the bonds which
have to be broken. If is difficult to break a carbon-fluorine bond, but easy to
break a carbon-iodine one.

http://www.chemguide.co.uk/organicprops/haloalkanes/background.html (9 of 10)30/12/2004 11:20:42

an introduction to halogenoalkanes (haloalkanes)

Note: You will find all of this discussed in detail in the


mechanism section of this site under nucleophilic substitution
reactions. This link will take you to the most relevant page, but
you may also want to explore the rest of the nucleophilic
substitution menu.
Use the BACK button (or HISTORY file or GO menu) on your
browser if you want to return to this page later.

Where would you like to go now?


To the halogenoalkanes menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/haloalkanes/background.html (10 of 10)30/12/2004 11:20:42

Halogenoalkanes (haloalkanes) Menu

Understanding Chemistry

HALOGENOALKANES MENU
(Also known as haloalkanes or alkyl halides)

Background . . .
An introduction to the halogenoalkanes, their reactivity and their
physical properties.
Making halogenoalkanes . . .
Making halogenoalkanes in the lab from alcohols.
The reaction with hydroxide ions . . .
The reaction with hydroxide ions from, for example, sodium
hydroxide solution leading to substitution and elimination reactions.
The reaction with cyanide ions . . .
The reaction with cyanide ions from, for example, a solution of
potassium cyanide in alcohol.
The reaction with ammonia . . .
The reaction with ammonia to produce various kinds of amines.
Reactions involving silver nitrate solution . . .
Includes testing for halogenoalkanes, and the use of silver nitrate
solution as a guide to the reactivity of the various halogenoalkanes.
http://www.chemguide.co.uk/organicprops/haloalkanemenu.html (1 of 2)30/12/2004 11:20:43

Halogenoalkanes (haloalkanes) Menu

Grignard reagents . . .
An introduction to Grignard reagents, including making them from
halogenoalkanes and some of their simple reactions.
Uses of halogenoalkanes . . .
Including the problems associated with CFCs.

Go to menu of other organic compounds . . .


Go to Main Menu . . .

You might also be interested in:


aryl halides (halogenoarenes) . . .
This includes the reasons that halogen atoms attached to benzene rings are
much less easily replaced than when they are attached to carbon chains.

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/haloalkanemenu.html (2 of 2)30/12/2004 11:20:43

making halogenoalkanes (haloalkanes)

MAKING HALOGENOALKANES (haloalkanes or


alkyl halides)

This page looks at ways of making halogenoalkanes in the lab starting


from alcohols. It covers more than enough detail for UK A level purposes,
but isn't intended to be all-inclusive!

Summarising the methods of preparation


Halogenoalkanes can be made from the reaction between alkenes and
hydrogen halides, but they are more commonly made by replacing the -OH
group in an alcohol by a halogen atom. That's the method we'll
concentrate on in this page.

Making halogenoalkanes from alcohols using hydrogen halides


The general reaction looks like this:

Making chloroalkanes
It is possible to make tertiary chloroalkanes successfully from the
corresponding alcohol and concentrated hydrochloric acid, but to make
primary or secondary ones you really need to use a different method - the
reaction rates are too slow.
A tertiary chloroalkane can be made by shaking the corresponding alcohol
with concentrated hydrochloric acid at room temperature.

http://www.chemguide.co.uk/organicprops/haloalkanes/making.html (1 of 8)30/12/2004 11:20:54

making halogenoalkanes (haloalkanes)

Note: If you don't know what primary, secondary and tertiary


halogenoalkanes are, you should read the introduction to
halogenoalkanes before you go on.
Use the BACK button on your browser to return to this page.

Making bromoalkanes
Rather than using hydrobromic acid, you usually treat the alcohol with a
mixture of sodium or potassium bromide and concentrated sulphuric acid.
This produces hydrogen bromide which reacts with the alcohol. The
mixture is warmed to distil off the bromoalkane. You will find practical
details of a reaction of this kind further down the page.

Making iodoalkanes
In this case the alcohol is reacted with a mixture of sodium or potassium
iodide and concentrated phosphoric(V) acid, H3PO4, and the iodoalkane is
distilled off. The mixture of the iodide and phosphoric(V) acid produces
hydrogen iodide which reacts with the alcohol.

Phosphoric(V) acid is used instead of concentrated sulphuric acid because


sulphuric acid oxidises iodide ions to iodine and produces hardly any
hydrogen iodide. A similar thing happens to some extent with bromide ions
in the preparation of bromoalkanes, but not enough to get in the way of the
main reaction.

http://www.chemguide.co.uk/organicprops/haloalkanes/making.html (2 of 8)30/12/2004 11:20:54

making halogenoalkanes (haloalkanes)

Note: If you are interested in the reactions between halide


ions and concentrated sulphuric acid you could follow this link.
In the present context, all you would need to do is read the
beginning of that page.
If you choose to follow this link, use the BACK button on your
browser to return to this page.

Making halogenoalkanes from alcohols using phosphorus halides


Making chloroalkanes
Chloroalkanes can be made by reacting an alcohol with liquid phosphorus
(III) chloride, PCl3.

Note: The reason for choosing a slightly bigger alcohol as an


example rather than using ethanol is that chloroethane (which
would be formed with ethanol) is a gas, and the preparation
would have to be handled differently. 1-chloropropane (formed
in this equation) is a liquid.

http://www.chemguide.co.uk/organicprops/haloalkanes/making.html (3 of 8)30/12/2004 11:20:54

making halogenoalkanes (haloalkanes)

They can also be made by adding solid phosphorus(V) chloride, PCl5 to an


alcohol.
This reaction is violent at room temperature, producing clouds of hydrogen
chloride gas. It isn't a good choice as a way of making halogenoalkanes,
although it is used as a test for -OH groups in organic chemistry.

There are also side reactions involving the POCl3 reacting with the alcohol.
Making bromoalkanes and iodoalkanes
These are both made in the same general way. Instead of using
phosphorus(III) bromide or iodide, the alcohol is heated under reflux with a
mixture of red phosphorus and either bromine or iodine.
The phosphorus first reacts with the bromine or iodine to give the
phosphorus(III) halide.

These then react with the alcohol to give the corresponding


halogenoalkane which can be distilled off.

Making bromoethane in the lab


This is a simple example of an organic preparation, and is one of the most
commonly used in A level chemistry courses.

http://www.chemguide.co.uk/organicprops/haloalkanes/making.html (4 of 8)30/12/2004 11:20:54

making halogenoalkanes (haloalkanes)

Warning! The method described below is NOT complete. It


deliberately avoids mentioning quantities and doesn't cover
any of the safety aspects of the preparation. You can find full
details in most organic practical books.

Making impure bromoethane


Concentrated sulphuric acid is added slowly with lots of shaking and
cooling to some ethanol in a flask, and then solid potassium bromide is
added. The flask is then connected to a condenser so that the
bromoethane formed can be distilled off.

Bromoethane has a low boiling point but is denser than water and almost
insoluble in it. To prevent it from evaporating, it is often collected under
water in a flask surrounded by ice. Sometimes it is simply collected in a
tube surrounded by ice without any water.
The reaction flask is heated gently until no more droplets of bromoethane
collect.
Purifying the bromoethane
Impurities in the bromoethane include:

hydrogen bromide (although most of that will dissolve in the water if


you are collecting the bromoethane under water);
bromine - from the oxidation of bromide ions by the concentrated
sulphuric acid;
sulphur dioxide - formed when concentrated sulphuric acid oxidises
the bromide ions;
unreacted ethanol;

http://www.chemguide.co.uk/organicprops/haloalkanes/making.html (5 of 8)30/12/2004 11:20:54

making halogenoalkanes (haloalkanes)

ethoxyethane (diethyl ether) - formed by a side reaction between


the ethanol and the concentrated sulphuric acid.

The purification sequence


Stage 1
If you have collected the bromoethane under water, transfer the contents
of the collection flask to a separating funnel. Otherwise, pour the impure
bromoethane into the separating funnel, add some water and shake it.
Pour off and keep the bromoethane layer.

The water you discard will contain almost all of the hydrogen bromide, and
quite a lot of any bromine, sulphur dioxide and ethanol present as
impurities.
Stage 2
To get rid of any remaining acidic impurities (including the bromine and
sulphur dioxide), return the bromoethane to the separating funnel and
shake it with either sodium carbonate or sodium hydrogencarbonate
solution.
This reacts with any acids present liberating carbon dioxide and forming
http://www.chemguide.co.uk/organicprops/haloalkanes/making.html (6 of 8)30/12/2004 11:20:54

making halogenoalkanes (haloalkanes)

soluble salts.
Separate and retain the lower bromoethane layer as before.
Stage 3
Now wash the bromoethane with water in a separating funnel to remove
any remaining inorganic impurities (excess sodium carbonate solution,
etc). This time, transfer the lower bromoethane layer to a dry test tube.
Stage 4
Add some anhydrous calcium chloride to the tube, shake well and leave to
stand. The anhydrous calcium chloride is a drying agent and removes any
remaining water. It also absorbs ethanol, and so any remaining ethanol
may be removed as well (depending on how much calcium chloride you
use).
Stage 5
Transfer the dry bromoethane to a distillation flask and fractionally distil it,
collecting what distils over at between 35 and 40C.
In principle, this should remove any remaining organic impurities. In
practice, though, any ethoxyethane (which is perhaps the most likely
impurity left at this stage) has a boiling point very, very close to that of
bromoethane. It is unlikely that you will be able to separate the two.
If there is any ethanol left which hadn't been absorbed by the calcium
chloride, that would certainly be removed because its boiling point is much
higher.
What's left?
In my experience with A level students, virtually nothing if you are working
on a small scale!! Every time you go through a purification stage, you
inevitably lose some of what you are trying to collect.

Where would you like to go now?

http://www.chemguide.co.uk/organicprops/haloalkanes/making.html (7 of 8)30/12/2004 11:20:54

making halogenoalkanes (haloalkanes)

To the halogenoalkanes menu . . .


To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/haloalkanes/making.html (8 of 8)30/12/2004 11:20:54

halogenoalkanes (haloalkanes) and cyanide ions

REACTIONS BETWEEN HALOGENOALKANES


AND CYANIDE IONS

This page looks at the reaction between halogenoalkanes (haloalkanes


or alkyl halides) and cyanide ions from sodium or potassium cyanide
solution.
You will find a link to a separate page about the mechanisms for the
reaction.

Replacing a halogen by -CN


Reactions
If a halogenoalkane is heated under reflux with a solution of sodium or
potassium cyanide in ethanol, the halogen is replaced by a -CN group
and a nitrile is produced. Heating under reflux means heating with a
condenser placed vertically in the flask to prevent loss of volatile
substances from the mixture.
The solvent is important. If water is present you tend to get substitution
by -OH instead of -CN.

Note: A solution of potassium cyanide in water is quite


alkaline, and contains significant amounts of hydroxide ions.
These react with the halogenoalkane. This is discussed on
the page about the reactions between halogenoalkanes and
hydroxide ions.
Use the BACK button on your browser to return to this page if
you choose to follow this link.

http://www.chemguide.co.uk/organicprops/haloalkanes/cyanide.html (1 of 3)30/12/2004 11:20:56

halogenoalkanes (haloalkanes) and cyanide ions

For example, using 1-bromopropane as a typical primary halogenoalkane:

You could write the full equation rather than the ionic one, but it slightly
obscures what's going on:

The bromine (or other halogen) in the halogenoalkane is simply replaced


by a -CN group - hence a substitution reaction. In this example,
butanenitrile is formed.

Note: When you are naming nitriles, you have to remember


to include the carbon in the -CN group when you count the
longest chain. In this example, there are 4 carbons in the
longest chain - hence butanenitrile.

Secondary and tertiary halogenoalkanes behave similarly, although the


mechanism will vary depending on which sort of halogenoalkane you are
using.

Note: If you aren't sure what primary, secondary and tertiary


halogenoalkanes are, you should read the beginning of the
introduction to halogenoalkanes.
You will find the mechanisms for these reactions by following
this link.
Use the BACK button (or GO menu or HISTORY file) on your
browser if you want to return to this page after following
either of these links.

http://www.chemguide.co.uk/organicprops/haloalkanes/cyanide.html (2 of 3)30/12/2004 11:20:56

halogenoalkanes (haloalkanes) and cyanide ions

Why these reactions matter


This reaction with cyanide ions is a useful way of lengthening carbon
chains. For example, in the equations above, you start with a 3-carbon
chain and end up with a 4-carbon chain. There aren't very many simple
ways of making new carbon-carbon bonds.
It is fairly easy to change the -CN group at the end of the new chain into
other groups.

Note: If you want to find out more about nitriles and


reactions of the -CN group, you could explore the nitriles
menu.

Where would you like to go now?


To the halogenoalkanes menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/haloalkanes/cyanide.html (3 of 3)30/12/2004 11:20:56

Nitriles Menu

Understanding Chemistry

NITRILES MENU

Background . . .
An introduction to nitriles including their physical properties.
Preparation of nitriles . . .
Their preparation from halogenoalkanes, amides, and aldehydes
and ketones.
The hydrolysis of nitriles . . .
The hydrolysis of nitriles using acids or alkalis.
The reduction of nitriles . . .
The reduction of nitriles to primary amines.

Go to menu of other organic compounds . . .


Go to Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/nitrilemenu.html (1 of 2)30/12/2004 11:20:58

Nitriles Menu

http://www.chemguide.co.uk/organicprops/nitrilemenu.html (2 of 2)30/12/2004 11:20:58

an introduction to nitriles

INTRODUCING NITRILES

This page explains what nitriles are and looks at their simple physical
properties such as solubility and boiling points.

What are nitriles?


Nitriles contain the -CN group, and used to be known as cyanides.
Some simple nitriles
The smallest organic nitrile is ethanenitrile, CH3CN, (old name: methyl
cyanide or acetonitrile - and sometimes now called ethanonitrile).
Hydrogen cyanide, HCN, doesn't usually count as organic, even though it
contains a carbon atom.

Notice the triple bond between the carbon and nitrogen in the -CN group.
The three simplest nitriles are:

CH3CN

ethanenitrile

CH3CH2CN

propanenitrile

CH3CH2CH2CN

butanenitrile

When you are counting the length of the carbon chain, don't forget the
carbon in the -CN group. If the chain is branched, this carbon usually
counts as the number 1 carbon.

http://www.chemguide.co.uk/organicprops/nitriles/background.html (1 of 5)30/12/2004 11:21:14

an introduction to nitriles

Note: Compounds like this are formed when aldehydes react


with hydrogen cyanide. This is therefore the sort of branched
nitrile that you are most likely to come across at this level.

Physical properties
Boiling points
The small nitriles are liquids at room temperature.

nitrile

boiling point (C)

CH3CN

82

CH3CH2CN

97

CH3CH2CH2CN

116 - 118

Note: The majority of the data sheets I have looked at quote


this boiling range for butanenitrile. I don't know why it doesn't
seem to have a precise boiling point.

http://www.chemguide.co.uk/organicprops/nitriles/background.html (2 of 5)30/12/2004 11:21:14

an introduction to nitriles

These boiling points are very high for the size of the molecules - similar
to what you would expect if they were capable of forming hydrogen
bonds.
However, they don't form hydrogen bonds - they don't have a hydrogen
atom directly attached to an electronegative element.
They are just very polar molecules. The nitrogen is very electronegative
and the electrons in the triple bond are very easily pulled towards the
nitrogen end of the bond.
Nitriles therefore have strong permanent dipole-dipole attractions as well
as van der Waals dispersion forces between their molecules.

Note: If you aren't happy about intermolecular attractions


then you really ought to follow this link before you go on.
Use the BACK button on your browser to return to this page.

Solubility in water
Ethanenitrile is completely soluble in water, and the solubility then falls
as chain length increases.

nitrile
CH3CN
CH3CH2CN
CH3CH2CH2CN

solubility at 20C
miscible
10 g per 100 cm3 of water
3 g per 100 cm3 of water

http://www.chemguide.co.uk/organicprops/nitriles/background.html (3 of 5)30/12/2004 11:21:14

an introduction to nitriles

The reason for the solubility is that although nitriles can't hydrogen bond
with themselves, they can hydrogen bond with water molecules.
One of the slightly positive hydrogen atoms in a water molecule is
attracted to the lone pair on the nitrogen atom in a nitrile and a hydrogen
bond is formed.

There will also, of course, be dispersion forces and dipole-dipole


attractions between the nitrile and water molecules.
Forming these attractions releases energy. This helps to supply the
energy needed to separate water molecule from water molecule and
nitrile molecule from nitrile molecule before they can mix together.
As chain lengths increase, the hydrocarbon parts of the nitrile molecules
start to get in the way.
By forcing themselves between water molecules, they break the
relatively strong hydrogen bonds between water molecules without
replacing them by anything as good. This makes the process
energetically less profitable, and so solubility decreases.

Where would you like to go now?


To the nitriles menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

http://www.chemguide.co.uk/organicprops/nitriles/background.html (4 of 5)30/12/2004 11:21:14

an introduction to nitriles

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/nitriles/background.html (5 of 5)30/12/2004 11:21:14

the preparation of nitriles

MAKING NITRILES

This page looks at various ways of making nitriles - from


halogenoalkanes (haloalkanes or alkyl halides), from amides, and from
aldehydes and ketones. It pulls together information from pages dealing
with each of these kinds of compounds

Making nitriles from halogenoalkanes


The halogenoalkane is heated under reflux with a solution of sodium or
potassium cyanide in ethanol. The halogen is replaced by a -CN group
and a nitrile is produced. Heating under reflux means heating with a
condenser placed vertically in the flask to prevent loss of volatile
substances from the mixture.
The solvent is important. If water is present you tend to get substitution
by -OH instead of -CN.

Note: A solution of potassium cyanide in water is quite


alkaline, and contains significant amounts of hydroxide ions.
These react with the halogenoalkane. This reaction is
discussed on the page about the reactions between
halogenoalkanes and hydroxide ions.
Use the BACK button on your browser to return to this page if
you choose to follow this link.

http://www.chemguide.co.uk/organicprops/nitriles/preparation.html (1 of 5)30/12/2004 11:21:20

the preparation of nitriles

For example, using 1-bromopropane as a typical halogenoalkane:

You could write the full equation rather than the ionic one, but it slightly
obscures what's going on:

The bromine (or other halogen) in the halogenoalkane is simply replaced


by a -CN group - hence a substitution reaction. In this example,
butanenitrile is formed.

Note: If you want the mechanisms for these reactions (which


differ depending on exactly what sort of halogenoalkane you
are talking about), you can find them by following this link.
Use the BACK button on your browser to return to this page if
you choose to follow this link.

Making a nitrile by this method is a useful way of increasing the length of


a carbon chain. Having made the nitrile, the -CN group can easily be
modified to make other things - as you will find if you explore the nitriles
menu (link a the bottom of the page).

Making nitriles from amides


Nitriles can be made by dehydrating amides.
Amides are dehydrated by heating a solid mixture of the amide and
phosphorus(V) oxide, P4O10.
Water is removed from the amide group to leave a nitrile group, -CN. The
liquid nitrile is collected by simple distillation.
http://www.chemguide.co.uk/organicprops/nitriles/preparation.html (2 of 5)30/12/2004 11:21:20

the preparation of nitriles

For example, you will get ethanenitrile by dehydrating ethanamide.

Note: This is a just a flow scheme rather than a proper


equation. I haven't been able to find a single example of the
use of the full equation for this reaction. In fact the
phosphorus(V) oxide reacts with the water to produce
mixtures of phosphorus-containing acids.

Making nitriles from aldehydes and ketones


Aldehydes and ketones undergo an addition reaction with hydrogen
cyanide. The hydrogen cyanide adds across the carbon-oxygen double
bond in the aldehyde or ketone to produce a hydroxynitrile.
Hydroxynitriles used to be known as cyanohydrins.
For example, with ethanal (an aldehyde) you get 2-hydroxypropanenitrile:

With propanone (a ketone) you get 2-hydroxy-2-methylpropanenitrile:

In every example of this kind, the -OH group will be on the number 2
carbon atom - the one next to the -CN group.

http://www.chemguide.co.uk/organicprops/nitriles/preparation.html (3 of 5)30/12/2004 11:21:20

the preparation of nitriles

The reaction isn't normally done using hydrogen cyanide itself, because
this is an extremely poisonous gas. Instead, the aldehyde or ketone is
mixed with a solution of sodium or potassium cyanide in water to which a
little sulphuric acid has been added. The pH of the solution is adjusted to
about 4 - 5, because this gives the fastest reaction. The reaction
happens at room temperature.
The solution will contain hydrogen cyanide (from the reaction between
the sodium or potassium cyanide and the sulphuric acid), but still
contains some free cyanide ions. This is important for the mechanism.

Note: If you want the mechanism for this reaction, you will
find it explained if you follow this link.
Use the BACK button on your browser to return to this page.

These are useful reactions because they not only increase the number of
carbon atoms in a chain, but also introduce another reactive group as
well as the -CN group. The -OH group behaves just like the -OH group in
any alcohol with a similar structure.
For example, starting from a hydroxynitrile made from an aldehyde, you
can quite easily produce relatively complicated molecules like 2-amino
acids - the amino acids which are used to construct proteins.

http://www.chemguide.co.uk/organicprops/nitriles/preparation.html (4 of 5)30/12/2004 11:21:20

the preparation of nitriles

Note: The first step is the replacement of -OH in an alcohol


by chlorine.
The second step is the reaction between a halogenoalkane
and ammonia.
The final step is hydrolysis of a nitrile.
Use the BACK button on your browser to return to this page if
you follow any of these links.

Where would you like to go now?


To the nitriles menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/nitriles/preparation.html (5 of 5)30/12/2004 11:21:20

replacing the OH in alcohols by a halogen

REPLACING THE -OH GROUP IN ALCOHOLS BY


A HALOGEN

This page looks at reactions in which the -OH group in an alcohol is


replaced by a halogen such as chlorine or bromine. It includes a simple
test for an -OH group using phosphorus(V) chloride.

Reactions involving hydrogen halides


The general reaction looks like this:

Reaction with hydrogen chloride


Tertiary alcohols react reasonably rapidly with concentrated hydrochloric
acid, but for primary or secondary alcohols the reaction rates are too slow
for the reaction to be of much importance.
A tertiary alcohol reacts if it is shaken with with concentrated hydrochloric
acid at room temperature. A tertiary halogenoalkane (haloalkane or alkyl
halide) is formed

http://www.chemguide.co.uk/organicprops/alcohols/halogen.html (1 of 6)30/12/2004 11:21:31

replacing the OH in alcohols by a halogen

Note: If you don't know what primary, secondary and tertiary


alcohols are, you should read the introduction to alcohols
before you go on. The terms primary, secondary and tertiary
are used in exactly the same way with halogenoalkanes.
Use the BACK button on your browser to return to this page.

Replacing -OH by bromine


Rather than using hydrobromic acid, you usually treat the alcohol with a
mixture of sodium or potassium bromide and concentrated sulphuric acid.
This produces hydrogen bromide which reacts with the alcohol. The
mixture is warmed to distil off the bromoalkane.

Note: You will find practical details of this reaction on the page
about preparation of halogenoalkanes. You don't need to read
the beginning of that page because it is just a modification of
this one!
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/alcohols/halogen.html (2 of 6)30/12/2004 11:21:31

replacing the OH in alcohols by a halogen

Replacing -OH by iodine


In this case the alcohol is reacted with a mixture of sodium or potassium
iodide and concentrated phosphoric(V) acid, H3PO4, and the iodoalkane is
distilled off. The mixture of the iodide and phosphoric(V) acid produces
hydrogen iodide which reacts with the alcohol.

Phosphoric(V) acid is used instead of concentrated sulphuric acid because


sulphuric acid oxidises iodide ions to iodine and produces hardly any
hydrogen iodide.
A similar thing happens to some extent with bromide ions in the
preparation of bromoalkanes, but not enough to get in the way of the main
reaction. There is no reason why you couldn't use phosphoric(V) acid in
the bromide case instead of sulphuric acid if you wanted to.

Note: If you are interested in the reactions between halide


ions and concentrated sulphuric acid you could follow this link.
In the present context, all you would need to do is read the
beginning of that page.
If you choose to follow this link, use the BACK button on your
browser to return to this page.

http://www.chemguide.co.uk/organicprops/alcohols/halogen.html (3 of 6)30/12/2004 11:21:31

replacing the OH in alcohols by a halogen

Reacting alcohols with phosphorus halides


Reaction with phosphorus(III) chloride, PCl3
Alcohols react with liquid phosphorus(III) chloride (also called phosphorus
trichloride) to make chloroalkanes.

Reaction with phosphorus(V) chloride, PCl5


Solid phosphorus(V) chloride (phosphorus pentachloride) reacts violently
with alcohols at room temperature, producing clouds of hydrogen chloride
gas. It isn't a good choice as a way of making chloroalkanes, although it is
used as a test for -OH groups in organic chemistry.
To show that a substance was an alcohol, you would first have to eliminate
all the other things which also react with phosphorus(V) chloride. For
example, carboxylic acids (containing the -COOH group) react with it
(because of the -OH in -COOH), and so does water (H-OH).
If you have a neutral liquid not contaminated with water, and get clouds of
hydrogen chloride when you add phosphorus(V) chloride, then you have
an alcohol group present.

There are also side reactions involving the POCl3 reacting with the alcohol.
Other reactions involving phosphorus halides
Instead of using phosphorus(III) bromide or iodide, the alcohol is usually
heated under reflux with a mixture of red phosphorus and either bromine
or iodine.
The phosphorus first reacts with the bromine or iodine to give the
phosphorus(III) halide.

http://www.chemguide.co.uk/organicprops/alcohols/halogen.html (4 of 6)30/12/2004 11:21:31

replacing the OH in alcohols by a halogen

These then react with the alcohol to give the corresponding


halogenoalkane which can be distilled off.

Reacting alcohols with sulphur dichloride oxide (thionyl


chloride)
The reaction
Sulphur dichloride oxide (thionyl chloride) has the formula SOCl2.
Traditionally, the formula is written as shown, despite the fact that the
modern name writes the chlorine before the oxygen (alphabetical order).
The sulphur dichloride oxide reacts with alcohols at room temperature to
produce a chloroalkane. Sulphur dioxide and hydrogen chloride are given
off. Care would have to be taken because both of these are poisonous.

Why this reaction is useful


The big advantage that this reaction has over the use of either of the
phosphorus chlorides is that the two other products of the reaction
(sulphur dioxide and HCl) are both gases. That means that they separate
themselves from the reaction mixture.

Where would you like to go now?


To the alcohols menu . . .
http://www.chemguide.co.uk/organicprops/alcohols/halogen.html (5 of 6)30/12/2004 11:21:31

replacing the OH in alcohols by a halogen

To the menu of other organic compounds . . .


To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/alcohols/halogen.html (6 of 6)30/12/2004 11:21:31

halogenoalkanes (haloalkanes) and ammonia

REACTIONS BETWEEN HALOGENOALKANES


AND AMMONIA

This page looks at the reaction between halogenoalkanes (haloalkanes or


alkyl halides) and ammonia. This is a potentially very complicated series of
reactions, so it is important to know exactly what your examiners want.
You need to check your syllabus and past exam papers and mark
schemes.

Important! If you are a UK A level student, and don't have any


of this information, then it is essential that you get it! Find out
how by going to the syllabuses page.

The current page summarises the main facts about the reactions. If you
need details about the mechanisms, you will find a link below.

Reaction details and products


The halogenoalkane is heated with a concentrated solution of ammonia in
ethanol. The reaction is carried out in a sealed tube. You couldn't heat this
mixture under reflux, because the ammonia would simply escape up the
condenser as a gas.
We'll talk about the reaction using 1-bromoethane as a typical primary
halogenoalkane. There is no difference in the details of the reaction if you
chose a secondary or tertiary halogenoalkane instead. The equations
would just look more complicated than they already are!
You get a series of amines formed together with their salts. The reactions
happen one after another.
Making a primary amine

http://www.chemguide.co.uk/organicprops/haloalkanes/nh3.html (1 of 5)30/12/2004 11:21:37

halogenoalkanes (haloalkanes) and ammonia

The reaction happens in two stages. In the first stage, a salt is formed - in
this case, ethylammonium bromide. This is just like ammonium bromide,
except that one of the hydrogens in the ammonium ion is replaced by an
ethyl group.

There is then the possibility of a reversible reaction between this salt and
excess ammonia in the mixture.

The ammonia removes a hydrogen ion from the ethylammonium ion to


leave a primary amine - ethylamine.
The more ammonia there is in the mixture, the more the forward reaction is
favoured.

Note: You will find considerable disagreement in textbooks


and other sources about the exact nature of the products in this
reaction. Some of the information you'll come across is simply
wrong!
You can read the arguments about the products of this reaction
by following this link.
Warning! That page is in the mechanism section of the site.
Return to the current page using the BACK button on your
browser. If you use the links at the bottom of that page, you
could get seriously lost!

http://www.chemguide.co.uk/organicprops/haloalkanes/nh3.html (2 of 5)30/12/2004 11:21:37

halogenoalkanes (haloalkanes) and ammonia

Making a secondary amine


The reaction doesn't stop at a primary amine. The ethylamine also reacts
with bromoethane - in the same two stages as before.
In the first stage, you get a salt formed - this time, diethylammonium
bromide. Think of this as ammonium bromide with two hydrogens replaced
by ethyl groups.

There is again the possibility of a reversible reaction between this salt and
excess ammonia in the mixture.

The ammonia removes a hydrogen ion from the diethylammonium ion to


leave a secondary amine - diethylamine. A secondary amine is one which
has two alkyl groups attached to the nitrogen.
Making a tertiary amine
And still it doesn't stop! The diethylamine also reacts with bromoethane - in
the same two stages as before.
In the first stage, you get triethylammonium bromide.

There is again the possibility of a reversible reaction between this salt and
excess ammonia in the mixture.
http://www.chemguide.co.uk/organicprops/haloalkanes/nh3.html (3 of 5)30/12/2004 11:21:37

halogenoalkanes (haloalkanes) and ammonia

The ammonia removes a hydrogen ion from the triethylammonium ion to


leave a tertiary amine - triethylamine. A tertiary amine is one which has
three alkyl groups attached to the nitrogen.
Making a quaternary ammonium salt
The final stage! The triethylamine reacts with bromoethane to give
tetraethylammonium bromide - a quaternary ammonium salt (one in which
all four hydrogens have been replaced by alkyl groups).

This time there isn't any hydrogen left on the nitrogen to be removed. The
reaction stops here.

Note: This whole reaction sequence is a complete pain if you


are going to have to learn it. It is much, much easier to work it
out if you need to, provided you understand the mechanisms
for the reactions.
You can explore the mechanisms for the various stages of the
reaction by following this link. This will lead you to several
pages in the mechanism section of this site. If all you want to
do is make some sense of the above reactions, it would
probably pay you to just read the parts of those pages
concerned with primary halogenoalkanes like bromoethane.

http://www.chemguide.co.uk/organicprops/haloalkanes/nh3.html (4 of 5)30/12/2004 11:21:37

halogenoalkanes (haloalkanes) and ammonia

What do you actually get if you react bromoethane with ammonia?


Whatever you do, you get a mixture of all of the products (including both
amines and their salts) shown on this page.
To get mainly the quaternary ammonium salt, you can use a large excess
of bromoethane. If you look at the reactions going on, each one needs
additional bromoethane. If you provide enough, then the chances are that
the reaction will go to completion, given enough time.
On the other hand, if you use a very large excess of ammonia, the
chances are always greatest that a bromoethane molecule will hit an
ammonia molecule rather than one of the amines being formed. That will
help to prevent the formation of secondary (etc) amines.

Where would you like to go now?


To the halogenoalkanes menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/haloalkanes/nh3.html (5 of 5)30/12/2004 11:21:37

hydrolysis of nitriles

HYDROLYSING NITRILES

This page looks at the hydrolysis of nitriles under either acidic or alkaline
conditions to make carboxylic acids or their salts.

The hydrolysis of nitriles


Introduction
When nitriles are hydrolysed you can think of them reacting with water in
two stages - first to produce an amide, and then the ammonium salt of a
carboxylic acid.
For example, ethanenitrile would end up as ammonium ethanoate going
via ethanamide.

In practice, the reaction between nitriles and water would be so slow as


to be completely negligible. The nitrile is instead heated with either a
dilute acid such as dilute hydrochloric acid, or with an alkali such as
sodium hydroxide solution.
The end result is similar in all the cases, but the exact nature of the final
product varies depending on the conditions you use for the reaction.

Acidic hydrolysis of nitriles


The nitrile is heated under reflux with dilute hydrochloric acid. Instead of
getting an ammonium salt as you would do if the reaction only involved
water, you produce the free carboxylic acid.
For example, with ethanenitrile and hydrochloric acid you would get
http://www.chemguide.co.uk/organicprops/nitriles/hydrolysis.html (1 of 3)30/12/2004 11:21:41

hydrolysis of nitriles

ethanoic acid and ammonium chloride.

Why is the free acid formed rather than the ammonium salt? The
ethanoate ions in the ammonium ethanoate react with hydrogen ions
from the hydrochloric acid to produce ethanoic acid. Ethanoic acid is only
a weak acid and so once it has got the hydrogen ion, it tends to hang on
to it.

Alkaline hydrolysis of nitriles


The nitrile is heated under reflux with sodium hydroxide solution. This
time, instead of getting an ammonium salt as you would do if the reaction
only involved water, you get the sodium salt. Ammonia gas is given off as
well.
For example, with ethanenitrile and sodium hydroxide solution you would
get sodium ethanoate and ammonia.

The ammonia is formed from reaction between ammonium ions and


hydroxide ions.
If you wanted the free carboxylic acid in this case, you would have to
acidify the final solution with a strong acid such as dilute hydrochloric
acid or dilute sulphuric acid. The ethanoate ion in the sodium ethanoate
will react with hydrogen ions as mentioned above.

Where would you like to go now?


To the nitriles menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .
http://www.chemguide.co.uk/organicprops/nitriles/hydrolysis.html (2 of 3)30/12/2004 11:21:41

hydrolysis of nitriles

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/nitriles/hydrolysis.html (3 of 3)30/12/2004 11:21:41

reduction of nitriles

REDUCING NITRILES TO PRIMARY AMINES

This page looks at the reduction of nitriles to primary amines using either
lithium tetrahydridoaluminate(III) (lithium aluminium hydride) or hydrogen
and a metal catalyst.

The reduction of nitriles using LiAlH4


The reducing agent
Despite its name, the structure of the reducing agent is very simple.
There are four hydrogens ("tetrahydido") around the aluminium in a
negative ion (shown by the "ate" ending).
The "(III)" shows the oxidation state of the aluminium, and is often left out
because aluminium only ever shows the +3 oxidation state in its
compounds. To make the name shorter, that's what I shall do for the rest
of this page.

Note: It isn't important as far as the current page is


concerned, but if you want to understand more about
oxidation states (oxidation numbers), you will find them
explained if you follow this link.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/nitriles/reduction.html (1 of 4)30/12/2004 11:21:45

reduction of nitriles

The structure of LiAlH4 is:

In the negative ion, one of the bonds is a co-ordinate covalent (dative


covalent) bond using the lone pair on a hydride ion (H-) to form a bond
with an empty orbital on the aluminium.

Note: Follow this link if you aren't happy about co-ordinate


covalent (dative covalent) bonding. Again, it isn't particularly
important as far as the current page is concerned.
Use the BACK button on your browser to return to this page.

The overall reaction


The nitrile reacts with the lithium tetrahydridoaluminate in solution in
ethoxyethane (diethyl ether, or just "ether") followed by treatment of the
product of that reaction with a dilute acid.
Overall, the carbon-nitrogen triple bond is reduced to give a primary
amine. Primary amines contain the -NH2 group.
For example, with ethanenitrile you get ethylamine:

Notice that this is a simplified equation - perfectly acceptable to UK A


level examiners. [H] means "hydrogen from a reducing agent".

http://www.chemguide.co.uk/organicprops/nitriles/reduction.html (2 of 4)30/12/2004 11:21:45

reduction of nitriles

Note: If you know about the reduction of aldehydes and


ketones, you may know that they are also reduced by the
similar compound NaBH4.
However, NaBH4 isn't a strong enough reducing agent to
reduce nitriles.

The reduction of nitriles using hydrogen and a metal


catalyst
The carbon-nitrogen triple bond in a nitrile can also be reduced by
reaction with hydrogen gas in the presence of a variety of metal catalysts.
Commonly quoted catalysts are palladium, platinum or nickel.
The reaction will take place at a raised temperature and pressure. It is
impossible to give exact details because it will vary from catalyst to
catalyst.
For example, ethanenitrile can be reduced to ethylamine by reaction with
hydrogen in the presence of a palladium catalyst.

Note: Notice that this time the hydrogen is written normally


as H2. This is a proper equation involving hydrogen gas - not
a simplification.

http://www.chemguide.co.uk/organicprops/nitriles/reduction.html (3 of 4)30/12/2004 11:21:45

reduction of nitriles

Where would you like to go now?


To the nitriles menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/nitriles/reduction.html (4 of 4)30/12/2004 11:21:45

reduction of aldehydes and ketones

REDUCTION OF ALDEHYDES AND KETONES

This page looks at the reduction of aldehydes and ketones by two similar
reducing agents - lithium tetrahydridoaluminate(III) (also known as lithium
aluminium hydride) and sodium tetrahydridoborate(III) (sodium
borohydride).

Background to the reactions


The reducing agents
Despite the fearsome names, the structures of the two reducing agents
are very simple. In each case, there are four hydrogens ("tetrahydido")
around either aluminium or boron in a negative ion (shown by the "ate"
ending).
The "(III)" shows the oxidation state of the aluminium or boron, and is
often left out because these elements only ever show the +3 oxidation
state in their compounds. To make the names shorter, that's what I shall
do for the rest of this page.

Note: It isn't important as far as the current page is


concerned, but if you want to understand more about oxidation
states (oxidation numbers), you will find them explained if you
follow this link.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/carbonyls/reduction.html (1 of 7)30/12/2004 11:21:52

reduction of aldehydes and ketones

The formulae of the two compounds are LiAlH4 and NaBH4.


Their structures are:

In each of the negative ions, one of the bonds is a co-ordinate covalent


(dative covalent) bond using the lone pair on a hydride ion (H-) to form a
bond with an empty orbital on the aluminium or boron.

Note: Follow this link if you aren't happy about co-ordinate


covalent (dative covalent) bonding.
Use the BACK button on your browser to return to this page.

The overall reactions


The reduction of an aldehyde
You get exactly the same organic product whether you use lithium
tetrahydridoaluminate or sodium tetrahydridoborate.
For example, with ethanal you get ethanol:

Notice that this is a simplified equation - perfectly acceptable to UK A level


examiners. [H] means "hydrogen from a reducing agent".

http://www.chemguide.co.uk/organicprops/carbonyls/reduction.html (2 of 7)30/12/2004 11:21:52

reduction of aldehydes and ketones

In general terms, reduction of an aldehyde leads to a primary alcohol.

Note: If you aren't sure about types of alcohol it is essential to


follow this link before you go on. You only need to read the
beginning of that page.
Use the BACK button on your browser to return to this page.

The reduction of a ketone


Again the product is the same whichever of the two reducing agents you
use.
For example, with propanone you get propan-2-ol:

Reduction of a ketone leads to a secondary alcohol.

Reaction details
Using lithium tetrahydridoaluminate (lithium aluminium hydride)
Lithium tetrahydridoaluminate is much more reactive than sodium
tetrahydridoborate. It reacts violently with water and alcohols, and so any
reaction must exclude these common solvents.
The reactions are usually carried out in solution in an ether such as
ethoxyethane (diethyl ether). The reaction happens at room temperature,
and takes place in two separate stages.
In the first stage, a salt is formed containing a complex aluminium ion. The
http://www.chemguide.co.uk/organicprops/carbonyls/reduction.html (3 of 7)30/12/2004 11:21:52

reduction of aldehydes and ketones

following equations show what happens if you start with a general


aldehyde or ketone. R and R' can be any combination of hydrogen or alkyl
groups.

The product is then treated with a dilute acid (such as dilute sulphuric acid
or dilute hydrochloric acid) to release the alcohol from the complex ion.

The alcohol formed can be recovered from the mixture by fractional


distillation.

Note: You can see why UK A level examiners are happy with
the equations showing H in square brackets!
On the other hand, you may well be expected to know that the
reaction is done initially in solution in ethoxyethane followed
by treatment with acid.

http://www.chemguide.co.uk/organicprops/carbonyls/reduction.html (4 of 7)30/12/2004 11:21:52

reduction of aldehydes and ketones

Using sodium tetrahydridoborate (sodium borohydride)


Sodium tetrahydridoborate is a more gentle (and therefore safer) reagent
than lithium tetrahydridoaluminate. It can be used in solution in alcohols or
even solution in water - provided the solution is alkaline.
I have a problem over describing the reaction conditions, because it
seems to be used in so many different ways. Practical details found from
various university sites vary widely, and don't necessarily agree with what
theoretical sources say!
In what follows, I am choosing one out of many different methods. I have
chosen this one largely because I think I understand what is going on!
Solid sodium tetrahydridoborate is added to a solution of the aldehyde or
ketone in an alcohol such as methanol, ethanol or propan-2-ol. Depending
on which recipe you read, it is either heated under reflux or left for some
time around room temperature. This almost certainly varies depending on
the nature of the aldehyde or ketone.
At the end of this time, a complex similar to the previous one is formed.

In the second stage of the reaction, water is added and the mixture is
boiled to release the alcohol from the complex.

Again, the alcohol formed can be recovered from the mixture by fractional
distillation.

http://www.chemguide.co.uk/organicprops/carbonyls/reduction.html (5 of 7)30/12/2004 11:21:52

reduction of aldehydes and ketones

Note: Experimental variants on this use sodium hydroxide


solution or dilute acids instead of water in the second stage.
Another variation treats the original aldehyde or ketone with
the sodium tetrahydridoborate dissolved in sodium hydroxide
solution, adding acid at the end to destroy excess sodium
tetrahydridoborate. Unfortunately, none of the experimental
sources I have looked at explain what is happening in the
second stage in any detail.
The complication doesn't actually stop here! The boron
complex formed at the end of the first stage is almost certainly
more complicated than this. At least a part of the organic bits
attached to the boron will probably come from whatever
alcohol molecules are used as the solvent for the reaction.
I have included this detail (even though UK A level students
don't need it) because I object to giving simplifications that you
might have to completely unlearn in the future if you do a
Chemistry degree.
You will find a simplified (for UK A level purposes) mechanism
for the reaction by following this link. This mechanism is
simplified to the point of being wrong, so if you are working
outside of the UK A level system, don't bother to look at it!
This will take you to another part of the site. Use the BACK
button on your browser to return to this page if you want to
continue to explore aldehyde and ketone reactions.

http://www.chemguide.co.uk/organicprops/carbonyls/reduction.html (6 of 7)30/12/2004 11:21:52

reduction of aldehydes and ketones

Where would you like to go now?


To the aldehydes and ketones menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/carbonyls/reduction.html (7 of 7)30/12/2004 11:21:52

halogenoalkanes (haloalkanes) and silver nitrate

REACTIONS INVOLVING HALOGENOALKANES


AND SILVER NITRATE SOLUTION

This page looks at how silver nitrate solution can be used as part of a test
for halogenoalkanes (haloalkanes or alkyl halides), and also as a means
of measuring their relative reactivities.

Testing for halogenoalkanes


Silver nitrate solution can be used to find out which halogen is present in
a suspected halogenoalkane. The most effective way is to do a
substitution reaction which turns the halogen into a halide ion, and then to
test for that ion with silver nitrate solution.
Doing the reaction
The halogenoalkane is warmed with some sodium hydroxide solution in a
mixture of ethanol and water. Everything will dissolve in this mixture and
so you can get a good reaction.
The halogen atom is displaced as a halide ion:

Note: This reaction is described in more detail on the page


about reactions of halogenoalkanes with hydroxide ions.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/haloalkanes/agno3.html (1 of 6)30/12/2004 11:22:00

halogenoalkanes (haloalkanes) and silver nitrate

There is no need to make this reaction go to completion. The silver nitrate


test is sensitive enough to detect fairly small concentrations of halide ions.
The mixture is acidified by adding dilute nitric acid. This prevents
unreacted hydroxide ions reacting with the silver ions. Then silver nitrate
solution is added.
Various precipitates may be formed from the reaction between the silver
and halide ions:

ion present

observation

Cl-

white precipitate

Br-

very pale cream precipitate

I-

very pale yellow precipitate

Confirming the precipitates


It is actually quite difficult to distinguish between these colours, especially
if there isn't much precipitate. You can sort out which precipitate you
have by adding ammonia solution.

original precipitate

AgCl

observation
precipitate dissolves to give a colourless
solution

http://www.chemguide.co.uk/organicprops/haloalkanes/agno3.html (2 of 6)30/12/2004 11:22:00

halogenoalkanes (haloalkanes) and silver nitrate

AgBr

precipitate is almost unchanged using dilute


ammonia solution, but dissolves in
concentrated ammonia solution to give a
colourless solution

AgI

precipitate is insoluble in ammonia solution


of any concentration

Note: You can read more about these tests (including an


explanation of this reaction involving ammonia) by following
this link to the page about testing for halide ions.
Use the BACK button on your browser to return to this page.

Comparing halogenoalkane reactivities


Background
In this case, various halogenoalkanes are treated with a solution of silver
nitrate in a mixture of ethanol and water. Nothing else is added. After
varying lengths of time precipitates appear as halide ions (produced from
reactions of the halogenoalkanes) react with the silver ions present.
As long as you are doing everything under controlled conditions (same
amounts of everything, same temperature and so on), the time taken
gives a good guide to the reactivity of the halogenoalkanes - the quicker
the precipitate appears, the more reactive the halogenoalkane.
The halide ion is formed in one of two ways, depending on the type of
halogenoalkane you have present - primary, secondary or tertiary.

http://www.chemguide.co.uk/organicprops/haloalkanes/agno3.html (3 of 6)30/12/2004 11:22:00

halogenoalkanes (haloalkanes) and silver nitrate

Note: If you aren't sure what primary, secondary and tertiary


halogenoalkanes are, you should read the beginning of the
introduction to halogenoalkanes by following this link before
you go on.
Use the BACK button on your browser to return to this page.

For a primary halogenoalkane, the main reaction is one between the


halogenoalkane and water in the solvent.

A tertiary halogenoalkane ionises to a very small extent of its own accord.

Secondary halogenoalkanes do a bit of both of these.

Comparing the reaction rates as you change the halogen


You would have to keep the type of halogenoalkane (primary, secondary
or tertiary) constant, but vary the halogen. You might, for example,
compare the times taken to produce a precipitate from this series of
primary halogenoalkanes:

Obviously, the time taken for a precipitate of silver bromide to appear will
depend on how much of everything you use and the temperature at which
the reaction is carried out. But the pattern of results is always the same.
For example:

http://www.chemguide.co.uk/organicprops/haloalkanes/agno3.html (4 of 6)30/12/2004 11:22:00

halogenoalkanes (haloalkanes) and silver nitrate

A primary iodo compound produces a precipitate quite quickly.

A primary bromo compound takes longer to give a precipitate.

A primary chloro compound probably won't give any precipitate


until well after you have lost interest in the whole thing!

The order of reactivity reflects the strengths of the carbon-halogen bonds.


The carbon-iodine bond is the weakest and the carbon-chlorine the
strongest of the three bonds. In order for a halide ion to be produced, the
carbon-halogen bond has to be broken. The weaker the bond, the easier
that is.

Comparing the reaction rates of primary, secondary and tertiary


halogenoalkanes
You would need to keep the halogen atom constant. It is common to use
bromides because they have moderate reaction rates. You could, for
example, compare the reactivity of these compounds:

Again, the actual times taken will vary with reaction conditions, but the
pattern will always be the same.
For example:

The tertiary halide produces a precipitate almost instantly.


The secondary halide gives a slight precipitate after a few
seconds. The precipitate thickens up with time.
The primary halide may take considerably longer to produce a

http://www.chemguide.co.uk/organicprops/haloalkanes/agno3.html (5 of 6)30/12/2004 11:22:00

halogenoalkanes (haloalkanes) and silver nitrate

precipitate.
It is more difficult to explain the reason for this, because it needs a fairly
intimate knowledge of the mechanisms involved in the reactions. It
reflects the change in the way that the halide ion is produced as you go
from primary to secondary to tertiary halogenoalkanes.

Note: You can read about the different mechanisms for this
reaction in the mechanism section of this site if you are
interested.
Use the BACK button on your browser to return to this page.

Where would you like to go now?


To the halogenoalkanes menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/haloalkanes/agno3.html (6 of 6)30/12/2004 11:22:00

grignard reagents

AN INTRODUCTION TO GRIGNARD REAGENTS

This page takes an introductory look at how Grignard reagents are made
from halogenoalkanes (haloalkanes or alkyl halides), and introduces some
of their reactions.

Making Grignard reagents


What are Grignard reagents?
A Grignard reagent has a formula RMgX where X is a halogen, and R is
an alkyl or aryl (based on a benzene ring) group. For the purposes of this
page, we shall take R to be an alkyl group.
A typical Grignard reagent might be CH3CH2MgBr.
The preparation of a Grignard reagent
Grignard reagents are made by adding the halogenoalkane to small bits of
magnesium in a flask containing ethoxyethane (commonly called diethyl
ether or just "ether"). The flask is fitted with a reflux condenser, and the
mixture is warmed over a water bath for 20 - 30 minutes.

Everything must be perfectly dry because Grignard reagents react with


water (see below).

Warning! Ethoxyethane (ether) is very dangerous to work


with. It is an anaesthetic, and is extremely inflammable. Under
no circumstances should you try to carry out this reaction
without properly qualified guidance.

http://www.chemguide.co.uk/organicprops/haloalkanes/grignard.html (1 of 10)30/12/2004 11:22:14

grignard reagents

Any reactions using the Grignard reagent are carried out with the mixture
produced from this reaction. You can't separate it out in any way.

Reactions of Grignard reagents


Grignard reagents and water
Grignard reagents react with water to produce alkanes. This is the reason
that everything has to be very dry during the preparation above.
For example:

The inorganic product, Mg(OH)Br, is referred to as a "basic bromide". You


can think of it as a sort of half-way stage between magnesium bromide
and magnesium hydroxide.

Grignard reagents and carbon dioxide


Grignard reagents react with carbon dioxide in two stages. In the first, you
get an addition of the Grignard reagent to the carbon dioxide.
Dry carbon dioxide is bubbled through a solution of the Grignard reagent
in ethoxyethane, made as described above.
For example:

http://www.chemguide.co.uk/organicprops/haloalkanes/grignard.html (2 of 10)30/12/2004 11:22:14

grignard reagents

The product is then hydrolysed (reacted with water) in the presence of a


dilute acid. Typically, you would add dilute sulphuric acid or dilute
hydrochloric acid to the solution formed by the reaction with the CO2.
A carboxylic acid is produced with one more carbon than the original
Grignard reagent.
The usually quoted equation is (without the red bits):

Almost all sources quote the formation of a basic halide such as Mg(OH)
Br as the other product of the reaction. That's actually misleading because
these compounds react with dilute acids. What you end up with would be
a mixture of ordinary hydrated magnesium ions, halide ions and sulphate
or chloride ions - depending on which dilute acid you added.

Note: What you need to learn about this depends on what


your examiners want. The only way to find that out is to look at
old exam papers and mark schemes. If you are a UK A level
student and haven't got copies of these, find out how to get
hold of them by going to the syllabuses page to find your
Exam Board's web address.

http://www.chemguide.co.uk/organicprops/haloalkanes/grignard.html (3 of 10)30/12/2004 11:22:14

grignard reagents

Grignard reagents and carbonyl compounds


What are carbonyl compounds?
Carbonyl compounds contain the C=O double bond. The simplest ones
have the form:

R and R' can be the same or different, and can be an alkyl group or
hydrogen.

Note: Other carbonyl compounds also react with Grignard


reagents, but these are all that are required for UK A level
purposes.

If one (or both) of the R groups are hydrogens, the compounds are called
aldehydes. For example:

If both of the R groups are alkyl groups, the compounds are called
ketones. Examples include:

http://www.chemguide.co.uk/organicprops/haloalkanes/grignard.html (4 of 10)30/12/2004 11:22:14

grignard reagents

The general reaction between Grignard reagents and carbonyl


compounds
The reactions between the various sorts of carbonyl compounds and
Grignard reagents can look quite complicated, but in fact they all react in
the same way - all that changes are the groups attached to the carbonoxygen double bond.
It is much easier to understand what is going on by looking closely at the
general case (using "R" groups rather than specific groups) - and then
slotting in the various real groups as and when you need to.
The reactions are essentially identical to the reaction with carbon dioxide all that differs is the nature of the organic product.
In the first stage, the Grignard reagent adds across the carbon-oxygen
double bond:

Dilute acid is then added to this to hydrolyse it. (I am using the normally
accepted equation ignoring the fact that the Mg(OH)Br will react further
with the acid.)

An alcohol is formed. One of the key uses of Grignard reagents is the


ability to make complicated alcohols easily.
What sort of alcohol you get depends on the carbonyl compound you
started with - in other words, what R and R' are.

http://www.chemguide.co.uk/organicprops/haloalkanes/grignard.html (5 of 10)30/12/2004 11:22:14

grignard reagents

The reaction between Grignard reagents and methanal


In methanal, both R groups are hydrogen. Methanal is the simplest
possible aldehyde.

Assuming that you are starting with CH3CH2MgBr and using the general
equation above, the alcohol you get always has the form:

Since both R groups are hydrogen atoms, the final product will be:

A primary alcohol is formed. A primary alcohol has only one alkyl group
attached to the carbon atom with the -OH group on it.
You could obviously get a different primary alcohol if you started from a
different Grignard reagent.

The reaction between Grignard reagents and other aldehydes


The next biggest aldehyde is ethanal. One of the R groups is hydrogen
and the other CH3.

http://www.chemguide.co.uk/organicprops/haloalkanes/grignard.html (6 of 10)30/12/2004 11:22:14

grignard reagents

Again, think about how that relates to the general case. The alcohol
formed is:

So this time the final product has one CH3 group and one hydrogen
attached:

A secondary alcohol has two alkyl groups (the same or different) attached
to the carbon with the -OH group on it.
You could change the nature of the final secondary alcohol by either:

changing the nature of the Grignard reagent - which would change


the CH3CH2 group into some other alkyl group;
changing the nature of the aldehyde - which would change the CH3
group into some other alkyl group.

The reaction between Grignard reagents and ketones


Ketones have two alkyl groups attached to the carbon-oxygen double
bond. The simplest one is propanone.

http://www.chemguide.co.uk/organicprops/haloalkanes/grignard.html (7 of 10)30/12/2004 11:22:14

grignard reagents

This time when you replace the R groups in the general formula for the
alcohol produced you get a tertiary alcohol.

A tertiary alcohol has three alkyl groups attached to the carbon with the OH attached. The alkyl groups can be any combination of same or
different.
You could ring the changes on the product by

changing the nature of the Grignard reagent - which would change


the CH3CH2 group into some other alkyl group;
changing the nature of the ketone - which would change the CH3
groups into whatever other alkyl groups you choose to have in the
original ketone.

Why do Grignard reagents react with carbonyl compounds?


The mechanisms for these reactions aren't required by any UK A level
syllabuses, but you might need to know a little about the nature of
Grignard reagents.
The bond between the carbon atom and the magnesium is polar. Carbon
is more electronegative than magnesium, and so the bonding pair of
electrons is pulled towards the carbon.
That leaves the carbon atom with a slight negative charge.
http://www.chemguide.co.uk/organicprops/haloalkanes/grignard.html (8 of 10)30/12/2004 11:22:14

grignard reagents

Note: If you aren't sure about electronegativity, you can read


about it in an organic context by following this link.
Use the BACK button on your browser to return to this page.

The carbon-oxygen double bond is also highly polar with a significant


amount of positive charge on the carbon atom. The nature of this bond is
described in detail elsewhere on this site.

Note: If you are interested, you could follow this link to the
bonding in a carbonyl group. Reading the last couple of
paragraphs on that page would be enough in the present
context.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/haloalkanes/grignard.html (9 of 10)30/12/2004 11:22:14

grignard reagents

The Grignard reagent can therefore serve as a nucleophile because of


the attraction between the slight negativeness of the carbon atom in the
Grignard reagent and the positiveness of the carbon in the carbonyl
compound.
A nucleophile is a species that attacks positive (or slightly positive)
centres in other molecules or ions.

Note: I have included this because one of the UK A level


syllabuses says that candidates should "recall that Grignard
reagents act as nucleophiles". That is all you need to know!
The mechanism is more complex than this suggests at first
sight, and isn't required. You won't find these mechanisms
anywhere on this site.

Where would you like to go now?


To the halogenoalkanes menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/haloalkanes/grignard.html (10 of 10)30/12/2004 11:22:14

uses of halogenoalkanes

USES OF HALOGENOALKANES

This page looks at some of the uses of halogenoalkanes (haloalkanes or


alkyl halides) as required by UK A level syllabuses.

CFCs and their replacements


What are CFCs?
CFCs are chlorofluorocarbons - compounds containing carbon with
chlorine and fluorine atoms attached. Two common CFCs are:
CFC-11

CCl3F

CFC-12

CCl2F2

How to work out the formula from the CFC number: You
add 90 to the CFC number to give a new number. For
example, CFC-11 generates the number 101 (11+90)
The first digit of the new number tells you how many carbon
atoms there are. The second number tells you the number of
hydrogens, and the third the number of fluorines. You
calculate the number of chlorines from the formula Cl = 2(C
+1)-H-F.

http://www.chemguide.co.uk/organicprops/haloalkanes/uses.html (1 of 5)30/12/2004 11:22:17

uses of halogenoalkanes

Uses of CFCs
CFCs are non-flammable and not very toxic. They therefore had a large
number of uses.
They were used as refrigerants, propellants for aerosols, for generating
foamed plastics like expanded polystyrene or polyurethane foam, and as
solvents for dry cleaning and for general degreasing purposes.
Unfortunately, CFCs are largely responsible for destroying the ozone
layer. In the high atmosphere, the carbon-chlorine bonds break to give
chlorine free radicals. It is these radicals which destroy ozone. CFCs are
now being replaced by less environmentally harmful compounds.

Note: You will find the mechanism for the depletion of ozone
by the chlorine radicals produced from CFCs on the
introduction to catalysis page in the physical chemistry
section of this site.
Use the BACK button on your browser to return to this page.

CFCs can also cause global warming. One molecule of CFC-11, for
example, has a global warming potential about 5000 times greater than a
molecule of carbon dioxide.
On the other hand, there is far more carbon dioxide in the atmosphere
than CFCs, so global warming isn't the major problem associated with
them.

Replacements for CFCs


These are still mainly halogenoalkanes, although simple alkanes such as
butane can be used for some applications (for example, as aerosol
propellants).

http://www.chemguide.co.uk/organicprops/haloalkanes/uses.html (2 of 5)30/12/2004 11:22:17

uses of halogenoalkanes

Hydrochlorofluorocarbons, HCFCs
These are carbon compounds which contain hydrogen as well as
halogen atoms. For example:
HCFC-22

CHClF2

The formula can be worked out from the number in the name in exactly
the same way as for CFCs.
These have a shorter life in the atmosphere than CFCs, and much of
them is destroyed in the low atmosphere and so doesn't reach the ozone
layer. HCFC-22 has only about one-twentieth of the effect on the ozone
layer as a typical CFC.
Hydrofluorocarbons, HFCs
These are compounds containing only hydrogen and fluorine attached to
carbon. For example:
HFC-134a

CH2F-CF3

Because these HCFCs don't contain any chlorine, they have zero effect
on the ozone layer. HFC-134a is now widely used in refrigerants, for
blowing foamed plastics and as a propellant in aerosols.
Hydrocarbons
Again, these have no effect on the ozone layer, but they do have a downside. They are highly flammable and are involved in environmental
problems such as the formation of photochemical smog.

Other uses of organic halogeno compounds


In making plastics

http://www.chemguide.co.uk/organicprops/haloalkanes/uses.html (3 of 5)30/12/2004 11:22:17

uses of halogenoalkanes

Strictly speaking, the compounds we are talking about here are


halogenoalkenes, not halogenoalkanes, but one of the UK A level
syllabuses includes them in this section.
Chloroethene, CH2=CHCl, is used to make poly(chloroethene) - usually
known as PVC.
Tetrafluoroethene, CF2=CF2, is used to make poly(tetrafluoroethene) PTFE.

Note: These are both discussed in detail on the page about


polymerisation of alkenes.
Use the BACK button on your browser to return to this page.

Lab uses of halogenoalkanes


If you refer to the halogenoalkanes menu (see below), you will find that
they react with lots of things leading to a wide range of different organic
products.
Halogenoalkanes are therefore useful in the lab as intermediates in
making other organic chemicals.

Where would you like to go now?


To the halogenoalkanes menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

http://www.chemguide.co.uk/organicprops/haloalkanes/uses.html (4 of 5)30/12/2004 11:22:17

uses of halogenoalkanes

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/haloalkanes/uses.html (5 of 5)30/12/2004 11:22:17

polymerisation of alkenes

THE POLYMERISATION OF ALKENES

This page looks at the polymerisation of alkenes to produce polymers


like poly(ethene) (usually known as polythene, and sometimes as
polyethylene), poly(propene) (old name: polypropylene), PVC and PTFE.
It also looks briefly at how the structure of the polymers affects their
properties and uses.

Poly(ethene) (polythene or polyethylene)


Low density poly(ethene): LDPE
Manufacture
In common with everything else on this page, this is an example of
addition polymerisation.
An addition reaction is one in which two or more molecules join together
to give a single product. During the polymerisation of ethene, thousands
of ethene molecules join together to make poly(ethene) - commonly
called polythene.

The number of molecules joining up is very variable, but is in the region


of 2000 to 20000.
Conditions
Temperature: about 200C
Pressure:
about 2000 atmospheres
a small amount of oxygen as an
Initiator:
impurity

http://www.chemguide.co.uk/organicprops/alkenes/polymerisation.html (1 of 12)30/12/2004 11:22:27

polymerisation of alkenes

Note: If you want the mechanism for this reaction, you can
get it by following this link to the mechanism section of the
site.
Use the BACK button on your browser if you want to return to
this page.

Properties and uses


Low density poly(ethene) has quite a lot of branching along the
hydrocarbon chains, and this prevents the chains from lying tidily close to
each other. Those regions of the poly(ethene) where the chains lie close
to each other and are regularly packed are said to be crystalline. Where
the chains are a random jumble, it is said to be amorphous. Low density
poly(ethene) has a significant proportion of amorphous regions.

Note: Various sources quote estimated values for the


proportion of low density poly(ethene) which is crystalline.
These vary from 50% to 75%. I have no idea what the correct
value is!

One chain is held to its neighbours in the structure by van der Waals
dispersion forces. Those attractions will be greater if the chains are close
to each other. The amorphous regions where the chains are inefficiently
packed lower the effectiveness of the van der Waals attractions and so
lower the melting point and strength of the polymer. They also lower the
density of the polymer (hence: "low density poly(ethene)").

http://www.chemguide.co.uk/organicprops/alkenes/polymerisation.html (2 of 12)30/12/2004 11:22:27

polymerisation of alkenes

Note: Follow this link if you don't understand van der Waals
forces.
Use the BACK button on your browser to return to this page.

Low density poly(ethene) is used for familiar things like plastic carrier
bags and other similar low strength and flexible sheet materials.

High density poly(ethene): HDPE


Manufacture
This is made under quite different conditions from low density poly
(ethene).
Conditions
Temperature: about 60C
Pressure:
low - a few atmospheres
Ziegler-Natta catalysts or other
Catalyst:
metal compounds
Ziegler-Natta catalysts are mixtures of titanium compounds like titanium
(III) chloride, TiCl3, or titanium(IV) chloride, TiCl4, and compounds of
aluminium like aluminium triethyl, Al(C2H5)3. There are all sorts of other
catalysts constantly being developed.
These catalysts work by totally different mechanisms from the high
pressure process used to make low density poly(ethene). The chains
grow in a much more controlled - much less random - way.
Properties and uses
High density poly(ethene) has very little branching along the hydrocarbon
http://www.chemguide.co.uk/organicprops/alkenes/polymerisation.html (3 of 12)30/12/2004 11:22:27

polymerisation of alkenes

chains - the crystallinity is 95% or better. This better packing means that
van der Waals attractions between the chains are greater and so the
plastic is stronger and has a higher melting point. Its density is also
higher because of the better packing and smaller amount of wasted
space in the structure.

Note: Despite the names, HDPE is only about 3% denser


than LDPE! HDPE has a density of about 0.95 g cm-3
compared with about 0.92 for LDPE.

High density poly(ethene) is used to make things like plastic milk bottles
and similar containers, washing up bowls, plastic pipes and so on. Look
for the letters HDPE near the recycling symbol.

Poly(propene) (polypropylene): PP
Poly(propene) is manufactured using Ziegler-Natta and other modern
catalysts. There are three variants on the structure of poly(propene)
which you may need to know about, but we'll start from the beginning
with a general structure which fits all of them.

Note: You should check your syllabus to find out exactly


how much you need to know. It is pointless getting bogged
down in this if you don't need to! If you are a UK A level
student and haven't got a copy of your syllabus, follow this
link to find out how to get one.
Use the BACK button (or HISTORY file or GO menu) on your
browser to return to this page later.

http://www.chemguide.co.uk/organicprops/alkenes/polymerisation.html (4 of 12)30/12/2004 11:22:27

polymerisation of alkenes

The general structure


If your syllabus simply mentions the structure of poly(propene) with no
more detail, this is adequate.
The trick is to think about the shape of the propene in the right way:

Now line lots of them up in a row and join them together. Notice that the
double bonds are all replaced by single bonds in the process.

In a simple equation form, this is normally written as:

The three variations on this structure


You have got to remember that the diagrams above are 2-dimensional.
Real poly(propene) chains are 3-dimensional. There are three different
http://www.chemguide.co.uk/organicprops/alkenes/polymerisation.html (5 of 12)30/12/2004 11:22:27

polymerisation of alkenes

sorts of poly(propene) depending in how the CH3 groups are arranged in


space.
These are called isotactic, atactic and syndiotactic poly(propene). The
commonly used version is isotactic poly(propene).
Isotactic poly(propene)
A bit of the isotactic poly(propene) chain looks like this:

Note: Dotted lines show bonds going back into the screen or
paper, and wedge shapes show them coming out towards
you. If you aren't very happy about the various ways of
drawing organic structures it might be worth following this link
before you go on.
Use the BACK button on your browser to return to this page.

This very regular arrangement of the CH3 groups makes it possible for
the chains to pack close together and so maximise the amount of van der
Waals bonding between them. That means that isotactic poly(propene) is
quite strong either as a solid object or when it is drawn into fibres.
This is the common form of poly(propene) which is used to make plastic
crates and ropes amongst many other things. Look for the letters PP
near the recycling symbol.
Atactic poly(propene)
In atactic poly(propene) the CH3 groups are orientated randomly along
http://www.chemguide.co.uk/organicprops/alkenes/polymerisation.html (6 of 12)30/12/2004 11:22:27

polymerisation of alkenes

the chain.

This lack of regularity makes it impossible for the chains to lie closely
together and so the van der Waals attractions between them are weaker.
Atactic poly(propene) is much softer with a lower melting point.
It is formed as a waste product during the manufacture of isotactic poly
(propene) and its uses are limited. It is used, for example, in road paint,
in making roofing materials like "roofing felt", and in some sealants and
adhesives.
Syndiotactic poly(propene)
Syndiotactic poly(propene) is a relatively new material and is another
regularly arranged version of poly(propene). In this case, every alternate
CH3 group is orientated in the same way.

This regularity means that the chains can pack closely, and van der
Waals attractions will be fairly strong. However, the attractions aren't as
strong as in isotactic poly(propene). This makes syndiotactic poly
(propene) softer and gives it a lower melting point.
Because syndiotactic poly(propene) is relatively new, at the time of
writing uses were still being developed. It has uses in packaging - for
example, in plastic film for shrink wrapping food. There are also medical
uses - for example, in medical tubing and for medical bags and pouches.
There are a wide range of other potential uses - either on its own, or in
mixtures with isotactic poly(propene).
http://www.chemguide.co.uk/organicprops/alkenes/polymerisation.html (7 of 12)30/12/2004 11:22:27

polymerisation of alkenes

Poly(chloroethene) (polyvinyl chloride): PVC


Poly(chloroethene) is commonly known by the initials of its old name,
PVC.
Structure
Poly(chloroethene) is made by polymerising chloroethene, CH2=CHCl.
Working out its structure is no different from working out the structure of
poly(propene) (see above). As long as you draw the chloroethene
molecule in the right way, the structure is pretty obvious.

The equation is usually written:

It doesn't matter which carbon you attach the chlorine to in the original
molecule. Just be consistent on both sides of the equation.
The polymerisation process produces mainly atactic polymer molecules with the chlorines orientated randomly along the chain. The structure is
no different from atactic poly(propene) - just replace the CH3 groups by
chlorine atoms.

http://www.chemguide.co.uk/organicprops/alkenes/polymerisation.html (8 of 12)30/12/2004 11:22:27

polymerisation of alkenes

Because of the way the chlorine atoms stick out from the chain at
random, and because ot their large size, it is difficult for the chains to lie
close together. Poly(chloroethene) is mainly amorphous with only small
areas of crystallinity.
Properties and uses
You normally expect amorphous polymers to be more flexible than
crystalline ones because the forces of attraction between the chains tend
to be weaker. However, pure poly(chloroethene) tends to be rather hard
and rigid.
This is because of the presence of additional dipole-dipole interactions
due to the polarity of the carbon-chlorine bonds. Chlorine is more
electronegative than carbon, and so attracts the electrons in the bond
towards itself. That makes the chlorine atoms slightly negative and the
carbons slightly positive.
These permanent dipoles add to the attractions due to the temporary
dipoles which produce the dispersion forces.

Note: If you aren't happy about the various sorts of


intermolecular forces, it is important to follow this link.
If you don't understand about electronegativity and polar
bonds, then follow this one as well.
Use the BACK button (or HISTORY file or GO menu) on your
browser to return to this page.

http://www.chemguide.co.uk/organicprops/alkenes/polymerisation.html (9 of 12)30/12/2004 11:22:27

polymerisation of alkenes

Plasticisers are added to the poly(chloroethene) to reduce the


effectiveness of these attractions and make the plastic more flexible. The
more plasticiser you add, the more flexible it becomes.
Poly(chloroethene) is used to make a wide range of things including
guttering, plastic windows, electrical cable insulation, sheet materials for
flooring and other uses, footwear, clothing, and so on and so on.

Poly(tetrafluoroethene): PTFE
You may have come across this under the brand names of Teflon or
Fluon.
Structure
Structurally, PTFE is just like poly(ethene) except that each hydrogen in
the structure is replaced by a fluorine atom.

The PTFE chains tend to pack well and PTFE is fairly crystalline.
Because of the fluorine atoms, the chains also contain more electrons
(for an equal length) than a corresponding poly(ethene) chain. Taken
together (the good packing and the extra electrons) that means that the
van der Waals dispersion forces will be stronger than in even high
density poly(ethene).

http://www.chemguide.co.uk/organicprops/alkenes/polymerisation.html (10 of 12)30/12/2004 11:22:27

polymerisation of alkenes

Note: There won't be any dipole-dipole attractions between


the chains in addition to dispersion forces (unlike PVC). The
fluorines are arranged regularly around the carbon backbone.
Although each bond is very polar, overall they cancel each
other out.
(In fact, there might even be some repulsions because the
outsides of the chains all consist of slightly negative fluorine
atoms, but these obviously aren't enough to have a
significant effect on the strong dispersion forces. I haven't
been able to find any mention of this either in textbooks or on
the Web.)
You might like to read about CCl4 (a simple example of a nonpolar molecule containing polar bonds) on the main page
about electronegativity (different from the link above).
Use the BACK button on your browser to return to this page.

Properties and uses


PTFE has a relatively high melting point (due to the strength of the
attractions between the chains) and is very resistant to chemical attack.
The carbon chain is so wrapped up in fluorine atoms that nothing can get
at it to react with it. This makes it useful in the chemical and food
industries to coat vessels and make them resistant to almost everything
which might otherwise corrode them.
Equally important is that PTFE has remarkable non-stick properties which is the basis for its most familiar uses in non-stick kitchen and
garden utensils. For the same reason, it can also be used in things like
low-friction bearings

http://www.chemguide.co.uk/organicprops/alkenes/polymerisation.html (11 of 12)30/12/2004 11:22:27

polymerisation of alkenes

Note: Despite an extensive Web search, I haven't found any


convincing explanation for the non-stick properties of PTFE
at the molecular level. Many sources talk about it in terms of
surface tension or surface energy, which actually beg the
question. Unless you can explain the origin of these in terms
of attractions or repulsions at the molecular level, it seems to
me that you aren't actually explaining anything just by using a
posh-sounding term!
If you know of an explanation for the non-stick properties of
PTFE in terms of intermolecular forces that is simple enough
to use at this level, I would be grateful if you could contact me
using the address on the about this site page.

Where would you like to go now?


To the alkenes menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/alkenes/polymerisation.html (12 of 12)30/12/2004 11:22:27

Aryl halides Menu

Understanding Chemistry

ARYL HALIDES (HALOGENOARENES) MENU

Aryl halides are compounds like chlorobenzene - in which a halogen


atom is attached directly to a benzene ring. This covers the limited
amount of aryl halide chemistry required by UK A level syllabuses.

Background . . .
An introduction to the aryl halides and their physical properties.
The preparation of aryl halides . . .
Making chlorobenzene and bromobenzene by reacting chlorine or
bromine with benzene, and making iodobenzene from
benzenediazonium chloride.
The reactions of aryl halides . . .
. . . or more particularly the reasons for their lack of reaction
relative to halogenoalkanes (haloalkanes or alkyl halides).

Go to menu of other organic compounds . . .


Go to Main Menu . . .

http://www.chemguide.co.uk/organicprops/arylhalidesmenu.html (1 of 2)30/12/2004 11:22:28

Aryl halides Menu

You might also be interested in:


halogenoalkanes (haloalkanes or alkyl halides) . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/arylhalidesmenu.html (2 of 2)30/12/2004 11:22:28

an introduction to aryl halides (halogenoarenes)

INTRODUCTION TO THE ARYL HALIDES


(HALOGENOARENES)

This page looks at the structure and physical properties of three simple
aryl halides - chlorobenzene, bromobenzene and iodobenzene. An aryl
halide has a halogen atom attached directly to a benzene ring.

The structure of chlorobenzene


We'll look in some detail at the structure of chlorobenzene.
Bromobenzene and iodobenzene are just the same.
The simplest way to draw the structure of chlorobenzene is:

. . . but to understand chlorobenzene properly, you need to dig a bit


deeper than this.

Warning! You need to understand about the bonding in


benzene in order to make sense of this next bit.
If you follow this link, you may have to explore several other
pages before you are ready to come back here again. Use
the BACK button (or HISTORY file or GO menu) on your
browser to return to this page.

http://www.chemguide.co.uk/organicprops/arylhalides/background.html (1 of 4)30/12/2004 11:22:32

an introduction to aryl halides (halogenoarenes)

There is an interaction between the delocalised electrons in the benzene


ring and one of the lone pairs on the chlorine atom. This overlaps with
the delocalised ring electron system . . .

. . . giving a structure rather like this:

This delocalisation is by no means complete, but it does have a


significant effect on the properties of both the carbon-chlorine bond and
the polarity of the molecule.
The delocalisation introduces some extra bonding between the carbon
and the chlorine, making the bond stronger. This has a major effect on
the reactions of compounds like chlorobenzene.
There is also some movement of electrons away from the chlorine
towards the ring. Chlorine is quite electronegative and usually draws
electrons in the carbon-chlorine bond towards itself. In this case, this is
offset to some extent by the movement of electrons back towards the
ring in the delocalisation. The molecule is less polar than you would
otherwise have expected.

http://www.chemguide.co.uk/organicprops/arylhalides/background.html (2 of 4)30/12/2004 11:22:32

an introduction to aryl halides (halogenoarenes)

Physical properties
Boiling points
Chlorobenzene, bromobenzene and iodobenzene are all oily liquids. The
boiling points increase as the halogen atom gets bigger.
boiling point (C)
C6H5Cl

132

C6H5Br

156

C6H5I

189

Note: If you aren't happy about intermolecular forces then


you really ought to follow this link before you go on. This
section won't make much sense to you if you aren't familiar
with the various sorts of intermolecular forces.
Use the BACK button on your browser to return to this page.

The main attractions between the molecules will be van der Waals
dispersion forces. These increase as the number of electrons in the
molecule increases. This is the reason that the boiling points increase as
the halogen atom gets bigger.
There will also be permanent dipole-dipole attractions involved in the
chlorobenzene and bromobenzene, but very little in the iodobenzene.
Iodine has much the same electronegativity as carbon.
These dipole-dipole attractions must be very unimportant relative to the
dispersion forces because the most polar molecule (the chlorobenzene)
has the lowest boiling point of the three.

http://www.chemguide.co.uk/organicprops/arylhalides/background.html (3 of 4)30/12/2004 11:22:32

an introduction to aryl halides (halogenoarenes)

Solubility in water
The aryl halides are insoluble in water. They are denser than water and
form a separate lower layer.
The molecules are quite large compared with a water molecule. In order
for chlorobenzene to dissolve it would have to break lots of existing
hydrogen bonds between water molecules. You also have to break the
quite strong van der Waals dispersion forces between chlorobenzene
molecules. Both of these cost energy.
The only new forces between the chlorobenzene and the water would be
van der Waals dispersion forces. These aren't as strong as hydrogen
bonds (or the original dispersion forces in the chlorobenzene), and so
you wouldn't get much energy released when they form.
It simply isn't energetically profitable for chlorobenzene (and the others)
to dissolve in water.

Where would you like to go now?


To the aryl halides menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/arylhalides/background.html (4 of 4)30/12/2004 11:22:32

preparation of chlorobenzene, bromobenzene and iodobenzene

MAKING ARYL HALIDES (HALOGENOARENES)

This page only looks at the ways of making the aryl halides,
chlorobenzene, bromobenzene and iodobenzene, as required by some
UK A level syllabuses. It is not intended to be an overall survey of the
topic.

Making chlorobenzene
Benzene reacts with chlorine in the presence of a catalyst, replacing one
of the hydrogen atoms on the ring by a chlorine atom.
The reaction happens at room temperature. The catalyst is either
aluminium chloride or iron.
Strictly speaking iron isn't a catalyst, because it gets permanently
changed during the reaction. It reacts with some of the chlorine to form
iron(III) chloride, FeCl3.

This compound acts as the catalyst and behaves exactly like aluminium
chloride, AlCl3, in this reaction.
The reaction between benzene and chlorine in the presence of either
aluminium chloride or iron gives chlorobenzene.

or, written more compactly:

http://www.chemguide.co.uk/organicprops/arylhalides/preparation.html (1 of 4)30/12/2004 11:22:36

preparation of chlorobenzene, bromobenzene and iodobenzene

Note: Follow this link if you want the mechanism for the
chlorination of benzene.
Use the BACK button on your browser to return to this page
later.

Making bromobenzene
The reaction between benzene and bromine in the presence of either
aluminium bromide (rather than aluminium chloride) or iron gives
bromobenzene. Iron is usually used because it is cheaper and more
readily available. If you use iron, it is first converted into iron(III) bromide
by the reaction between the iron and bromine.

or:

Note: The mechanism for this reaction is exactly the same


as the chlorine reaction. Use the link further up the page if
you are interested.

http://www.chemguide.co.uk/organicprops/arylhalides/preparation.html (2 of 4)30/12/2004 11:22:36

preparation of chlorobenzene, bromobenzene and iodobenzene

Making iodobenzene
Iodobenzene can be made from the reaction of benzene with iodine if
they are heated under reflux in the presence of concentrated nitric acid,
but it is normally made from benzenediazonium chloride solution. That's
what we will concentrate on here.
If you add cold potassium iodide solution to ice-cold benzenediazonium
chloride solution, nitrogen gas is given off, and you get oily droplets of
iodobenzene formed.
There is a simple reaction between the diazonium ions present in the
benzenediazonium chloride solution and the iodide ions from the
potassium iodide solution.

Note: If you want to find out more about diazonium


compounds (including their preparation and their reactions),
follow this link to explore the phenylamine menu.

Where would you like to go now?


To the aryl halides menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004


http://www.chemguide.co.uk/organicprops/arylhalides/preparation.html (3 of 4)30/12/2004 11:22:36

preparation of chlorobenzene, bromobenzene and iodobenzene

http://www.chemguide.co.uk/organicprops/arylhalides/preparation.html (4 of 4)30/12/2004 11:22:36

reactions of aryl halides (halogenoarenes)

THE REACTIONS OF ARYL HALIDES


(HALOGENOARENES)

This page only looks at one aspect of the chemistry of the aryl halides
such as chlorobenzene - the fact that they are very unreactive compared
with halogenoalkanes (haloalkanes or alkyl halides). This is the only bit
of their chemistry asked for by any UK A level syllabuses.

Nucleophilic substitution reactions


A nucleophile can be either a negative ion or a molecule which carries a
partial negative charge somewhere on it. On this page, we are just going
to be looking at a simple nucleophile - a hydroxide ion.
A nucleophilic substitution reaction is one in which a part of a molecule is
replaced after it has been attacked by a nucleophile.

Note: To make sense of this page, you will have to


understand how nucleophilic substitution happens. Follow
this link to the introductory page on nucleophilic substitution
reactions if you aren't absolutely confident about this.
It would also be worthwhile looking at the page specifically
about nucleophilic substitution by hydroxide ions.
Both of these pages deal with nucleophilic substitution in the
halogenoalkanes. The rest of this page is going to be a
comparison with these reactions, and so it is important that
you understand them.
Use the BACK button (or HISTORY file or GO menu) on your
browser to return to this page later.

http://www.chemguide.co.uk/organicprops/arylhalides/reactions.html (1 of 4)30/12/2004 11:22:40

reactions of aryl halides (halogenoarenes)

Nucleophilic substitution in the halogenoalkanes


Here is a quick summary of the two ways that halogenoalkanes can react
with hydroxide ions. We'll compare these with the aryl halides afterwards.
The two different ways in which these reactions can happen depends on
what kind of halogenoalkane you are talking about.
Here is the mechanism for the reaction involving bromoethane - a
primary halogenoalkane. A hydroxide ion attacks the slightly positive
carbon atom and pushes off the bromine as a bromide ion.

A tertiary halogenoalkane reacts differently.


The mechanism this time involves an initial ionisation of the
halogenoalkane:

. . . followed by a very rapid attack by the hydroxide ion on the


carbocation (carbonium ion) formed:

http://www.chemguide.co.uk/organicprops/arylhalides/reactions.html (2 of 4)30/12/2004 11:22:40

reactions of aryl halides (halogenoarenes)

Note: If this isn't all completely obvious to you, you should


have followed (or spent more time on) the links above!

Nucleophilic substitution in the aryl halides


Simple aryl halides like chlorobenzene are very resistant to nucleophilic
substitution. It is possible to replace the chlorine by -OH, but only under
very severe industrial conditions - for example at 200C and 200
atmospheres.
In the lab, these reactions don't happen. There are two reasons for this depending on which of the above mechanisms you are talking about.

The extra strength of the carbon-halogen bond in aryl halides


The carbon-chlorine bond in chlorobenzene is stronger than you might
expect. There is an interaction between one of the lone pairs on the
chlorine atom and the delocalised ring electrons, and this strengthens the
bond.

Note: This is explained in more detail (with some diagrams)


on the introduction to aryl halides page.
Use the BACK button on your browser to return to this page if
you follow this link.

http://www.chemguide.co.uk/organicprops/arylhalides/reactions.html (3 of 4)30/12/2004 11:22:40

reactions of aryl halides (halogenoarenes)

Both of the mechanisms above involve breaking the carbon-halogen


bond at some stage. The more difficult it is to break, the slower the
reaction will be.

Repulsion by the ring electrons


This will only apply if the hydroxide ion attacked the chlorobenzene by a
mechanism like the first one described above. In that mechanism, the
hydroxide ion attacks the slightly positive carbon that the halogen atom is
attached to.
If the halogen atom is attached to a benzene ring, the incoming
hydroxide ion is going to be faced with the delocalised ring electrons
above and below that carbon atom. The negative hydroxide ion will
simply be repelled.
That particular mechanism is simply a non-starter!

Where would you like to go now?


To the aryl halides menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/arylhalides/reactions.html (4 of 4)30/12/2004 11:22:40

oxidation of alcohols

OXIDATION OF ALCOHOLS

This page looks at the oxidation of alcohols using acidified sodium or


potassium dichromate(VI) solution. This reaction is used to make
aldehydes, ketones and carboxylic acids, and as a way of distinguishing
between primary, secondary and tertiary alcohols.

Important! It is pointless reading this page unless you are


confident you know what primary, secondary and tertiary
alcohols are. If you aren't sure, you must read the
introduction to alcohols before you go on.
This page will also refer constantly to aldehydes and ketones.
Follow this link if you haven't come across these compounds
before.
Use the BACK button on your browser to return to this page.

Oxidising the different types of alcohols


The oxidising agent used in these reactions is normally a solution of
sodium or potassium dichromate(VI) acidified with dilute sulphuric acid. If
oxidation occurs, the orange solution containing the dichromate(VI) ions
is reduced to a green solution containing chromium(III) ions.
The electron-half-equation for this reaction is

http://www.chemguide.co.uk/organicprops/alcohols/oxidation.html (1 of 9)30/12/2004 11:22:55

oxidation of alcohols

Note: If you don't yet know about electron-half-equations just


ignore this reference for now. If you should already know
about them, but aren't very good at handling them, you might
like to have a look at this link. It isn't particularly important for
the purposes of the current page.
If you choose to follow this link, use the BACK button on your
browser to return to this page.

Primary alcohols
Primary alcohols can be oxidised to either aldehydes or carboxylic acids
depending on the reaction conditions. In the case of the formation of
carboxylic acids, the alcohol is first oxidised to an aldehyde which is then
oxidised further to the acid.
Partial oxidation to aldehydes
You get an aldehyde if you use an excess of the alcohol, and distil off the
aldehyde as soon as it forms.
The excess of the alcohol means that there isn't enough oxidising agent
present to carry out the second stage. Removing the aldehyde as soon
as it is formed means that it doesn't hang around waiting to be oxidised
anyway!
If you used ethanol as a typical primary alcohol, you would produce the
aldehyde ethanal, CH3CHO.
The full equation for this reaction is fairly complicated, and you need to
understand about electron-half-equations in order to work it out.

In organic chemistry, simplified versions are often used which


concentrate on what is happening to the organic substances. To do that,
oxygen from an oxidising agent is represented as [O]. That would
produce the much simpler equation:
http://www.chemguide.co.uk/organicprops/alcohols/oxidation.html (2 of 9)30/12/2004 11:22:55

oxidation of alcohols

It also helps in remembering what happens. You can draw simple


structures to show the relationship between the primary alcohol and the
aldehyde formed.

Important! This is not intended to suggest any sort of


mechanism for the reaction - it is just a way of helping you to
remember what happens.
If you are in the UK A level system, it is highly likely that your
examiners will accept equations involving [O]. To be sure,
consult your syllabus, past papers and mark schemes. If you
haven't got any of these things, follow this link to find out how
to get them.

Full oxidation to carboxylic acids


You need to use an excess of the oxidising agent and make sure that the
aldehyde formed as the half-way product stays in the mixture.
The alcohol is heated under reflux with an excess of the oxidising agent.
When the reaction is complete, the carboxylic acid is distilled off.
The full equation for the oxidation of ethanol to ethanoic acid is:

http://www.chemguide.co.uk/organicprops/alcohols/oxidation.html (3 of 9)30/12/2004 11:22:55

oxidation of alcohols

Note: This equation is worked out in detail on the page


about electron-half-equations mentioned above, if you are
interested.
If you choose to follow this link, use the BACK button on your
browser to return to this page.

The more usual simplified version looks like this:

Alternatively, you could write separate equations for the two stages of the
reaction - the formation of ethanal and then its subsequent oxidation.

This is what is happening in the second stage:

Secondary alcohols
Secondary alcohols are oxidised to ketones - and that's it. For example, if
you heat the secondary alcohol propan-2-ol with sodium or potassium
dichromate(VI) solution acidified with dilute sulphuric acid, you get
propanone formed.
Playing around with the reaction conditions makes no difference
http://www.chemguide.co.uk/organicprops/alcohols/oxidation.html (4 of 9)30/12/2004 11:22:55

oxidation of alcohols

whatsoever to the product.


Using the simple version of the equation and showing the relationship
between the structures:

If you look back at the second stage of the primary alcohol reaction, you
will see that an oxygen "slotted in" between the carbon and the hydrogen
in the aldehyde group to produce the carboxylic acid. In this case, there
is no such hydrogen - and the reaction has nowhere further to go.

Tertiary alcohols
Tertiary alcohols aren't oxidised by acidified sodium or potassium
dichromate(VI) solution. There is no reaction whatsoever.
If you look at what is happening with primary and secondary alcohols,
you will see that the oxidising agent is removing the hydrogen from the OH group, and a hydrogen from the carbon atom attached to the -OH.
Tertiary alcohols don't have a hydrogen atom attached to that carbon.
You need to be able to remove those two particular hydrogen atoms in
order to set up the carbon-oxygen double bond.

http://www.chemguide.co.uk/organicprops/alcohols/oxidation.html (5 of 9)30/12/2004 11:22:55

oxidation of alcohols

Using these reactions as a test for the different types of


alcohol
Doing the test
First you have to be sure that you have actually got an alcohol by testing
for the -OH group. You would need to show that it was a neutral liquid,
free of water and that it reacted with solid phosphorus(V) chloride to
produce a burst of acidic steamy hydrogen chloride fumes.

Note: You will find the chemistry of the phosphorus(V)


chloride test by following this link.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/alcohols/oxidation.html (6 of 9)30/12/2004 11:22:55

oxidation of alcohols

You would then add a few drops of the alcohol to a test tube containing
potassium dichromate(VI) solution acidified with dilute sulphuric acid.
The tube would be warmed in a hot water bath.
Results for the various kinds of alcohol
Picking out the tertiary alcohol
In the case of a primary or secondary alcohol, the orange solution turns
green. With a tertiary alcohol there is no colour change.
After heating:

Distinguishing between the primary and secondary alcohols


You need to produce enough of the aldehyde (from oxidation of a
primary alcohol) or ketone (from a secondary alcohol) to be able to test
them. There are various things which aldehydes do which ketones don't.
These include the reactions with Tollens' reagent, Fehling's solution and
Benedict's solution, and are covered on a separate page.

Note: You will find these tests for aldehydes by following this
link.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/alcohols/oxidation.html (7 of 9)30/12/2004 11:22:55

oxidation of alcohols

In my experience, these tests can be a bit of a bother to carry out and the
results aren't always as clear-cut as the books say. A much simpler but
fairly reliable test is to use Schiff's reagent. Schiff's reagent isn't
specifically mentioned by any of the UK A level syllabuses, but I have
always used it.
Schiff's reagent is a fuchsin dye decolourised by passing sulphur dioxide
through it. In the presence of even small amounts of an aldehyde, it turns
bright magenta.
It must, however, be used absolutely cold, because ketones react with it
very slowly to give the same colour. If you heat it, obviously the change
is faster - and potentially confusing.
While you are warming the reaction mixture in the hot water bath, you
can pass any vapours produced through some Schiff's reagent.

If the Schiff's reagent quickly becomes magenta, then you are


producing an aldehyde from a primary alcohol.

http://www.chemguide.co.uk/organicprops/alcohols/oxidation.html (8 of 9)30/12/2004 11:22:55

oxidation of alcohols

If there is no colour change in the Schiff's reagent, or only a trace


of pink colour within a minute or so, then you aren't producing an
aldehyde, and so haven't got a primary alcohol.
Because of the colour change to the acidified potassium
dichromate(VI) solution, you must therefore have a secondary
alcohol.

You should check the result as soon as the potassium dichromate(VI)


solution turns green - if you leave it too long, the Schiff's reagent might
start to change colour in the secondary alcohol case as well.

Where would you like to go now?


To the alcohols menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/alcohols/oxidation.html (9 of 9)30/12/2004 11:22:55

oxidation of aldehydes and ketones

OXIDATION OF ALDEHYDES AND KETONES

This page looks at ways of distinguishing between aldehydes and


ketones using oxidising agents such as acidified potassium dichromate
(VI) solution, Tollens' reagent, Fehling's solution and Benedict's solution.

Background
Why do aldehydes and ketones behave differently?
You will remember that the difference between an aldehyde and a ketone
is the presence of a hydrogen atom attached to the carbon-oxygen
double bond in the aldehyde. Ketones don't have that hydrogen.

The presence of that hydrogen atom makes aldehydes very easy to


oxidise. Or, put another way, they are strong reducing agents.

http://www.chemguide.co.uk/organicprops/carbonyls/oxidation.html (1 of 10)30/12/2004 11:23:05

oxidation of aldehydes and ketones

Note: If you aren't sure about oxidation and reduction, it


would be a good idea to follow this link to another part of the
site before you go on. Alternatively, come back to this link if
you feel you need help later on in this page.
Use the BACK button (or HISTORY file or GO menu if you get
seriously waylaid) on your browser to return to this page.

Because ketones don't have that particular hydrogen atom, they are
resistant to oxidation. Only very strong oxidising agents like potassium
manganate(VII) solution (potassium permanganate solution) oxidise
ketones - and they do it in a destructive way, breaking carbon-carbon
bonds.
Provided you avoid using these powerful oxidising agents, you can easily
tell the difference between an aldehyde and a ketone. Aldehydes are
easily oxidised by all sorts of different oxidising agents: ketones aren't.
You will find details of these reactions further down the page.

What is formed when aldehydes are oxidised?


It depends on whether the reaction is done under acidic or alkaline
conditions. Under acidic conditions, the aldehyde is oxidised to a
carboxylic acid. Under alkaline conditions, this couldn't form because it
would react with the alkali. A salt is formed instead.

http://www.chemguide.co.uk/organicprops/carbonyls/oxidation.html (2 of 10)30/12/2004 11:23:05

oxidation of aldehydes and ketones

Note: In the case of methanal, HCHO, the oxidation goes


further. The methanoic acid or methanoate ions formed are
easily oxidised to carbon dioxide and water.

Building equations for the oxidation reactions


If you need to work out the equations for these reactions, the only reliable
way of building them is to use electron-half-equations.
The half-equation for the oxidation of the aldehyde obviously varies
depending on whether you are doing the reaction under acidic or alkaline
conditions.
Under acidic conditions it is:

. . . and under alkaline conditions:

http://www.chemguide.co.uk/organicprops/carbonyls/oxidation.html (3 of 10)30/12/2004 11:23:05

oxidation of aldehydes and ketones

Note: These electron-half-equations are quite easy to work


out from scratch. There isn't any real need to remember them.
Follow this link if you aren't sure how to do it, or if you aren't
confident about using these equations. These particular
examples aren't covered, but the technique is exactly the
same whatever the equation.
Because I am going to be using electron-half-equations quite
a lot on the rest of this page, it would definitely be worth
following this link if you aren't happy about them.
Use the BACK button on your browser to return to this page.

These half-equations are then combined with the half-equations from


whatever oxidising agent you are using. Examples are given in detail
below.

Specific examples
In each of the following examples, we are assuming that you know that
you have either an aldehyde or a ketone. There are lots of other things
which could also give positive results.

Note: Follow this link to find out how to test for the carbonoxygen double bond in aldehydes and ketones.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/carbonyls/oxidation.html (4 of 10)30/12/2004 11:23:05

oxidation of aldehydes and ketones

Assuming that you know it has to be one or the other, in each case, a
ketone does nothing. Only an aldehyde gives a positive result.

Using acidified potassium dichromate(VI) solution


A small amount of potassium dichromate(VI) solution is acidified with
dilute sulphuric acid and a few drops of the aldehyde or ketone are
added. If nothing happens in the cold, the mixture is warmed gently for a
couple of minutes - for example, in a beaker of hot water.

ketone

No change in the orange solution.

aldehyde

Orange solution turns green.

The orange dichromate(VI) ions have been reduced to green chromium


(III) ions by the aldehyde. In turn the aldehyde is oxidised to the
corresponding carboxylic acid.
The electron-half-equation for the reduction of dichromate(VI) ions is:

Combining that with the half-equation for the oxidation of an aldehyde


under acidic conditions:

. . . gives the overall equation:

http://www.chemguide.co.uk/organicprops/carbonyls/oxidation.html (5 of 10)30/12/2004 11:23:05

oxidation of aldehydes and ketones

Note: You may wonder why I have gone to all the trouble of
working out a complete equation for this reaction (and the
next ones) rather than using symbols like [O] which are
frequently used in organic chemistry.
The problem is that what is important in using these reactions
as tests is the colour change in the oxidising agent. In this
particular reaction, you have to explain, for example, why the
solution turns green. Any equation that you write has got to
show the production of the chromium(III) ions.
If you aren't sure exactly how the two half-equations are
combined to give the final equation, follow the last link (further
up the page) to learn how to do it - and then come back and
have a go yourself to check that you can get the right answer.

Using Tollens' reagent (the silver mirror test)


Tollens' reagent contains the diamminesilver(I) ion, [Ag(NH3)2]+.
This is made from silver(I) nitrate solution. You add a drop of sodium
hydroxide solution to give a precipitate of silver(I) oxide, and then add just
enough dilute ammonia solution to redissolve the precipitate.
To carry out the test, you add a few drops of the aldehyde or ketone to
the freshly prepared reagent, and warm gently in a hot water bath for a
few minutes.

ketone

No change in the colourless solution.

aldehyde

The colourless solution produces a grey precipitate of


silver, or a silver mirror on the test tube.

http://www.chemguide.co.uk/organicprops/carbonyls/oxidation.html (6 of 10)30/12/2004 11:23:05

oxidation of aldehydes and ketones

Aldehydes reduce the diamminesilver(I) ion to metallic silver. Because


the solution is alkaline, the aldehyde itself is oxidised to a salt of the
corresponding carboxylic acid.

Note: If you actually get a silver mirror it is very satisfying but a grey precipitate is enough to show that the test has
worked. Whether you get a silver mirror or not seems a matter
of luck. I have watched really careful students clean
everything scrupulously and take great care over quantities,
and still get no more than a trace of a mirror. On the other
hand students who have just thrown everything together in
the first grubby test tube that came to hand can get a
wonderful mirror. Life isn't always fair!

The electron-half-equation for the reduction of of the diamminesilver(I)


ions to silver is:

Combining that with the half-equation for the oxidation of an aldehyde


under alkaline conditions:

. . . gives the overall equation:

Using Fehling's solution or Benedict's solution


Fehling's solution and Benedict's solution are variants of essentially the
same thing. Both contain complexed copper(II) ions in an alkaline solution.
Fehling's solution contains copper(II) ions complexed with tartrate ions in
sodium hydroxide solution. Complexing the copper(II) ions with tartrate
ions prevents precipitation of copper(II) hydroxide.
http://www.chemguide.co.uk/organicprops/carbonyls/oxidation.html (7 of 10)30/12/2004 11:23:05

oxidation of aldehydes and ketones

Benedict's solution contains copper(II) ions complexed with citrate ions in


sodium carbonate solution. Again, complexing the copper(II) ions
prevents the formation of a precipitate - this time of copper(II) carbonate.
Both solutions are used in the same way. A few drops of the aldehyde or
ketone are added to the reagent, and the mixture is warmed gently in a
hot water bath for a few minutes.

ketone

No change in the blue solution.

aldehyde

The blue solution produces a dark red precipitate of


copper(I) oxide.

Aldehydes reduce the complexed copper(II) ion to copper(I) oxide.


Because the solution is alkaline, the aldehyde itself is oxidised to a salt of
the corresponding carboxylic acid.

Note: I have watched students do this reaction with


aldehydes and Fehling's solution over many years. Getting
the dark red precipitate described in all the books was
actually pretty rare! A lot of imagination had to go in to
spotting the red colour in amongst all the other colours you
tend to get as well. It wasn't one of my favourite tests. I have
never used Benedict's solution - perhaps that behaves better!
Methanal is such a powerful reducing agent that the copper(II)
ions may be reduced to metallic copper - often seen as a very
nice copper mirror on the tube.

http://www.chemguide.co.uk/organicprops/carbonyls/oxidation.html (8 of 10)30/12/2004 11:23:05

oxidation of aldehydes and ketones

The equations for these reactions are always simplified to avoid having to
write in the formulae for the tartrate or citrate ions in the copper
complexes. The electron-half-equations for both Fehling's solution and
Benedict's solution can be written as:

Combining that with the half-equation for the oxidation of an aldehyde


under alkaline conditions:

. . . gives the overall equation:

Note: If you are really alert, you might wonder where the
hydroxide ions come from if you are talking about Benedict's
solution, which is made alkaline using sodium carbonate.
Sodium carbonate solution is alkaline because the carbonate
ions react reversibly with water to produce hydroxide ions
(and hydrogencarbonate ions).

Where would you like to go now?


To the aldehydes and ketones menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

http://www.chemguide.co.uk/organicprops/carbonyls/oxidation.html (9 of 10)30/12/2004 11:23:05

oxidation of aldehydes and ketones

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/carbonyls/oxidation.html (10 of 10)30/12/2004 11:23:05

addition-elimination reactions of aldehydes and ketones

ADDITION-ELIMINATION REACTIONS OF
ALDEHYDES AND KETONES

This page looks at the reaction of aldehydes and ketones with 2,4dinitrophenylhydrazine (Brady's reagent) as a test for the carbon-oxygen
double bond. It also looks briefly at some other similar reactions which
are all known as addition-elimination (or condensation) reactions.

The reaction with 2,4-dinitrophenylhydrazine


2,4-dinitrophenylhydrazine is often abbreviated to 2,4-DNP or 2,4-DNPH.
A solution of 2,4-dinitrophenylhydrazine in a mixture of methanol and
sulphuric acid is known as Brady's reagent.
What is 2,4-dinitrophenylhydrazine?
Although the name sounds complicated, and the structure looks quite
complicated, it is actually very easy to work out.
Start with the formula of hydrazine. That's almost all you need to
remember!
Hydrazine is:

In phenylhydrazine, one of the hydrogens is replaced by a phenyl group,


C6H5. This is based on a benzene ring.

http://www.chemguide.co.uk/organicprops/carbonyls/addelim.html (1 of 7)30/12/2004 11:23:12

addition-elimination reactions of aldehydes and ketones

Note: I'm using the common symbol for a benzene ring. If


you aren't familiar with this, there is a carbon atom at each
corner of the hexagon, together with a hydrogen atom if there
isn't anything else attached. The circle in the middle of the
hexagon suggests the delocalised electrons.
If you want to read more about the structure of benzene, you
could follow this link to another part of the site. It isn't
important for the purposes of the present page, and you
might have to explore quite a lot of other pages to help you to
understand that one.
If you do choose to follow this link, use the BACK button (or
HISTORY file or GO menu if you get seriously waylaid) on
your browser to return to this page.

In 2,4-dinitrophenylhydrazine, there are two nitro groups, NO2, attached


to the phenyl group in the 2- and 4- positions. The corner with the
nitrogen attached is counted as the number 1 position, and you just
number clockwise around the ring.

Doing the reaction


Details vary slightly depending on the nature of the aldehyde or ketone,
http://www.chemguide.co.uk/organicprops/carbonyls/addelim.html (2 of 7)30/12/2004 11:23:12

addition-elimination reactions of aldehydes and ketones

and the solvent that the 2,4-dinitrophenylhydrazine is dissolved in.


Assuming you are using Brady's reagent (a solution of the 2,4dinitrophenylhydrazine in methanol and sulphuric acid):
Add either a few drops of the aldehyde or ketone, or possibly a solution
of the aldehyde or ketone in methanol, to the Brady's reagent. A bright
orange or yellow precipitate shows the presence of the carbon-oxygen
double bond in an aldehyde or ketone.
This is the simplest test for an aldehyde or ketone.

The chemistry of the reaction


The overall reaction is given by the equation:

R and R' can be any combination of hydrogen or hydrocarbon groups


(such as alkyl groups). If at least one of them is a hydrogen, then the
original compound is an aldehyde. If both are hydrocarbon groups, then it
is a ketone.
Look carefully at what has happened.

http://www.chemguide.co.uk/organicprops/carbonyls/addelim.html (3 of 7)30/12/2004 11:23:12

addition-elimination reactions of aldehydes and ketones

Provided you take care to draw the two starting molecules lined up right,
working out the structure of the product is easy.
The product is known as a "2,4-dinitrophenylhydrazone". Notice that all
that has changed is the ending from "-ine" to "-one". That's possibly
confusing!
The product from the reaction with ethanal would be called ethanal 2,4dinitrophenylhydrazone; from propanone, you would get propanone 2,4dinitrophenylhydrazone - and so on. That's not too difficult!
The reaction is known as a condensation reaction. A condensation
reaction is one in which two molecules join together with the loss of a
small molecule in the process. In this case, that small molecule is water.
In terms of mechanisms, this is a nucleophilic addition-elimination
reaction. The 2,4-dinitrophenylhydrazine first adds across the carbonoxygen double bond (the addition stage) to give an intermediate
compound which then loses a molecule of water (the elimination stage).

Note: This mechanism isn't required by any of the UK A


level syllabuses, and so you won't find it anywhere on this
site. Sorry!

http://www.chemguide.co.uk/organicprops/carbonyls/addelim.html (4 of 7)30/12/2004 11:23:12

addition-elimination reactions of aldehydes and ketones

Using the reaction


The reaction has two uses in testing for aldehydes and ketones.

First, you can just use it to test for the presence of the carbonoxygen double bond. You only get an orange or yellow precipitate
from a carbon-oxygen double bond in an aldehyde or ketone.
Secondly, you can use it to help to identify the specific aldehyde or
ketone.
The precipitate is filtered and washed with, for example, methanol
and then recrystallised from a suitable solvent which will vary
depending on the nature of the aldehyde or ketone. For example,
you can recrystallise the products from the small aldehydes and
ketones from a mixture of ethanol and water.
The crystals are dissolved in the minimum quantity of hot solvent.
When the solution cools, the crystals are re-precipitated and can
be filtered, washed with a small amount of solvent and dried. They
should then be pure.
If you then find the melting point of the crystals, you can compare
it with tables of the melting points of 2,4-dinitrophenylhydrazones
of all the common aldehydes and ketones to find out which one
you are likely to have got.

Some other similar reactions


If you go back and look at the equations, nothing in the 2,4dinitrophenylhydrazine changes during the reaction apart from the -NH2
group. You can get a similar reaction if the -NH2 group is attached to
other things.
In each case, the reaction would look like this:

http://www.chemguide.co.uk/organicprops/carbonyls/addelim.html (5 of 7)30/12/2004 11:23:12

addition-elimination reactions of aldehydes and ketones

In what follows, all that changes is the nature of the "X".

with hydrazine

The product is a "hydrazone". If you started from propanone, it would be


propanone hydrazone.

with phenylhydrazine

The product is a "phenylhydrazone".

with hydroxylamine

The product is an "oxime" - for example, ethanal oxime.

http://www.chemguide.co.uk/organicprops/carbonyls/addelim.html (6 of 7)30/12/2004 11:23:12

addition-elimination reactions of aldehydes and ketones

Where would you like to go now?


To the aldehydes and ketones menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/carbonyls/addelim.html (7 of 7)30/12/2004 11:23:12

esterification - alcohols and carboxylic acids

ESTERIFICATION

This page looks at esterification - mainly the reaction between alcohols


and carboxylic acids to make esters. It also looks briefly at making esters
from the reactions between acyl chlorides (acid chlorides) and alcohols,
and between acid anhydrides and alcohols.

What are esters?


Esters are derived from carboxylic acids. A carboxylic acid contains the COOH group, and in an ester the hydrogen in this group is replaced by a
hydrocarbon group of some kind. We shall just be looking at cases where
it is replaced by an alkyl group, but it could equally well be an aryl group
(one based on a benzene ring).
A common ester - ethyl ethanoate
The most commonly discussed ester is ethyl ethanoate. In this case, the
hydrogen in the -COOH group has been replaced by an ethyl group. The
formula for ethyl ethanoate is:

Notice that the ester is named the opposite way around from the way the
formula is written. The "ethanoate" bit comes from ethanoic acid. The
"ethyl" bit comes from the ethyl group on the end.

http://www.chemguide.co.uk/organicprops/alcohols/esterification.html (1 of 8)30/12/2004 11:23:20

esterification - alcohols and carboxylic acids

Note: In my experience, students starting organic chemistry


get more confused about writing names and formulae for
esters than for almost anything else - particularly when it
comes to less frequently met esters like the ones coming up
next. Take time and care to make sure you understand!

A few more esters


In each case, be sure that you can see how the names and formulae
relate to each other.

Notice that the acid is named by counting up the total number of carbon
atoms in the chain - including the one in the -COOH group. So, for
example, CH3CH2COOH is propanoic acid, and CH3CH2COO is the
propanoate group.

http://www.chemguide.co.uk/organicprops/alcohols/esterification.html (2 of 8)30/12/2004 11:23:20

esterification - alcohols and carboxylic acids

Note: You can find more about naming acids and esters by
following this link to a different part of this site.
Use the BACK button on your browser to return to this page.

Making esters from carboxylic acids and alcohols


The chemistry of the reaction
Esters are produced when carboxylic acids are heated with alcohols in the
presence of an acid catalyst. The catalyst is usually concentrated
sulphuric acid. Dry hydrogen chloride gas is used in some cases, but
these tend to involve aromatic esters (ones containing a benzene ring). If
you are a UK A level student you won't have to worry about these.
The esterification reaction is both slow and reversible. The equation for
the reaction between an acid RCOOH and an alcohol R'OH (where R and
R' can be the same or different) is:

So, for example, if you were making ethyl ethanoate from ethanoic acid
and ethanol, the equation would be:

http://www.chemguide.co.uk/organicprops/alcohols/esterification.html (3 of 8)30/12/2004 11:23:20

esterification - alcohols and carboxylic acids

Note: The mechanism for the esterification reaction is


covered in the catalysis section of this site. It is not required
for any UK A level chemistry syllabus.
If you follow this link, use the BACK button on your browser to
return to this page.

Doing the reactions


On a test tube scale
Carboxylic acids and alcohols are often warmed together in the presence
of a few drops of concentrated sulphuric acid in order to observe the smell
of the esters formed.
You would normally use small quantities of everything heated in a test
tube stood in a hot water bath for a couple of minutes.
Because the reactions are slow and reversible, you don't get a lot of ester
produced in this time. The smell is often masked or distorted by the smell
of the carboxylic acid. A simple way of detecting the smell of the ester is
to pour the mixture into some water in a small beaker.
Apart from the very small ones, esters are fairly insoluble in water and
tend to form a thin layer on the surface. Excess acid and alcohol both
dissolve and are tucked safely away under the ester layer.
Small esters like ethyl ethanoate smell like typical organic solvents (ethyl
ethanoate is a common solvent in, for example, glues).
As the esters get bigger, the smells tend towards artificial fruit flavouring "pear drops", for example.
On a larger scale

http://www.chemguide.co.uk/organicprops/alcohols/esterification.html (4 of 8)30/12/2004 11:23:20

esterification - alcohols and carboxylic acids

If you want to make a reasonably large sample of an ester, the method


used depends to some extent on the size of the ester. Small esters are
formed faster than bigger ones.
To make a small ester like ethyl ethanoate, you can gently heat a mixture
of ethanoic acid and ethanol in the presence of concentrated sulphuric
acid, and distil off the ester as soon as it is formed.
This prevents the reverse reaction happening. It works well because the
ester has the lowest boiling point of anything present. The ester is the only
thing in the mixture which doesn't form hydrogen bonds, and so it has the
weakest intermolecular forces.

Note: Follow this link if you aren't sure about hydrogen


bonding.
Use the BACK button on your browser to return to this page.

Larger esters tend to form more slowly. In these cases, it may be


necessary to heat the reaction mixture under reflux for some time to
produce an equilibrium mixture. The ester can be separated from the
carboxylic acid, alcohol, water and sulphuric acid in the mixture by
fractional distillation.

Note: Providing full details for organic preparations (including


all the steps necessary in cleaning up the product) is beyond
the scope of this site. If you need this sort of detail, you should
be looking at an organic practical book.

http://www.chemguide.co.uk/organicprops/alcohols/esterification.html (5 of 8)30/12/2004 11:23:20

esterification - alcohols and carboxylic acids

Other ways of making esters


Esters can also be made from the reactions between alcohols and either
acyl chlorides or acid anhydrides.

Making esters from alcohols and acyl chlorides (acid chlorides)


If you add an acyl chloride to an alcohol, you get a vigorous (even violent)
reaction at room temperature producing an ester and clouds of steamy
acidic fumes of hydrogen chloride.
For example, if you add the liquid ethanoyl chloride to ethanol, you get a
burst of hydrogen chloride produced together with the liquid ester ethyl
ethanoate.

Note: If you want to find out more about acyl chlorides,


explore the acyl chlorides menu by following this link.
Use the BACK button (or GO menu or HISTORY file) on your
browser to return to this page.

http://www.chemguide.co.uk/organicprops/alcohols/esterification.html (6 of 8)30/12/2004 11:23:20

esterification - alcohols and carboxylic acids

Making esters from alcohols and acid anhydrides


The reactions of acid anhydrides are slower than the corresponding
reactions with acyl chlorides, and you usually need to warm the mixture.
Taking ethanol reacting with ethanoic anhydride as a typical reaction
involving an alcohol:
There is a slow reaction at room temperature (or faster on warming).
There is no visible change in the colourless liquids, but a mixture of ethyl
ethanoate and ethanoic acid is formed.

Note: If you want to find out more about acid anhydrides,


explore the acid anhydrides menu by following this link.
Use the BACK button (or GO menu or HISTORY file) on your
browser to return to this page.

Where would you like to go now?


To the alcohols menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

http://www.chemguide.co.uk/organicprops/alcohols/esterification.html (7 of 8)30/12/2004 11:23:20

esterification - alcohols and carboxylic acids

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/alcohols/esterification.html (8 of 8)30/12/2004 11:23:20

mechanism for the esterification reaction

THE MECHANISM FOR THE ESTERIFICATION


REACTION

This page looks in detail at the mechanism for the formation of esters
from carboxylic acids and alcohols in the presence of concentrated
sulphuric acid acting as the catalyst. It uses the formation of ethyl
ethanoate from ethanoic acid and ethanol as a typical example.

The mechanism for the formation of ethyl ethanoate


A reminder of the facts
Ethanoic acid reacts with ethanol in the presence of concentrated
sulphuric acid as a catalyst to produce the ester, ethyl ethanoate. The
reaction is slow and reversible. To reduce the chances of the reverse
reaction happening, the ester is distilled off as soon as it is formed.

The mechanism

Warning! This is a fairly complex mechanism, and is


definitley NOT required for any UK A' level syllabus. I have
included it in case it is of use to my many non-A'level visitors.

http://www.chemguide.co.uk/physical/catalysis/esterify.html (1 of 7)30/12/2004 11:23:29

mechanism for the esterification reaction

All the steps in the mechanism below are shown as one-way reactions
because it makes the mechanism look less confusing. The reverse
reaction is actually done sufficiently differently that it affects the way the
mechanism is written. You will find a link to the hydrolysis of esters
further down the page if you are interested.

Note: The explanation assumes that you know about the use
of curly arrows in organic reaction mechanisms. If you aren't
happy about these follow this link before you go any further.
(To be honest, if you are that unsure about the conventions
used in reaction mechanisms, you probably shouldn't be
reading this page anyway - you will find it distinctly scary!)
Use the BACK button on your browser to return to this page.

Step 1
In the first step, the ethanoic acid takes a proton (a hydrogen ion) from
the concentrated sulphuric acid. The proton becomes attached to one of
the lone pairs on the oxygen which is double-bonded to the carbon.

The transfer of the proton to the oxygen gives it a positive charge, but it is
actually misleading to draw the structure in this way (although nearly
everybody does!).
The positive charge is delocalised over the whole of the right-hand end of
the ion, with a fair amount of positiveness on the carbon atom. In other
words, you can think of an electron pair shifting to give this structure:

http://www.chemguide.co.uk/physical/catalysis/esterify.html (2 of 7)30/12/2004 11:23:29

mechanism for the esterification reaction

You could also imagine another electron pair shift producing a third
structure:

So which of these is the correct structure of the ion formed? None of


them! The truth lies somewhere in between all of them. One way of
writing the delocalised structure of the ion is like this:

The double headed arrows are telling you that each of the individual
structures makes a contribution to the real structure of the ion. They
don't mean that the bonds are flipping back and forth between one
structure and another. The various structures are known as resonance
structures or canonical forms.
There will be some degree of positive charge on both of the oxygen
atoms, and also on the carbon atom. Each of the bonds between the
carbon and the two oxygens will be the same - somewhere between a
single bond and a double bond.

http://www.chemguide.co.uk/physical/catalysis/esterify.html (3 of 7)30/12/2004 11:23:29

mechanism for the esterification reaction

Note: You will find a more pictorial look at a similar case to


this in a page discussing the acidity of organic acids.
Amongst other things, that page looks at the structure of ions
like the ethanoate ions which also have delocalised charges.
Use the BACK button on your browser to return easily to this
page.

For the purposes of the rest of this discussion, we are going to use the
structure where the positive charge is on the carbon atom.
Step 2
The positive charge on the carbon atom is attacked by one of the lone
pairs on the oxygen of the ethanol molecule.

Note: You could work out precisely why that particular


oxygen carries the positive charge on the right-hand side. On
the other hand, you could realise that there has to be a
positive charge somewhere (because you started with one),
and that particular oxygen doesn't look right - it has too many
bonds. Put the charge on there!
That's a quick rough-and-ready reasoning which works every
time I use it!

http://www.chemguide.co.uk/physical/catalysis/esterify.html (4 of 7)30/12/2004 11:23:29

mechanism for the esterification reaction

Step 3
What happens next is that a proton (a hydrogen ion) gets transferred
from the bottom oxygen atom to one of the others. It gets picked off by
one of the other substances in the mixture (for example, by attaching to a
lone pair on an unreacted ethanol molecule), and then dumped back onto
one of the oxygens more or less at random.
The net effect is:

Step 4
Now a molecule of water is lost from the ion.

The product ion has been drawn in a shape to reflect the product which
we are finally getting quite close to!
The structure for the latest ion is just like the one we discusssed at length
back in step 1. The positive charge is actually delocalised all over that
end of the ion, and there will also be contributions from structures where
the charge is on the either of the oxygens:

It is easier to follow what is happening if we keep going with the structure


with the charge on the carbon.
http://www.chemguide.co.uk/physical/catalysis/esterify.html (5 of 7)30/12/2004 11:23:29

mechanism for the esterification reaction

Step 5
The hydrogen is removed from the oxygen by reaction with the
hydrogensulphate ion which was formed way back in the first step.

And there we are! The ester has been formed, and the sulphuric acid
catalyst has been regenerated.

Where would you like to go now?


To explore the reverse of this reaction . . .
To the catalysis menu . . .
To the Physical Chemistry menu . . .
To the organic mechanisms menu . . .
To Main Menu . . .

http://www.chemguide.co.uk/physical/catalysis/esterify.html (6 of 7)30/12/2004 11:23:29

mechanism for the esterification reaction

You might also be interested in:


properties and reactions of alcohols . . .
Covers all the physical and chemical properties of alcohols required
by UK A level syllabuses.
properties and reactions of carboxylic acids . . .
A similar survey of carboxylic acid chemistry.
properties and reactions of esters . . .
. . . and of esters.

Jim Clark 2002 (modified 2004)

http://www.chemguide.co.uk/physical/catalysis/esterify.html (7 of 7)30/12/2004 11:23:29

mechanism for the acid catalysed hydrolysis of esters

THE MECHANISM FOR THE ACID CATALYSED


HYDROLYSIS OF ESTERS

This page looks in detail at the mechanism for the hydrolysis of esters in
the presence of a dilute acid (such as hydrochloric acid or sulphuric acid)
acting as the catalyst. It uses ethyl ethanoate as a typical ester.

The mechanism for the hydrolysis of ethyl ethanoate


A reminder of the facts
Ethyl ethanoate is heated under reflux with a dilute acid such as dilute
hydrochloric acid or dilute sulphuric acid. The ester reacts with the water
present to produce ethanoic acid and ethanol.
Because the reaction is reversible, an equilibrium mixture is produced
containing all four of the substances in the equation. In order to get as
much hydrolysis as possible, a large excess of water can be used. The
dilute acid provides both the acid catalyst and the water.

The mechanism

Warning! This is a fairly complex mechanism, and is


definitley NOT required for any UK A' level syllabus. I have
included it in case it is of use to my many non-A'level visitors.

http://www.chemguide.co.uk/physical/catalysis/hydrolyse.html (1 of 7)30/12/2004 11:23:50

mechanism for the acid catalysed hydrolysis of esters

All the steps in the mechanism below are shown as one-way reactions
because it makes the mechanism look less confusing. The reverse
reaction is actually done sufficiently differently that it affects the way the
mechanism is written. You will find a link to the esterification reaction
further down the page if you are interested.

Note: The explanation assumes that you know about the use
of curly arrows in organic reaction mechanisms. If you aren't
happy about these follow this link before you go any further.
(To be honest, if you are that unsure about the conventions
used in reaction mechanisms, you probably shouldn't be
reading this page anyway - you will find it distinctly scary!)
Use the BACK button on your browser to return to this page.

Step 1
The actual catalyst in this case is the hydroxonium ion, H3O+, present in
all solutions of acids in water.
In the first step, the ester takes a proton (a hydrogen ion) from the
hydroxonium ion. The proton becomes attached to one of the lone pairs
on the oxygen which is double-bonded to the carbon.

The transfer of the proton to the oxygen gives it a positive charge, but the
charge is actually delocalised (spread around) much more widely than
this shows.
One way of representing this delocalisation is to draw a number of
structures called resonance structures or canonical forms joined by
http://www.chemguide.co.uk/physical/catalysis/hydrolyse.html (2 of 7)30/12/2004 11:23:50

mechanism for the acid catalysed hydrolysis of esters

double-headed arrows. You could, if you wished, put in some curly


arrows to show the movements of electrons which change one of these
structures into the next.

None of these formulae represents the true structure of the ion - but each
gives you some information about it. For example, notice where the
positive charge is in the three structures. What this means in reality is
that the positive charge is spread around over those three atoms - the
two oxygens and the carbon.

Note: If you haven't come across canonical forms as a way


of representing delocalisation, don't worry about it particularly.
It is important, however, that you don't imagine that the
molecule is rapidly flipping from one structure to another. The
double-headed arrows mean something different.
A mule is a hybrid of a donkey and a horse. In this notation,
you could represent a mule by writing donkey and horse
connected by a double-headed arrow. Neitherdonkey nor
horse accurately represents what a mule looks like, but with a
bit of imagination you could build up a fairly good picture of a
mule by combining together the characteristics of both
donkey and horse. But a mule obviously doesn't spend its
time rapidly changing back and forth between being a donkey
and a horse!

http://www.chemguide.co.uk/physical/catalysis/hydrolyse.html (3 of 7)30/12/2004 11:23:50

mechanism for the acid catalysed hydrolysis of esters

The next stage of the mechanism involves an attack on the carbon, and
so it is convenient to use the structure showing the positive charge on
that carbon in the next step.
Step 2
The positive charge on the carbon atom is attacked by one of the lone
pairs on the oxygen of a water molecule.

Note: You could work out precisely why that particular


oxygen carries the positive charge on the right-hand side. On
the other hand, you could realise that there has to be a
positive charge somewhere (because you started with one),
and that particular oxygen doesn't look right - it has too many
bonds. Put the charge on there!
That's a quick rough-and-ready reasoning which works every
time I use it!

http://www.chemguide.co.uk/physical/catalysis/hydrolyse.html (4 of 7)30/12/2004 11:23:50

mechanism for the acid catalysed hydrolysis of esters

Step 3
What happens next is that a proton (a hydrogen ion) gets transferred
from the bottom oxygen atom to one of the others. It gets picked off by
one of the other substances in the mixture (for example, by attaching to a
lone pair on a water molecule), and then dumped back onto one of the
oxygens more or less at random.
Eventually, by chance, it will join to the oxygen with the ethyl group
attached. When that happens, the net effect is:

Step 4
Now a molecule of ethanol is lost from the ion. That's one of the products
of the reaction.

The structure for the latest ion is just like the one we discusssed at length
back in step 1. The positive charge is actually delocalised all over that
end of the ion. The real structure will be a hybrid of these:

http://www.chemguide.co.uk/physical/catalysis/hydrolyse.html (5 of 7)30/12/2004 11:23:50

mechanism for the acid catalysed hydrolysis of esters

It is easier to follow what is happening if we keep going with the structure


with the charge on the carbon.
Step 5
The hydrogen is removed from the oxygen by reaction with a water
molecule.

And there we are! We have produced the ethanoic acid (the other product
of the reaction) and the hydroxonium ion catalyst has been regenerated.

Where would you like to go now?


To explore the reverse of this reaction . . .
To the catalysis menu . . .
To the Physical Chemistry menu . . .
To the organic mechanisms menu . . .
To Main Menu . . .

http://www.chemguide.co.uk/physical/catalysis/hydrolyse.html (6 of 7)30/12/2004 11:23:50

mechanism for the acid catalysed hydrolysis of esters

You might also be interested in:


properties and reactions of alcohols . . .
Covers all the physical and chemical properties of alcohols required
by UK A level syllabuses.
properties and reactions of carboxylic acids . . .
A similar survey of carboxylic acid chemistry.
properties and reactions of esters . . .
. . . and of esters.

Jim Clark 2002 (modified 2004)

http://www.chemguide.co.uk/physical/catalysis/hydrolyse.html (7 of 7)30/12/2004 11:23:50

Carboxylic Acids Menu

Understanding Chemistry

CARBOXYLIC ACIDS MENU

Background . . .
An introduction to carboxylic acids and their salts, including their
bonding and their physical properties.
Preparation of carboxylic acids . . .
Their preparation by the oxidation of primary alcohols and
aldehydes, and by the hydrolysis of nitriles.
Simple reactions of carboxylic acids as acids . . .
Reactions with metals, sodium hydroxide, carbonates and
hydrogencarbonates, ammonia and amines.
Making esters from carboxylic acids . . .
Reactions with alcohols to produce esters.
Reduction of carboxylic acids . . .
The reduction of carboxylic acids to primary alcohols using lithium
tetrahydridoaluminate(III) (lithium aluminium hydride).
Making acyl chlorides (acid chlorides) . . .
Replacing the -OH in the -COOH group with chlorine using
phosphorus(V) chloride, phosphorus(III) chloride or sulphur
dichloride oxide (thionyl chloride).

http://www.chemguide.co.uk/organicprops/acidmenu.html (1 of 2)30/12/2004 11:23:51

Carboxylic Acids Menu

The decarboxylation of carboxylic acids and their salts . . .


Formation of hydrocarbons by heating the salts of carboxylic acids
(and some acids) with soda lime.

Go to menu of other organic compounds . . .


Go to Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/acidmenu.html (2 of 2)30/12/2004 11:23:51

an introduction to carboxylic acids

INTRODUCING CARBOXYLIC ACIDS

This page explains what carboxylic acids are, and looks at the ions that
they form in their salts. It also considers their simple physical properties
such as solubility and boiling points. Details of the chemical reactions of
carboxylic acids are described on separate pages.

What are carboxylic acids?


Carboxylic acids contain a -COOH group
Carboxylic acids are compounds which contain a -COOH group. For the
purposes of this page we shall just look at compounds where the -COOH
group is attached either to a hydrogen atom or to an alkyl group.

Note: There is no very significant reason for this. For


simplicity, I am just trying to avoid making it look complicated
by having either another active group present in the molecule
as well as the -COOH, or the presence of a benzene ring.
Benzoic acid (benzenecarboxylic acid) has the -COOH group
attached to a benzene ring. Its physical and chemical
properties are in line with those of any other carboxylic acid
of a similar size, so I haven't felt it necessary to write about it
separately.
If you are interested in amino acids, you could follow this link
to the amino acids and proteins menu.

http://www.chemguide.co.uk/organicprops/acids/background.html (1 of 6)30/12/2004 11:23:56

an introduction to carboxylic acids

Examples of carboxylic acids

The name counts the total number of carbon atoms in the longest chain including the one in the -COOH group. If you have side groups attached
to the chain, notice that you always count from the carbon atom in the COOH group as being number 1.

Note: If you aren't confident about naming organic


compounds, then you might like to follow this link at some
point.
Use the BACK button on your browser if you want to return to
this page.

Salts of carboxylic acids


Carboxylic acids are acidic because of the hydrogen in the -COOH
group. When the acids form salts, this is lost and replaced by a metal.
Sodium ethanoate, for example, has the structure:

Depending on whether or not you wanted to stress the ionic nature of the
compound, this would be simplified to CH3COO- Na+ or just CH3COONa.
http://www.chemguide.co.uk/organicprops/acids/background.html (2 of 6)30/12/2004 11:23:56

an introduction to carboxylic acids

Notice:

The bond between the sodium and the ethanoate is ionic. Don't
draw a line between the two (implying a covalent bond). That's
absolutely wrong!
Although the name is written with the sodium first, the formula is
always written in one of the ways shown. This is something you
just have to get used to.

Note: We often write the formula of the ion showing the


negative charge on one of the oxygen atoms (as above). This
is OK for many purposes, but is technically wrong. In fact the
negative charge is delocalised over the whole of the -COO
end of the ion and the two carbon-oxygen bonds are identical
- not one single and one double.
This is discussed in more detail elsewhere on the site about
half-way down a page about the acidity of organic
compounds, although you would probably have to refer to
other pages as well to understand this properly.
This isn't essential for the purposes of the present page, but if
you choose to follow this link use the BACK button (or
HISTORY file or GO menu) on your browser to return to this
page.

http://www.chemguide.co.uk/organicprops/acids/background.html (3 of 6)30/12/2004 11:23:56

an introduction to carboxylic acids

Physical properties of carboxylic acids


The physical properties (for example, boiling point and solubility) of the
carboxylic acids are governed by their ability to form hydrogen bonds.
Boiling points
Before we look at carboxylic acids, a reminder about alcohols:
The boiling points of alcohols are higher than those of alkanes of similar
size because the alcohols can form hydrogen bonds with each other as
well as van der Waals dispersion forces and dipole-dipole interactions.

Note: Hydrogen bonding in alcohols is discussed in detail in


the introduction to alcohols. If you aren't confident about
hydrogen bonding and other intermolecular forces and their
relationship to physical properties it would be a good idea to
read this before you go on. It is all done in more detail than I
have used on this present page.
Use the BACK button on your browser to return to this page.

The boiling points of carboxylic acids of similar size are higher still.
For example:

propan-1-ol

CH3CH2CH2OH

97.2C

ethanoic acid

CH3COOH

118C

These are chosen for comparison because they have identical relative
molecular masses and almost the same number of electrons (which

http://www.chemguide.co.uk/organicprops/acids/background.html (4 of 6)30/12/2004 11:23:56

an introduction to carboxylic acids

affects van der Waals dispersion forces).


The higher boiling points of the carboxylic acids are still caused by
hydrogen bonding, but operating in a different way.
In a pure carboxylic acid, hydrogen bonding can occur between two
molecules of acid to produce a dimer.

This immediately doubles the size of the molecule and so increases the
van der Waals dispersion forces between one of these dimers and its
neighbours - resulting in a high boiling point.

Solubility in water
In the presence of water, the carboxylic acids don't dimerise. Instead,
hydrogen bonds are formed between water molecules and individual
molecules of acid.
The carboxylic acids with up to four carbon atoms will mix with water in
any proportion. When you mix the two together, the energy released
when the new hydrogen bonds form is much the same as is needed to
break the hydrogen bonds in the pure liquids.
The solubility of the bigger acids decreases very rapidly with size. This is
because the longer hydrocarbon "tails" of the molecules get between
water molecules and break hydrogen bonds. In this case, these broken
hydrogen bonds are only replaced by much weaker van der Waals
dispersion forces.

http://www.chemguide.co.uk/organicprops/acids/background.html (5 of 6)30/12/2004 11:23:56

an introduction to carboxylic acids

Note: The similar case with the solubility of alcohols is


discussed in detail in the introduction to alcohols. If you aren't
happy about the effect of chain length on solubility then it
would definitely be worth following this link.
Use the BACK button on your browser to return to this page.

The energetics of dissolving carboxylic acids in water is made more


complicated because some of the acid molecules actually react with the
water rather than just dissolving in it. This is the basis for the acidity of
these compounds and is discussed on another page.

Where would you like to go now?


To the carboxylic acids menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/acids/background.html (6 of 6)30/12/2004 11:23:56

amino acids and proteins menu

Understanding Chemistry

AMINO ACIDS AND PROTEINS MENU

Background . . .
An introduction to amino acids including their physical properties.
Acid-base reactions of amino acids . . .
Amino acids as zwitterions and the influence of this on their
reactions with acids and bases.
The structure of proteins . . .
A brief introduction to protein structure including primary,
secondary and tertiary structure.
Hydrolysis of proteins . . .
Hydrolysing proteins using hydrochloric acid.

Go to menu of other organic compounds . . .


Go to Main Menu . . .

http://www.chemguide.co.uk/organicprops/aminoacidmenu.html (1 of 2)30/12/2004 11:23:57

amino acids and proteins menu

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/aminoacidmenu.html (2 of 2)30/12/2004 11:23:57

an introduction to amino acids

INTRODUCING AMINO ACIDS

This page explains what amino acids are, concentrating on the 2-amino
acids that are biologically important. It looks in some detail at their simple
physical properties such as solubility and melting points.

What are amino acids?


Structures and names
Amino acids are exactly what they say they are! They are compounds
containing an amino group, -NH2, and a carboxylic acid group, -COOH.
The biologically important amino acids have the amino group attached to
the carbon atom next door to the -COOH group. They are known as 2amino acids. They are also known (slightly confusingly) as alpha-amino
acids. These are the ones we will concentrated on.
The two simplest of these amino acids are 2-aminoethanoic acid and 2aminopropanoic acid.
Because of the biological importance of molecules like these, they are
normally known by their traditional biochemical names.
2-aminoethanoic acid, for example, is usually called glycine, and 2aminopropanoic acid is usually known as alanine.

http://www.chemguide.co.uk/organicprops/aminoacids/background.html (1 of 7)30/12/2004 11:24:03

an introduction to amino acids

The general formula for a 2-amino acid is:

. . . where "R" can be quite a complicated group containing other active


groups like -OH, -SH, other amine or carboxylic acid groups, and so on.
It is definitely NOT necessarily a simple hydrocarbon group!

Physical properties
Melting points
The amino acids are crystalline solids with surprisingly high melting
points. It is difficult to pin the melting points down exactly because the
amino acids tend to decompose before they melt. Decomposition and
melting tend to be in the 200 - 300C range.
For the size of the molecules, this is very high. Something unusual must
be happening.
If you look again at the general structure of an amino acid, you will see
that it has both a basic amine group and an acidic carboxylic acid group.

There is an internal transfer of a hydrogen ion from the -COOH group to


the -NH2 group to leave an ion with both a negative charge and a
positive charge.
This is called a zwitterion.

http://www.chemguide.co.uk/organicprops/aminoacids/background.html (2 of 7)30/12/2004 11:24:03

an introduction to amino acids

This is the form that amino acids exist in even in the solid state. Instead
of the weaker hydrogen bonds and other intermolecular forces that you
might have expected, you actually have much stronger ionic attractions
between one ion and its neighbours.
These ionic attractions take more energy to break and so the amino
acids have high melting points for the size of the molecules.

Solubility
Amino acids are generally soluble in water and insoluble in non-polar
organic solvents such as hydrocarbons.
This again reflects the presence of the zwitterions. In water, the ionic
attractions between the ions in the solid amino acid are replaced by
strong attractions between polar water molecules and the zwitterions.
This is much the same as any other ionic substance dissolving in water.
The extent of the solubility in water varies depending on the size and
nature of the "R" group.

Note: At this point I would normally try to relate the actual


values for solubility of the various amino acids to their
structures. Unfortunately, from the solubility values that I
have got (and I'm not convinced they are necessarily right), I
can't find any obvious patterns - in fact, there are some very
strange cases indeed.

http://www.chemguide.co.uk/organicprops/aminoacids/background.html (3 of 7)30/12/2004 11:24:03

an introduction to amino acids

The lack of solubility in non-polar organic solvents such as hydrocarbons


is because of the lack of attraction between the solvent molecules and
the zwitterions. Without strong attractions between solvent and amino
acid, there won't be enough energy released to pull the ionic lattice apart.

Optical activity
If you look yet again at the general formula for an amino acid, you will
see that (apart from glycine, 2-aminoethanoic acid) the carbon at the
centre of the structure has four different groups attached. In glycine, the
"R" group is another hydrogen atom.

This is equally true if you draw the structure of the zwitterion instead of
this simpler structure.
Because of these four different groups attached to the same carbon
atom, amino acids (apart from glycine) are chiral.

Important: If you don't know exactly what that means, follow


this link to the page about optical isomerism. You will find the
optical activity of amino acids discussed at the bottom of that
page, but read the whole page to be sure that you
understand what is going on.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/aminoacids/background.html (4 of 7)30/12/2004 11:24:03

an introduction to amino acids

The lack of a plane of symmetry means that there will be two


stereoisomers of an amino acid (apart from glycine) - one the nonsuperimposable mirror image of the other.
For a general 2-amino acid, the isomers are:

Note: If you don't know what the various bond symbols


mean, you shouldn't have got this far! Follow the link
mentioned above to the page about optical isomerism. Read
that page and follow the further link on that page to drawing
organic molecules.
Use the BACK button on your browser to return to this page.

All the naturally occurring amino acids have the right-hand structure in
this diagram. This is known as the "L-" configuration. The other one is
known as the "D-" configuration.
You almost certainly don't need to know this for UK A level chemistry
purposes, but if you are interested, you can recognise the Lconfiguration by imagining that you are looking down from above on the
right-hand structure in the last diagram. If you read around the other
groups in a clockwise direction, you get the word CORN.

http://www.chemguide.co.uk/organicprops/aminoacids/background.html (5 of 7)30/12/2004 11:24:03

an introduction to amino acids

You can't tell by looking at a structure whether that isomer will rotate the
plane of polarisation of plane polarised light clockwise or anticlockwise.
All the naturally occurring amino acids have the same L- configuration,
but they include examples which rotate the plane clockwise (+) and those
which do the opposite (-).
For example:

(+)alanine

(-)cysteine

(-)tyrosine

(+)valine

It is quite common for natural systems to only work with one of the optical
isomers (enantiomers) of an optically active substance like the amino
acids. It isn't too difficult to see why that might be. Because the
molecules have different spatial arrangements of their various groups,
only one of them is likely to fit properly into the active sites on the
enzymes they work with.

Where would you like to go now?


To the amino acid and proteins menu . . .
To the menu of other organic compounds . . .
http://www.chemguide.co.uk/organicprops/aminoacids/background.html (6 of 7)30/12/2004 11:24:03

an introduction to amino acids

To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/aminoacids/background.html (7 of 7)30/12/2004 11:24:03

the acid base behaviour of amino acids

THE ACID-BASE BEHAVIOUR OF AMINO ACIDS

This page looks at what happens to amino acids as you change the pH
by adding either acids or alkalis to their solutions.
For simplicity, the page only looks at amino acids which contain a single NH2 group and a single -COOH group.

Amino acids as zwitterions


Zwitterions in simple amino acid solutions
An amino acid has both a basic amine group and an acidic carboxylic
acid group.

There is an internal transfer of a hydrogen ion from the -COOH group to


the -NH2 group to leave an ion with both a negative charge and a positive
charge.
This is called a zwitterion.

http://www.chemguide.co.uk/organicprops/aminoacids/acidbase.html (1 of 7)30/12/2004 11:24:09

the acid base behaviour of amino acids

This is the form that amino acids exist in even in the solid state. If you
dissolve the amino acid in water, a simple solution also contains this ion.

Adding an alkali to an amino acid solution


If you increase the pH of a solution of an amino acid by adding hydroxide
ions, the hydrogen ion is removed from the -NH3+ group.

You could show that the amino acid now existed as a negative ion using
electrophoresis.
In its simplest form, electrophoresis can just consist of a piece of
moistened filter paper on a microscope slide with a crocodile clip at each
end attached to a battery. A drop of amino acid solution is placed in the
centre of the paper.
Although the amino acid solution is colourless, its position after a time
can be found by spraying it with a solution of ninhydrin. If the paper is
allowed to dry and then heated gently, the amino acid shows up as a
coloured spot.
The amino acid would be found to travel towards the anode (the positive
electrode).

Adding an acid to an amino acid solution


If you decrease the pH by adding an acid to a solution of an amino acid,
the -COO- part of the zwitterion picks up a hydrogen ion.
http://www.chemguide.co.uk/organicprops/aminoacids/acidbase.html (2 of 7)30/12/2004 11:24:09

the acid base behaviour of amino acids

This time, during electrophoresis, the amino acid would move towards the
cathode (the negative electrode).

Shifting the pH from one extreme to the other


Suppose you start with the ion we've just produced under acidic
conditions and slowly add alkali to it.
That ion contains two acidic hydrogens - the one in the -COOH group and
the one in the -NH3+ group.
The more acidic of these is the one in the -COOH group, and so that is
removed first - and you get back to the zwitterion.

So when you have added just the right amount of alkali, the amino acid
no longer has a net positive or negative charge. That means that it
wouldn't move towards either the cathode or anode during
electrophoresis.
The pH at which this lack of movement during electrophoresis happens is
known as the isoelectric point of the amino acid. This pH varies from
amino acid to amino acid.
If you go on adding hydroxide ions, you will get the reaction we've
already seen, in which a hydrogen ion is removed from the -NH3+ group.

http://www.chemguide.co.uk/organicprops/aminoacids/acidbase.html (3 of 7)30/12/2004 11:24:09

the acid base behaviour of amino acids

Note: You might have expected the isoelectric point to be at


pH 7 - when the solution is neither acidic nor alkaline. In fact,
the isoelectric point for many amino acids is about pH 6.
Explaining why it isn't at pH 7 is quite long-winded and almost
certainly beyond what you will need for UK A level purposes.
If you are interested, the problem is discussed at the bottom
of this page.

You can, of course, reverse the whole process by adding an acid to the
ion we've just finished up with.
That ion contains two basic groups - the -NH2 group and the -COOgroup. The -NH2 group is the stronger base, and so picks up hydrogen
ions first. That leads you back to the zwitterion again.

. . . and, of course, you can keep going by then adding a hydrogen ion to
the -COO- group.

Why isn't the isoelectric point of an amino acid at pH 7?

http://www.chemguide.co.uk/organicprops/aminoacids/acidbase.html (4 of 7)30/12/2004 11:24:09

the acid base behaviour of amino acids

Warning: If you are a UK A level student, you are extremely


unlikely to need this for exam purposes. It is included only for
interest - you can probably safely ignore it. Check your
syllabus and past papers. If you haven't got these, follow this
link to find out how to get hold of them.

When an amino acid dissolves in water, the situation is a little bit more
complicated than we tend to pretend at this level. The zwitterion interacts
with water molecules - acting as both an acid and a base.
As an acid:

The -NH3+ group is a weak acid and donates a hydrogen ion to a water
molecule. Because it is only a weak acid, the position of equilibrium will
lie to the left.
As a base:

The -COO- group is a weak base and takes a hydrogen ion from a water
molecule. Again, the equilibrium lies to the left.
When you dissolve an amino acid in water, both of these reactions are
happening.
But . . .
The positions of the two equilibria aren't identical - they vary depending
on the influence of the "R" group. In practice, for the simple amino acids
we have been talking about, the position of the first equilibrium lies a bit
further to the right than the second one.

http://www.chemguide.co.uk/organicprops/aminoacids/acidbase.html (5 of 7)30/12/2004 11:24:09

the acid base behaviour of amino acids

That means that there will be rather more of the negative ion from the
amino acid in the solution than the positive one.
In those circumstances, if you carried out electrophoresis on the
unmodified solution, there would be a slight drift of amino acid towards
the positive electrode (the anode).
To stop that, you need to cut down the amount of the negative ion so that
the concentrations of the two ions are identical. You can do that by
adding a very small amount of acid to the solution, moving the position of
the first equilibrium further to the left.
Typically, the pH has to be lowered to about 6 to achieve this. For
glycine, for example, the isoelectric point is pH 6.07; for alanine, 6.11;
and for serine, 5.68.

Note: I need to stress again that these figures (and the


arguments) only hold where there is only one -NH2 group and
one -COOH group in the amino acid. The isoelectric points
are quite different if the molecule contains a second -NH2 or COOH group.

Where would you like to go now?


To the amino acid and proteins menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/aminoacids/acidbase.html (6 of 7)30/12/2004 11:24:09

the acid base behaviour of amino acids

http://www.chemguide.co.uk/organicprops/aminoacids/acidbase.html (7 of 7)30/12/2004 11:24:09

protein structure

THE STRUCTURE OF PROTEINS

This page explains how amino acids combine to make proteins and what
is meant by the primary, secondary and tertiary structures of proteins.
Quaternary structure isn't covered. It only applies to proteins consisting
of more than one polypeptide chain, and isn't required for UK A level
chemistry purposes.

The primary structure of proteins


Drawing the amino acids
In chemistry, if you were to draw the structure of a general 2-amino acid,
you would probably draw it like this:

However, for drawing the structures of proteins, we usually twist it so that


the "R" group sticks out at the side. It is much easier to see what is
happening if you do that.

That means that the two simplest amino acids, glycine and alanine,
would be shown as:

Peptides and polypeptides


http://www.chemguide.co.uk/organicprops/aminoacids/proteinstruct.html (1 of 14)30/12/2004 11:24:43

protein structure

Glycine and alanine can combine together with the elimination of a


molecule of water to produce a dipeptide. It is possible for this to
happen in one of two different ways - so you might get two different
dipeptides.
Either:

Or:

In each case, the linkage shown in blue in the structure of the dipeptide
is known as a peptide link. In chemistry, this would also be known as an
amide link, but since we are now in the realms of biochemistry and
biology, we'll use their terms.
If you joined three amino acids together, you would get a tripeptide. If
you joined lots and lots together (as in a protein chain), you get a
polypeptide.
A protein chain will have somewhere in the range of 50 to 2000 amino
acid residues. You have to use this term because strictly speaking a
peptide chain isn't made up of amino acids. When the amino acids
combine together, a water molecule is lost. The peptide chain is made up
from what is left after the water is lost - in other words, is made up of
amino acid residues.
By convention, when you are drawing peptide chains, the -NH2 group
which hasn't been converted into a peptide link is written at the left-hand
end. The unchanged -COOH group is written at the right-hand end.
http://www.chemguide.co.uk/organicprops/aminoacids/proteinstruct.html (2 of 14)30/12/2004 11:24:43

protein structure

A protein chain will therefore look like this:

The "R" groups come from the 20 amino acids which occur in proteins.
The peptide chain is known as the backbone, and the "R" groups are
known as side chains.

The primary structure of proteins


Now there's a problem! The term "primary structure" is used in two
different ways.
At its simplest, the term is used to describe the order of the amino acids
joined together to make the protein. In other words, if you replaced the
"R" groups in the last diagram by real groups you would have the primary
structure of a particular protein.
This primary structure is usually shown using abbreviations for the amino
acid residues. These abbreviations commonly consist of three letters or
one letter.
Using three letter abbreviations, a bit of a protein chain might be
represented by, for example:

If you look carefully, you will spot the abbreviations for glycine (Gly) and
alanine (Ala) amongst the others.
The wider definition of primary structure includes all the features of a
protein which are a result of covalent bonds. Obviously, all the peptide
http://www.chemguide.co.uk/organicprops/aminoacids/proteinstruct.html (3 of 14)30/12/2004 11:24:43

protein structure

links are made of covalent bonds, so that isn't a problem.


But there is an additional feature in proteins which is also covalently
bound. It involves the amino acid cysteine.

If two cysteine side chains end up next to each other because of folding
in the peptide chain, they can react to form a sulphur bridge. This is
another covalent link and so some people count it as a part of the
primary structure of the protein.

Because of the way sulphur bridges affect the way the protein folds,
other people count this as a part of the tertiary structure (see below).
This is obviously a potential source of confusion!

http://www.chemguide.co.uk/organicprops/aminoacids/proteinstruct.html (4 of 14)30/12/2004 11:24:43

protein structure

Important: You need to know where your particular


examiners are going to include sulphur bridges - as a part of
the primary structure or as a part of the tertiary structure. At
the time of writing, the UK A level Chemistry Exam Board
which wanted most detail on protein structure (OCR) seemed
to include it under tertiary structure. You need to check your
current syllabus and past papers. If you haven't got these,
follow this link to find out how to get hold of them.

The secondary structure of proteins


Within the long protein chains there are regions in which the chains are
organised into regular structures known as alpha-helices (alpha-helixes)
and beta-pleated sheets. These are the secondary structures in proteins.
These secondary structures are held together by hydrogen bonds. These
form as shown in the diagram between one of the lone pairs on an
oxygen atom and the hydrogen attached to a nitrogen atom:

http://www.chemguide.co.uk/organicprops/aminoacids/proteinstruct.html (5 of 14)30/12/2004 11:24:43

protein structure

Important: If you aren't happy about hydrogen bonding and


are unsure about what this diagram means, follow this link
before you go on. What follows is difficult enough to visualise
anyway without having to worry about what hydrogen bonds
are as well!
You must also find out exactly how much detail you need to
know about this next bit. It may well be that all you need is to
have heard of an alpha-helix and know that it is held together
by hydrogen bonds between the C=O and N-H groups. Once
again, you need to check your syllabus and past papers particularly mark schemes for the past papers.
If you follow either of these links, use the BACK button on
your browser to return to this page.

The alpha-helix
In an alpha-helix, the protein chain is coiled like a loosely-coiled spring.
The "alpha" means that if you look down the length of the spring, the
coiling is happening in a clockwise direction as it goes away from you.

Note: If your visual imagination is as hopeless as mine, the


only way to really understand this is to get a bit of wire and
coil it into a spring shape. The lead on your computer mouse
is fine for doing this!

http://www.chemguide.co.uk/organicprops/aminoacids/proteinstruct.html (6 of 14)30/12/2004 11:24:43

protein structure

The next diagram shows how the alpha-helix is held together by


hydrogen bonds. This is a very simplified diagram, missing out lots of
atoms. We'll talk it through in some detail after you have had a look at it.

What's wrong with the diagram? Two things:


First of all, only the atoms on the parts of the coils facing you are shown.
If you try to show all the atoms, the whole thing gets so complicated that
it is virtually impossible to understand what is going on.
Secondly, I have made no attempt whatsoever to get the bond angles
right. I have deliberately drawn all of the bonds in the backbone of the
chain as if they lie along the spiral. In truth they stick out all over the
place. Again, if you draw it properly it is virtually impossible to see the
spiral.
So, what do you need to notice?
Notice that all the "R" groups are sticking out sideways from the main
helix.
Notice the regular arrangement of the hydrogen bonds. All the N-H
groups are pointing upwards, and all the C=O groups pointing
downwards. Each of them is involved in a hydrogen bond.
And finally, although you can't see it from this incomplete diagram, each
http://www.chemguide.co.uk/organicprops/aminoacids/proteinstruct.html (7 of 14)30/12/2004 11:24:43

protein structure

complete turn of the spiral has 3.6 (approximately) amino acid residues
in it.
If you had a whole number of amino acid residues per turn, each group
would have an identical group underneath it on the turn below. Hydrogen
bonding can't happen under those circumstances.
Each turn has 3 complete amino acid residues and two atoms from the
next one. That means that each turn is offset from the ones above and
below, such that the N-H and C=O groups are brought into line with each
other.

Beta-pleated sheets
In a beta-pleated sheet, the chains are folded so that they lie alongside
each other. The next diagram shows what is known as an "anti-parallel"
sheet. All that means is that next-door chains are heading in opposite
directions. Given the way this particular folding happens, that would
seem to be inevitable.

It isn't, in fact, inevitable! It is possible to have some much more


complicated folding so that next-door chains are actually heading in the
same direction. We are getting well beyond the demands of UK A level
chemistry now.
The folded chains are again held together by hydrogen bonds involving
exactly the same groups as in the alpha-helix.

http://www.chemguide.co.uk/organicprops/aminoacids/proteinstruct.html (8 of 14)30/12/2004 11:24:43

protein structure

Note: Note that there is no reason why these sheets have to


be made from four bits of folded chain alongside each other
as shown in this diagram. That was an arbitrary choice which
produced a diagram which fitted nicely on the screen!

The tertiary structure of proteins


What is tertiary structure?
The tertiary structure of a protein is a description of the way the whole
chain (including the secondary structures) folds itself into its final 3dimensional shape. This is often simplified into models like the following
one for the enzyme dihydrofolate reductase. Enzymes are, of course,
based on proteins.

http://www.chemguide.co.uk/organicprops/aminoacids/proteinstruct.html (9 of 14)30/12/2004 11:24:43

protein structure

Note: This diagram was obtained from the RCSB Protein


Data Bank. If you want to find more information about
dihydrofolate reductase, their reference number for it is 7DFR.
There is nothing particularly special about this enzyme in
terms of structure. I chose it because it contained only a
single protein chain and had examples of both types of
secondary structure in it.

http://www.chemguide.co.uk/organicprops/aminoacids/proteinstruct.html (10 of 14)30/12/2004 11:24:43

protein structure

The model shows the alpha-helices in the secondary structure as coils of


"ribbon". The beta-pleated sheets are shown as flat bits of ribbon ending
in an arrow head. The bits of the protein chain which are just random
coils and loops are shown as bits of "string".
The colour coding in the model helps you to track your way around the
structure - going through the spectrum from dark blue to end up at red.
You will also notice that this particular model has two other molecules
locked into it (shown as ordinary molecular models). These are the two
molecules whose reaction this enzyme catalyses.

What holds a protein into its tertiary structure?


The tertiary structure of a protein is held together by interactions between
the the side chains - the "R" groups. There are several ways this can
happen.
Ionic interactions
Some amino acids (such as aspartic acid and glutamic acid) contain an
extra -COOH group. Some amino acids (such as lysine) contain an extra NH2 group.
You can get a transfer of a hydrogen ion from the -COOH to the -NH2
group to form zwitterions just as in simple amino acids.
You could obviously get an ionic bond between the negative and the
positive group if the chains folded in such a way that they were close to
each other.

http://www.chemguide.co.uk/organicprops/aminoacids/proteinstruct.html (11 of 14)30/12/2004 11:24:43

protein structure

Hydrogen bonds
Notice that we are now talking about hydrogen bonds between side
groups - not between groups actually in the backbone of the chain.
Lots of amino acids contain groups in the side chains which have a
hydrogen atom attached to either an oxygen or a nitrogen atom. This is a
classic situation where hydrogen bonding can occur.
For example, the amino acid serine contains an -OH group in the side
chain. You could have a hydrogen bond set up between two serine
residues in different parts of a folded chain.

http://www.chemguide.co.uk/organicprops/aminoacids/proteinstruct.html (12 of 14)30/12/2004 11:24:43

protein structure

You could easily imagine similar hydrogen bonding involving -OH groups,
or -COOH groups, or -CONH2 groups, or -NH2 groups in various
combinations - although you would have to be careful to remember that a
-COOH group and an -NH2 group would form a zwitterion and produce
stronger ionic bonding instead of hydrogen bonds.

van der Waals dispersion forces


Several amino acids have quite large hydrocarbon groups in their side
chains. A few examples are shown below. Temporary fluctuating dipoles
in one of these groups could induce opposite dipoles in another group on
a nearby folded chain.
The dispersion forces set up would be enough to hold the folded
structure together.

Important: If you aren't happy about van der Waals


dispersion forces you should follow this link.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/aminoacids/proteinstruct.html (13 of 14)30/12/2004 11:24:43

protein structure

Sulphur bridges
Sulphur bridges which form between two cysteine residues have already
been discussed under primary structures. Wherever you choose to place
them doesn't affect how they are formed!

Where would you like to go now?


To the amino acid and proteins menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/aminoacids/proteinstruct.html (14 of 14)30/12/2004 11:24:43

protein hydrolysis

THE HYDROLYSIS OF PROTEINS

This page looks briefly at the hydrolysis of proteins into their constituent
amino acids using hydrochloric acid.

Hydrolysing proteins using hydrochloric acid


The chemistry of the reaction
If you have already studied the hydrolysis of amides under acidic
conditions, you will find that this is basically the same reaction. That's not
surprising because what biologists and biochemists call a peptide link (in
proteins, for example) is what chemists call an amide link.
With an amide like ethanamide, the carbon-nitrogen bond in the amide
group is broken and you get a carboxylic acid formed:

Now imagine doing the same thing with a simple dipeptide made of any
two amino acids.

Instead of ammonium ions, you get positive ions made from the -NH2
http://www.chemguide.co.uk/organicprops/aminoacids/proteinhydrolysis.html (1 of 3)30/12/2004 11:24:48

protein hydrolysis

groups reacting with hydrogen ions.


You need the extra hydrogen ion in the equation (compared with the
amide equation) to react with the -NH2 group on the left-hand end of the
dipeptide - the one not involved in the peptide link.
If you scale this up to a polypeptide (a protein chain), each of the peptide
links will be broken in exactly the same way. That means that you will
end up with a mixture of the amino acids that made up the protein although in the form of their positive ions because of the presence of the
hydrogen ions from the hydrochloric acid.

Note: If you aren't sure about the formation of these positive


ions from the reactions between amino acids and hydrogen
ions, it would be useful to follow this link.
Use the BACK button on your browser to return to this page.

Doing the reaction


There are two ways of carrying out this reaction - an old, slow method,
and a new, fast one.
The old slow way
The protein is heated with 6 M hydrochloric acid for about 24 hours at
110C. (6M hydrochloric acid is slightly more than semi-concentrated.)
The new fast way
Protein samples are placed in tubes in a sealed container containing 6 M
hydrochloric acid in an atmosphere of nitrogen.
The whole container is then placed in a microwave oven for about 5 - 30
minutes (depending on the protein) with temperatures up to 200C.
The hydrochloric acid vaporises, comes into contact with the protein
http://www.chemguide.co.uk/organicprops/aminoacids/proteinhydrolysis.html (2 of 3)30/12/2004 11:24:48

protein hydrolysis

samples and hydrolyses them.


This method is used to hydrolyse small samples of protein during protein
analysis.

Where would you like to go now?


To the amino acid and proteins menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/aminoacids/proteinhydrolysis.html (3 of 3)30/12/2004 11:24:48

making carboxylic acids

MAKING CARBOXYLIC ACIDS

This page looks at ways of making carboxylic acids in the lab by the
oxidation of primary alcohols or aldehydes, and by the hydrolysis of
nitriles.

Note: If you are interested in the preparation of benzoic acid


(benzenecarboxylic acid) you will find it described in the
section on arenes (aromatic hydrocarbons like benzene and
methylbenzene). Benzoic acid is normally made by oxidising
hydrocarbon side chains attached to a benzene ring. This is
described towards the bottom of the page you will get to by
following this link.
If you choose to follow this link, use the BACK button on your
browser to return to this page.

Making carboxylic acids by oxidising primary alcohols or


aldehydes
Chemistry of the reactions
Primary alcohols and aldehydes are normally oxidised to carboxylic acids
using potassium dichromate(VI) solution in the presence of dilute
sulphuric acid. During the reaction, the potassium dichromate(VI)
solution turns from orange to green.
The potassium dichromate(VI) can just as well be replaced with sodium
dichromate(VI). Because what matters is the dichromate(VI) ion, all the
equations and colour changes would be identical.
Primary alcohols are oxidised to carboxylic acids in two stages - first to
an aldehyde and then to the acid. We often use simplified versions of
these equations using "[O]" to represent oxygen from the oxidising agent.

http://www.chemguide.co.uk/organicprops/acids/preparation.html (1 of 7)30/12/2004 11:24:57

making carboxylic acids

The formation of the aldehyde is shown by the simplified equation:

"R" is a hydrogen atom or a hydrocarbon group such as an alkyl group.

Note: Although "R" can in principle be a hydrogen atom, in


practice if it is a hydrogen, the oxidation eventually goes all
the way to carbon dioxide and water rather than stopping at
methanoic acid. Unlike most other carboxylic acids,
methanoic acid is very easily oxidised.

The aldehyde is then oxidised further to give the carboxylic acid:

If you start with an aldehyde, you are obviously just doing this second
stage.
Starting from the primary alcohol, you could combine these into one
single equation to give:

For example, if you were converting ethanol into ethanoic acid, the
simplified equation would be:

http://www.chemguide.co.uk/organicprops/acids/preparation.html (2 of 7)30/12/2004 11:24:57

making carboxylic acids

It is possible that you might want to write proper equations for these
reactions rather than these simplified ones. You can work these out from
electron-half-equations. How you do this is described in detail elsewhere
on the site.
The complete equation for the conversion of a primary alcohol to a
carboxylic acid is:

. . . or if you were starting from an aldehyde is:

Note: The equation for the conversion of ethanol to ethanoic


acid is worked out in detail on the page about electron-halfequations mentioned above. It isn't difficult to work out the
aldehyde one using the same principles - it would be a useful
exercise to check that you understand what to do.
If you choose to follow this link, use the BACK button on your
browser to return to this page.

Doing the reactions


It would actually be quite uncommon to make an acid starting from an
aldehyde, but very common to start from a primary alcohol. The
conversion of ethanol into ethanoic acid would be a typical example.
The alcohol is heated under reflux with an excess of a mixture of
potassium dichromate(VI) solution and dilute sulphuric acid.
Heating under reflux (heating in a flask with a condenser placed vertically
in it) prevents any aldehyde formed escaping before it has time to be
oxidised to the carboxylic acid.
http://www.chemguide.co.uk/organicprops/acids/preparation.html (3 of 7)30/12/2004 11:24:57

making carboxylic acids

Using an excess of oxidising agent is to be sure that there is enough


oxidising agent present for the oxidation to go all the way to the
carboxylic acid.
When oxidation is complete, the mixture can be distilled. You end up with
an aqueous solution of the acid.

Warning! The early stages of this reaction can be extremely


vigorous if great care isn't taken in producing the original
mixture in the flask. I have seen several cases of reaction
mixtures which suddenly boiled and spurted out of the top of
the condenser. In particular, the reagents have to be
thoroughly mixed before you start heating.

Making carboxylic acids by hydrolysing nitriles


What are nitriles?
Nitriles are compounds which contain -CN attached to a hydrocarbon
group. Some common examples include:

The name is based on the total number of carbons in the longest chain including the one in the -CN group. Where you have things substituted
into the chain (as in the third example), the -CN carbon counts as
number 1.
Where do nitriles come from?
Nitriles are produced in two important reactions - both of which result in
an increase in the length of the carbon chain because of the extra carbon
in the -CN group.
http://www.chemguide.co.uk/organicprops/acids/preparation.html (4 of 7)30/12/2004 11:24:57

making carboxylic acids

They are formed in the reaction between halogenoalkanes (haloalkanes


or alkyl halides) and cyanide ions. For example:

. . . or during the reaction between aldehydes or ketones and hydrogen


cyanide. For example, the reaction between ethanal and hydrogen
cyanide to make 2-hydroxypropanenitrile is:

Note: The first reaction is described on the page about


halogenoalkanes and cyanide ions. The second one is
described on a page about addition reactions of aldehydes
and ketones. Both of these pages include links to other
pages dealing with the mechanisms of these reactions.
If you choose to follow these links, use the BACK button on
your browser to return to this page.

Converting the nitrile into a carboxylic acid


There are two ways of doing this, both of which involve reacting the
carbon-nitrogen triple bond with water. This is described as hydrolysis.
The two methods produce slightly different products - you just have to be
careful to get this right.
Acid hydrolysis
The nitrile is heated under reflux with a dilute acid such as dilute
hydrochloric acid. A carboxylic acid is formed. For example, starting from
ethanenitrile you would get ethanoic acid. The ethanoic acid could be
distilled off the mixture.

http://www.chemguide.co.uk/organicprops/acids/preparation.html (5 of 7)30/12/2004 11:24:57

making carboxylic acids

Alkaline hydrolysis
The nitrile is heated under reflux with an alkali such as sodium hydroxide
solution.
This time you wouldn't, of course, get a carboxylic acid produced - any
acid formed would react with the sodium hydroxide present to give a salt.
You also wouldn't get ammonium ions because they would react with
sodium hydroxide to produce ammonia.
Starting from ethanenitrile, you would therefore get a solution containing
ethanoate ions (for example, sodium ethanoate if you used sodium
hydroxide solution) and ammonia.

You have to remember to convert the ions into the free carboxylic acid,
because that's what we are trying to make. To liberate the weak acid,
ethanoic acid, you just have to supply hydrogen ions from a strong acid
such as hydrochloric acid. You add enough hydrochloric acid to the
mixture to make it acidic.

Now you can distill off the carboxylic acid.

Note: If you have a choice in an exam, it obviously makes a


lot of sense to use acid hydrolysis rather than alkaline
hydrolysis - it saves a lot of bother!

http://www.chemguide.co.uk/organicprops/acids/preparation.html (6 of 7)30/12/2004 11:24:57

making carboxylic acids

Where would you like to go now?


To the carboxylic acids menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/acids/preparation.html (7 of 7)30/12/2004 11:24:57

some more reactions of benzene and methylbenzene

ASSORTED REACTIONS OF BENZENE AND


METHYLBENZENE

This page gives details of some reactions of benzene and


methylbenzene (toluene) not covered elsewhere in this section. It deals
with the combustion, hydrogenation and sulphonation of benzene and
methylbenzene (toluene), and with the oxidation of side chains attached
to benzene rings.
Remember that benzene, methylbenzene and similar hydrocarbons
based on benzene rings are collectively known as arenes.

Combustion
Like any other hydrocarbons, benzene and methylbenzene burn in a
plentiful supply of oxygen to give carbon dioxide and water. For example:
For benzene:

. . . and methylbenzene:

However, for these hydrocarbons, combustion is hardly ever complete,


especially if they are burnt in air. The high proportion of carbon in the
molecules means that you need a very high proportion of oxygen to
hydrocarbon to get complete combustion. Look at the equations.
As a general rule, the hydrogen in a hydrocarbon tends to get what
oxygen is available first, leaving the carbon to form carbon itself, or
carbon monoxide, if there isn't enough oxygen to go round.
The arenes tend to burn in air with extremely smoky flames - full of
carbon particles.
http://www.chemguide.co.uk/organicprops/arenes/other.html (1 of 7)30/12/2004 11:25:04

some more reactions of benzene and methylbenzene

You almost invariably get incomplete combustion, and the arenes can be
recognised by the smokiness of their flames.

Hydrogenation
Hydrogenation is an addition reaction in which hydrogen atoms are
added all the way around the benzene ring. A cycloalkane is formed. For
example:
With benzene:

. . . and methylbenzene:

These reactions destroy the electron delocalisation in the original


benzene ring, because those electrons are being used to form bonds
with the new hydrogen atoms.
Although the reactions are exothermic overall because of the strengths of
all the new carbon-hydrogen bonds being made, there is a high
activation barrier to the reaction.
The reactions are done using the same finely divided nickel catalyst that
is used in hydrogenating alkenes and at similar temperatures (around
150C), but the pressures used tend to be higher.
http://www.chemguide.co.uk/organicprops/arenes/other.html (2 of 7)30/12/2004 11:25:04

some more reactions of benzene and methylbenzene

Note: Pressure values quoted can be anywhere from about


20 atmospheres to 200 atmospheres. I have no way of
knowing which is right!
This hydrogenation reaction is important in estimating the
delocalisation energy for benzene. You can find more about
this by following this link to a page about the Kekul structure
for benzene.
Use the BACK button on your browser to return to this page
later.

Sulphonation
Sulphonation involves replacing one of the hydrogens on a benzene ring
by the sulphonic acid group, -SO3H.
The sulphonation of benzene
There are two equivalent ways of sulphonating benzene:

Heat benzene under reflux with concentrated sulphuric acid for


several hours.
Warm benzene under reflux at 40C with fuming sulphuric acid for
20 to 30 minutes. Fuming sulphuric acid, H2S2O7, can usefully be
thought of as a solution of sulphur trioxide in concentrated
sulphuric acid.

Or:
http://www.chemguide.co.uk/organicprops/arenes/other.html (3 of 7)30/12/2004 11:25:04

some more reactions of benzene and methylbenzene

The product is benzenesulphonic acid.

Note: You will find the mechanism for this reaction by


following this link.
Use the BACK button on your browser to return to this page.

The sulphonation of methylbenzene


Methylbenzene is more reactive than benzene because of the tendency
of the methyl group to "push" electrons towards the ring. Exactly how this
increases the rate of reaction is beyond UK A level - it is rather more
complicated than just an increase in the electron density of the ring.
The effect of this greater reactivity is that methylbenzene will react with
fuming sulphuric acid at 0C, and with concentrated sulphuric acid if they
are heated under reflux for about 5 minutes.
As well as the effect on the rate of reaction, with methylbenzene you also
have to think about where the sulphonic acid group ends up on the ring
relative to the methyl group.
Methyl groups have a tendency to "direct" new groups into the 2- and 4positions on the ring (assuming the methyl group is in the 1- position).
Methyl groups are said to be 2,4-directing. The origin of this directing
effect is also beyond UK A level.
So you get a mixture which mainly consists of two isomers. Only about 5
- 10% of the 3- isomer is formed.
The main reactions are:

http://www.chemguide.co.uk/organicprops/arenes/other.html (4 of 7)30/12/2004 11:25:04

some more reactions of benzene and methylbenzene

and:

In the case of sulphonation, the exact proportion of the isomers formed


depends on the temperature of the reaction. As the temperature
increases, you get increasing proportions of the 4- isomer and less of the
2- isomer.
This is because sulphonation is reversible. The sulphonic acid group can
fall off the ring again, and reattach somewhere else. This tends to favour
the formation of the most thermodynamically stable isomer. This
interchange happens more at higher temperatures.
The 4- isomer is more stable because there is no cluttering in the
molecule as there would be if the methyl group and sulphonic acid group
were next door to each other.

Side chain oxidation in alkylbenzenes


An alkylbenzene is simply a benzene ring with an alkyl group attached to
it. Methylbenzene is the simplest alkylbenzene.
Alkyl groups are usually fairly resistant to oxidation. However, when they
are attached to a benzene ring, they are easily oxidised by an alkaline
solution of potassium manganate(VII) (potassium permanganate).
Methylbenzene is heated under reflux with a solution of potassium
manganate(VII) made alkaline with sodium carbonate. The purple colour
of the potassium manganate(VII) is eventually replaced by a dark brown
precipitate of manganese(IV) oxide.
http://www.chemguide.co.uk/organicprops/arenes/other.html (5 of 7)30/12/2004 11:25:04

some more reactions of benzene and methylbenzene

The mixture is finally acidified with dilute sulphuric acid.


Overall, the methylbenzene is oxidised to benzoic acid.

Interestingly, any alkyl group is oxidised back to a -COOH group on the


ring under these conditions. So, for example, propylbenzene is also
oxidised to benzoic acid.

Note: These are flow schemes - deliberately not full


equations. To be honest, I don't really know whether you
could write an accurate single equation for anything more
complicated than a methyl group attached to the ring. In other
cases, you will certainly get some carbon dioxide produced,
but possibly some other organic molecules as well depending
on the conditions.
The separation of the benzoic acid is beyond the scope of
this site. The sulphuric acid converts benzoate ions (formed
under the alkaline conditions) into benzoic acid. The problem
lies in separating solid benzoic acid from solid manganese
(IV) oxide.
This reaction has some similarities to the oxidation of alkenes
by potassium manganate(VII), which you might also like to
have a look at if it is on your syllabus.
Use the BACK button on your browser to return to this page
later if you choose to follow this link.

http://www.chemguide.co.uk/organicprops/arenes/other.html (6 of 7)30/12/2004 11:25:04

some more reactions of benzene and methylbenzene

Where would you like to go now?


To the arenes menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/arenes/other.html (7 of 7)30/12/2004 11:25:04

alkenes and potassium manganate(VII) (permanganate)

ALKENES and POTASSIUM MANGANATE(VII)

This page looks at the reaction of the carbon-carbon double bond in


alkenes such as ethene with potassium manganate(VII) solution (potassium
permanganate solution).

Oxidation of alkenes
Experimental details
Alkenes react with potassium manganate(VII) solution in the cold. The
colour change depends on whether the potassium manganate(VII) is used
under acidic or alkaline conditions.
If the potassium manganate(VII) solution is acidified with dilute sulphuric
acid, the purple solution becomes colourless.
If the potassium manganate(VII) solution is made slightly alkaline (often by
adding sodium carbonate solution), the purple solution first becomes dark
green and then produces a dark brown precipitate.
Chemistry of the reaction
We'll look at the reaction with ethene. Other alkenes react in just the same
way.
Manganate(VII) ions are a strong oxidising agent, and in the first instance
oxidise ethene to ethane-1,2-diol (old name: ethylene glycol).
Looking at the equation purely from the point of view of the organic reaction:

http://www.chemguide.co.uk/organicprops/alkenes/kmno4.html (1 of 6)30/12/2004 11:25:09

alkenes and potassium manganate(VII) (permanganate)

Note: This type of equation is quite commonly used in organic


chemistry. Oxygen written in square brackets is taken to mean
"oxygen from an oxidising agent". The reason for this is that a
more normal equation tends to obscure the organic change in a
mass of other detail - as you will find below!
The full equations are given below, although you probably won't
need them.

The full equation depends on the conditions.


Under acidic conditions, the manganate(VII) ions are reduced to
manganese(II) ions.

Note: If you want to know how to write equations for redox


reactions like this you could follow this link, and explore in the
redox section of this site.
Use the BACK button (or HISTORY file or GO menu) on your
browser to return to this page later.

http://www.chemguide.co.uk/organicprops/alkenes/kmno4.html (2 of 6)30/12/2004 11:25:09

alkenes and potassium manganate(VII) (permanganate)

Under alkaline conditions, the manganate(VII) ions are first reduced to


green manganate(VI) ions . . .

. . . and then further to dark brown solid manganese(IV) oxide (manganese


dioxide).

This last reaction is also the one you would get if the reaction was done
under neutral conditions. You will notice that there are neither hydrogen
ions nor hydroxide ions on the left-hand side of the equation.

Note: You might possibly remember that further up the page it


says that potassium manganate(VII) is often made slightly
alkaline by adding sodium carbonate solution. Where are the
hydroxide ions in this?
Carbonate ions react with water to some extent to produce
hydrogencarbonate ions and hydroxide ions. It is the presence
of these hydroxide ions that gives sodium carbonate solution its
pH in the 10 - 11 region.

http://www.chemguide.co.uk/organicprops/alkenes/kmno4.html (3 of 6)30/12/2004 11:25:09

alkenes and potassium manganate(VII) (permanganate)

Complications
(. . . and you thought this was already complicated enough?)
The product, ethane-1,2-diol, is itself quite easily oxidised by manganate
(VII) ions, and so the reaction won't stop at this point unless the potassium
manganate(VII) solution is very dilute, very cold, and preferably not under
acidic conditions.
That means that this reaction has little use as a way of preparing ethane1,2-diol. Its only real use is in testing for carbon-carbon double bonds - and
even then it isn't very good!

Note: Ethane-1,2-diol is an alcohol, although unlike simple


ones like ethanol it contains two -OH groups. The oxidation of
alcohols is explored on another page if you are interested.
That page deals with the oxidation of alcohols by acidified
potassium dichromate(VI) solution - a slightly less powerful
oxidising agent than potassium manganate(VII). The essential
chemistry will be the same, although manganate(VII) ions
eventually oxidise ethane-1,2-diol all the way to carbon dioxide
and water. You only need read the first part of that page about
oxidation of primary alcohols.
Use the BACK button on your browser to return to this page if
you choose to follow this link.

http://www.chemguide.co.uk/organicprops/alkenes/kmno4.html (4 of 6)30/12/2004 11:25:09

alkenes and potassium manganate(VII) (permanganate)

Using the reaction to test for carbon-carbon double bonds


If an organic compound reacts with dilute alkaline potassium manganate
(VII) solution to give a green solution followed by a dark brown precipitate,
then it may contain a carbon-carbon double bond. But equally it could be
any one of a large number of other compounds all of which can be oxidised
by manganate(VII) ions under alkaline conditions.
The situation with acidified potassium manganate(VII) solution is even
worse because it has a tendency to break carbon-carbon bonds. It reacts
destructively with a large number of organic compounds and is rarely used
in organic chemistry.
You could use alkaline potassium manganate(VII) solution if, for example,
all you had to do was to find out whether a hydrocarbon was an alkane or
an alkene - in other words, if there was nothing else present which could be
oxidised.
It isn't a useful test. Bromine water is far more clear cut.

Note: You will find details of the use of bromine water in testing
for carbon-carbon double bonds in the page about the reactions
of alkenes with halogens.

Where would you like to go now?


To the alkenes menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

http://www.chemguide.co.uk/organicprops/alkenes/kmno4.html (5 of 6)30/12/2004 11:25:09

alkenes and potassium manganate(VII) (permanganate)

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/alkenes/kmno4.html (6 of 6)30/12/2004 11:25:09

halogenation of alkenes

THE HALOGENATION OF ALKENES

This page looks at the reaction of the carbon-carbon double bond in


alkenes such as ethene with halogens such as chlorine, bromine and
iodine. This is called halogenation.
Reactions where the chlorine or bromine are in solution (for example,
"bromine water") are slightly more complicated and are treated
separately at the end.

Simple reactions involving halogens


In each case, we will look at ethene as typical of all of the alkenes. There
are no complications as far as the basic facts are concerned as the
alkenes get bigger.
Ethene and fluorine
Ethene reacts explosively with fluorine to give carbon and hydrogen
fluoride gas. This isn't a useful reaction, and you aren't likely to need it
for UK A level purposes.

Ethene and chlorine or bromine or iodine


In each case you get an addition reaction. For example, bromine adds
to give 1,2-dibromoethane.

http://www.chemguide.co.uk/organicprops/alkenes/halogenation.html (1 of 4)30/12/2004 11:25:12

halogenation of alkenes

Note: Follow this link if you aren't happy about naming


organic compounds
Use the BACK button on your browser to return to this page.

The reaction with bromine happens at room temperature. If you have a


gaseous alkene like ethene, you can bubble it through either pure liquid
bromine or a solution of bromine in an organic solvent like
tetrachloromethane. The reddish-brown bromine is decolourised as it
reacts with the alkene.
A liquid alkene (like cyclohexene) can be shaken with liquid bromine or
its solution in tetrachloromethane.
Chlorine reacts faster than bromine, but the chemistry is similar. Iodine
reacts much, much more slowly, but again the chemistry is similar. You
are much more likely to meet the bromine case than either of these.

Note: If you are interested in the mechanism for the addition


of bromine to alkenes you will find it in the mechanism
section of this site.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/alkenes/halogenation.html (2 of 4)30/12/2004 11:25:12

halogenation of alkenes

Alkenes and bromine water


Using bromine water as a test for alkenes
If you shake an alkene with bromine water (or bubble a gaseous alkene
through bromine water), the solution becomes colourless. Alkenes
decolourise bromine water.
The chemistry of the test
This is complicated by the fact that the major product isn't 1,2dibromoethane. The water also gets involved in the reaction, and most of
the product is 2-bromoethanol.

However, there will still be some 1,2-dibromoethane formed, so at this


sort of level you can probably get away with quoting the simpler equation:

Note: Check your syllabus, past papers and mark schemes


to find out what you need to be able to do. If you are a UK A
level student and haven't got any of these, follow this link to
find out how to get hold of them. If you are working at a lower
level, use the simpler equation.
If you are interested in the mechanism for this reaction, I'm
afraid that you won't find it on this site because it isn't on any
of the UK A level syllabuses. You can, however, deduce it
fairly easily if you know the mechanism for the addition of
pure bromine. Instead of the intermediate ion being attacked
by a bromide ion, it is much more likely to be hit by a lone
pair on an oxygen of a water molecule, followed by loss of a
hydrogen ion from the product.

http://www.chemguide.co.uk/organicprops/alkenes/halogenation.html (3 of 4)30/12/2004 11:25:12

halogenation of alkenes

Where would you like to go now?


To the alkenes menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/alkenes/halogenation.html (4 of 4)30/12/2004 11:25:12

catalytic reforming to make benzene

MANUFACTURING ARENES

This page looks at the manufacture of arenes such as benzene and


methylbenzene (toluene) by the catalytic reforming of fractions from
petroleum (crude oil).

Catalytic reforming
What is reforming?
Reforming takes straight chain hydrocarbons in the C6 to C8 range from
the gasoline or naphtha fractions and rearranges them into compounds
containing benzene rings. Hydrogen is produced as a by-product of the
reactions.
For example, hexane, C6H14, loses hydrogen and turns into benzene. As
long as you draw the hexane bent into a circle, it is easy to see what is
happening.

Similarly, methylbenzene (toluene) is made from heptane:

http://www.chemguide.co.uk/organicprops/arenes/manufacture.html (1 of 3)30/12/2004 11:25:17

catalytic reforming to make benzene

The process
The feedstock
The feedstock is a mixture of the naphtha or gasoline fractions and
hydrogen. The hydrogen is there to help prevent the formation of carbon
by decomposition of the hydrocarbons at the high temperatures used.
The carbon would otherwise contaminate the catalyst.
The catalyst
A typical catalyst is a mixture of platinum and aluminium oxide. With a
platinum catalyst, the process is sometimes described as "platforming".
Temperature and pressure
The temperature is about 500C, and the pressure varies either side of
20 atmospheres.

Converting some of the methylbenzene into benzene


Methylbenzene is much less commercially valuable than benzene. The
methyl group can be removed from the ring by a process known as
"dealkylation".
The methylbenzene is mixed with hydrogen at a temperature of between
550 and 650C, and a pressure of between 30 and 50 atmospheres, with
a mixture of silicon dioxide and aluminium oxide as catalyst.

http://www.chemguide.co.uk/organicprops/arenes/manufacture.html (2 of 3)30/12/2004 11:25:17

catalytic reforming to make benzene

Where would you like to go now?


To the arenes menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/arenes/manufacture.html (3 of 3)30/12/2004 11:25:17

nitration of benzene and methylbenzene

NITRATION OF BENZENE AND


METHYLBENZENE

This page looks at the facts about the nitration of benzene and
methylbenzene. The mechanisms for these reactions are covered
elsewhere on the site, and you will find links to these.

The nitration of benzene


Nitration happens when one (or more) of the hydrogen atoms on the
benzene ring is replaced by a nitro group, NO2.
Benzene is treated with a mixture of concentrated nitric acid and
concentrated sulphuric acid at a temperature not exceeding 50C. The
mixture is held at this temperature for about half an hour. Yellow oily
nitrobenzene is formed.

You could write this in a more condensed form as:

The concentrated sulphuric acid is acting as a catalyst and so isn't


written into the equations.
At higher temperatures there is a greater chance of getting more than
one nitro group substituted onto the ring. You will get a certain amount of
1,3-dinitrobenzene formed even at 50C. Some of the nitrobenzene
formed reacts with the nitrating mixture of concentrated acids.

http://www.chemguide.co.uk/organicprops/arenes/nitration.html (1 of 5)30/12/2004 11:25:23

nitration of benzene and methylbenzene

Notice that the new nitro group goes into the 3 position on the ring. Nitro
groups "direct" new groups into the 3 and 5 positions. The reasons for
this "directing effect" are beyond UK A level.

Note: The numbering on the ring goes clockwise around the


ring starting with number 1 at the top.

It is also possible to get a third nitro group attached to the ring (in the 5
position). However, nitro groups make the ring much less reactive than
the original benzene ring. Two nitro groups on the ring make its reactions
so slow that virtually no trinitrobenzene is produced under these
conditions.

Note: Follow this link if you want the mechanism for the
nitration of benzene.
You will find the mechanism for the nitration of nitrobenzene
(the reaction producing 1,3-dinitrobenzene) at the bottom of
the page you will get to by following this second link.
If you choose to follow either of these links, use the BACK
button (or the HISTORY file or GO menu) on your browser to
return to this page later.

http://www.chemguide.co.uk/organicprops/arenes/nitration.html (2 of 5)30/12/2004 11:25:23

nitration of benzene and methylbenzene

The nitration of methylbenzene (toluene)


Methylbenzene reacts rather faster than benzene - in nitration, the
reaction is about 25 times faster. That means that you would use a lower
temperature to prevent more than one nitro group being substituted - in
this case, 30C rather than 50C. Apart from that, the reaction is just the
same - using the same nitrating mixture of concentrated sulphuric and
nitric acids.
You get a mixture of mainly two isomers formed: 2-nitromethylbenzene
and 4-nitromethylbenzene. Only about 5% of the product is 3nitromethylbenzene. Methyl groups are said to be 2,4-directing.
For 2-nitromethylbenzene:

. . . and for 4-nitromethylbenzene:

Note: You will find the mechanism for the nitration of


methylbenzene by following this link.
Use the BACK button (or HISTORY file or GO menu) on your
browser to return to this page later.

http://www.chemguide.co.uk/organicprops/arenes/nitration.html (3 of 5)30/12/2004 11:25:23

nitration of benzene and methylbenzene

Just as with benzene, you will get a certain amount of dinitro compound
formed under the conditions of the reaction, but virtually no trinitro
product because the reactivity of the ring decreases for every nitro group
added. From an experimental point of view this is just as well.
Trinitromethylbenzene used to be called trinitrotoluene or TNT!
The reason for the 2,4-directing effect of the methyl group is beyond UK
A level, but the increased reaction rate can be explained up to a point.
The reactivity of a benzene ring is governed by the electron density
around the ring. Methyl groups tend to "push" electrons towards the ring increasing the density, and so making the ring more attractive to
attacking reagents.
This is actually a simplification. In order to understand the rate effect
properly you have to think about the stability of the intermediate ions
formed during the reactions, because this affects the activation energy of
the reactions. This is also the basis for the directing effect and is again
beyond UK A level.

Warning! I have deliberately chosen not to use the strict


IUPAC names for the compounds in this section, because I
see them as illogical and inconsistent with the older names
for these compounds. The names for the compounds with
one nitro group substituted into methylbenzene should be
methyl-2-nitrobenzene and methyl-4-nitrobenzene. The
reason for these names is that you should strictly name the
attached groups in alphabetical order.
It seems to me to be illogical on three counts. Firstly, it
breaks up the name of the hydrocarbon, so that it is no longer
immediately obvious that you are talking about a derivative of
methylbenzene.
Secondly, it doesn't relate to the older names for these
compounds. Methyl-4-nitrobenzene, for example, used to be
called para-nitrotoluene. The "para" refers to the 4 position.
You could equally have called it 4-nitrotoluene. My name of 4nitromethylbenzene is a direct translation of this. I think that's

http://www.chemguide.co.uk/organicprops/arenes/nitration.html (4 of 5)30/12/2004 11:25:23

nitration of benzene and methylbenzene

entirely logical!
Thirdly, you can end up with completely different names for
compounds which are structurally similar. For example, if you
think about substituting chlorines into these positions instead
of nitro groups you will have to completely change the names
for purely alphabetical reasons. That's silly!
I may well be in a minority of one on this, but I am sticking to
my guns on it. In fact, IUPAC is much more flexible about
these things than they are sometimes given credit for. I doubt
if it would worry them, although it might upset your teachers
or lecturers. Obviously, if you are working in a system which
still calls methylbenzene "toluene", none of this bothers you!

Where would you like to go now?


To the arenes menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/arenes/nitration.html (5 of 5)30/12/2004 11:25:23

halogenation of benzene and methylbenzene

HALOGENATION OF BENZENE AND


METHYLBENZENE

This page looks at the reactions of benzene and methylbenzene


(toluene) with chlorine and bromine under various conditions. The
mechanisms for several of these reactions are covered elsewhere on the
site and you will find links to these other pages.

The halogenation of benzene


Substitution reactions
Benzene reacts with chlorine or bromine in the presence of a catalyst,
replacing one of the hydrogen atoms on the ring by a chlorine or bromine
atom.
The reactions happen at room temperature. The catalyst is either
aluminium chloride (or aluminium bromide if you are reacting benzene
with bromine) or iron.
Strictly speaking iron isn't a catalyst, because it gets permanently
changed during the reaction. It reacts with some of the chlorine or
bromine to form iron(III) chloride, FeCl3, or iron(III) bromide, FeBr3.

These compounds act as the catalyst and behave exactly like aluminium
chloride, AlCl3, or aluminium bromide, AlBr3, in these reactions.
The reaction with chlorine
The reaction between benzene and chlorine in the presence of either
aluminium chloride or iron gives chlorobenzene.
http://www.chemguide.co.uk/organicprops/arenes/halogenation.html (1 of 7)30/12/2004 11:25:32

halogenation of benzene and methylbenzene

or, written more compactly:

The reaction with bromine


The reaction between benzene and bromine in the presence of either
aluminium bromide or iron gives bromobenzene. Iron is usually used
because it is cheaper and more readily available.

or:

Note: Follow this link if you want the mechanism for the
halogenation of benzene.
Use the BACK button on your browser to return to this page
later.

http://www.chemguide.co.uk/organicprops/arenes/halogenation.html (2 of 7)30/12/2004 11:25:32

halogenation of benzene and methylbenzene

Addition reactions
In the presence of ultraviolet light (but without a catalyst present), hot
benzene will also undergo an addition reaction with chlorine or bromine.
The ring delocalisation is permanently broken and a chlorine or bromine
atom adds on to each carbon atom.
For example, if you bubble chlorine gas through hot benzene exposed to
UV light for an hour, you get 1,2,3,4,5,6-hexachlorocyclohexane.

Bromine would behave similarly.


The chlorines and hydrogens can stick up and down at random above
and below the ring and this leads to a number of geometric isomers.
Although there aren't any carbon-carbon double bonds, the bonds are
still "locked" and unable to rotate.
One of these isomers was once commonly used as an insecticide known
variously as BHC, HCH and Gammexane. This is one of the "chlorinated
hydrocarbons" which caused so much environmental harm.

Note: Geometric isomerism is explored on another page,


although only with regard to carbon-carbon double bonds.
It isn't too difficult to relate this to the ring case. For example,
the chlorine atoms in the 1 and 2 positions on the ring could
both be pointing up above the ring (or down below it) or one
could be pointing up and the other one down. That
corresponds exactly to the cis and trans positions that you
are probably familiar with in carbon-carbon double bonds.
Use the BACK button on your browser to return to this page if
you choose to follow this link.
http://www.chemguide.co.uk/organicprops/arenes/halogenation.html (3 of 7)30/12/2004 11:25:32

halogenation of benzene and methylbenzene

The halogenation of methylbenzene


Substitution reactions
It is possible to get two quite different substitution reactions between
methylbenzene and chlorine or bromine depending on the conditions
used. The chlorine or bromine can substitute into the ring or into the
methyl group.
Substitution into the ring
Substitution in the ring happens in the presence of aluminium chloride (or
aluminium bromide if you are using bromine) or iron, and in the absence
of UV light. The reactions happen at room temperature.
This is exactly the same as the reaction with benzene, except that you
have to worry about where the halogen atom attaches to the ring relative
to the position of the methyl group.
Methyl groups are 2,4-directing, which means that incoming groups will
tend to go into the 2 or 4 positions on the ring - assuming the methyl
group is in the 1 position. In other words, the new group will attach to the
ring next door to the methyl group or opposite it. The reason for the
directing effect is beyond UK A level.
With chlorine, substitution into the ring gives a mixture of 2chloromethylbenzene and 4-chloromethylbenzene.

http://www.chemguide.co.uk/organicprops/arenes/halogenation.html (4 of 7)30/12/2004 11:25:32

halogenation of benzene and methylbenzene

With bromine, you would get the equivalent bromine compounds.

Note: You will find the mechanism for this reaction half way
down the page you will get to by following this link.
Use the BACK button on your browser to return to this page
later.

Substitution into the methyl group


If chlorine or bromine react with boiling methylbenzene in the absence of
a catalyst but in the presence of UV light, substitution happens in the
methyl group rather than the ring.
For example, with chlorine (bromine would be similar):

The organic product is (chloromethyl)benzene. The brackets in the name


emphasise that the chlorine is part of the attached methyl group, and isn't
on the ring.
One of the hydrogen atoms in the methyl group has been replaced by a
chlorine atom. However, the reaction doesn't stop there, and all three
hydrogens in the methyl group can in turn be replaced by chlorine atoms.
That means that you could also get (dichloromethyl)benzene and
(trichloromethyl)benzene as the other hydrogen atoms in the methyl

http://www.chemguide.co.uk/organicprops/arenes/halogenation.html (5 of 7)30/12/2004 11:25:32

halogenation of benzene and methylbenzene

group are replaced one at a time.

If you use enough chlorine you will eventually get (trichloromethyl)


benzene, but any other proportions will always lead to a mixture of
products.

Note: You will find the mechanisms for these reactions


spread over 3 pages (depending on how much detail you
want) which you can get to by following this link
Use the BACK button (or HISTORY file or GO menu) on your
browser to return to this page.

Addition reactions
I haven't been able to track down anything similar to the reaction
between benzene and chlorine in which six chlorine atoms add around
the ring.
That perhaps isn't surprising. Chlorine adds to benzene in the presence
of ultraviolet light. With methylbenzene under those conditions, you get
substitution in the methyl group. That is energetically easier because it
doesn't involve breaking the delocalised electron system.
Whether you would get addition to the ring if you used a large excess of
chlorine and did the reaction for a long time, I don't know. Once all the
hydrogens in the methyl group had been substituted, perhaps you might
then get addition to the ring as well.
http://www.chemguide.co.uk/organicprops/arenes/halogenation.html (6 of 7)30/12/2004 11:25:32

halogenation of benzene and methylbenzene

None of this is needed for any of the UK A level syllabuses.

Note: If you have any solid information on this (either that


the reaction definitely does or definitely doesn't happen), I
would be grateful if you could contact me via the address on
the about this site page.

Where would you like to go now?


To the arenes menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/arenes/halogenation.html (7 of 7)30/12/2004 11:25:32

friedel-crafts reactions of benzene and methylbenzene

FRIEDEL-CRAFTS REACTIONS OF BENZENE


AND METHYLBENZENE

This page gives details of the Friedel-Crafts reactions of benzene and


methylbenzene (toluene). The mechanisms for some of these reactions
are covered elsewhere on the site and you will find links to those as you
go along.

Friedel-Crafts acylation
Friedel-Crafts acylation of benzene
What is acylation?
An acyl group is an alkyl group attached to a carbonoxygen double bond. If "R" represents any alkyl group,
then an acyl group has the formula RCO-. Acylation
means substituting an acyl group into something - in this
case, into a benzene ring.
The most commonly used acyl group is CH3CO-. This is called the
ethanoyl group, and in this case the reaction is sometimes called
"ethanoylation". In the example which follows we are substituting a
CH3CO- group into the ring, but you could equally well use any other acyl
group.
The facts
The most reactive substance containing an acyl group is an acyl chloride
(also known as an acid chloride). These have the general formula RCOCl.
Benzene is treated with a mixture of ethanoyl chloride, CH3COCl, and
aluminium chloride as the catalyst. The mixture is heated to about 60C
for about 30 minutes.

http://www.chemguide.co.uk/organicprops/arenes/fc.html (1 of 7)30/12/2004 11:25:38

friedel-crafts reactions of benzene and methylbenzene

A ketone called phenylethanone (old name: acetophenone) is formed.

Note: Don't worry too much about the name


"phenylethanone" - all that matters is that you can draw the
structure.
You may find slight variations on the conditions for this
reaction. Various recipes I have found vary the temperature
and time - for example by having a slightly lower temperature
for a longer time.

or, if you want a more compact form:

The aluminium chloride isn't written into these equations because it is


acting as a catalyst. If you wanted to include it, you could write AlCl3 over
the top of the arrow (see below).

Note: You will find the mechanism for this reaction in the
mechanisms section of this site.
Use the BACK button on your browser to return to this page
later.

http://www.chemguide.co.uk/organicprops/arenes/fc.html (2 of 7)30/12/2004 11:25:38

friedel-crafts reactions of benzene and methylbenzene

Friedel-Crafts acylation of methylbenzene (toluene)


The reaction is just the same with methylbenzene except that you have
to worry about where the acyl group attaches to the ring relative to the
methyl group.
Normally, the methyl group in methylbenzene directs new groups into the
2- and 4- positions (assuming the methyl group is in the 1- position). In
acylation, though, virtually all the substitution happens in the 4- position.

Note: The reason for the 2,4-directing effect of the methyl


group is beyond UK A level. The reason that you get virtually
none of the 2- isomer in this instance is because of the size
of the incoming acyl group. Everything gets too cluttered (and
therefore less stable) if you try to put the acyl group next door
to the methyl group.

Friedel-Crafts alkylation
Friedel-Crafts alkylation of benzene
What is alkylation?
Alkylation means substituting an alkyl group into something - in this case
into a benzene ring. A hydrogen on the ring is replaced by a group like
methyl or ethyl and so on.
The facts
Benzene reacts at room temperature with a chloroalkane (for example,
http://www.chemguide.co.uk/organicprops/arenes/fc.html (3 of 7)30/12/2004 11:25:38

friedel-crafts reactions of benzene and methylbenzene

chloromethane or chloroethane) in the presence of aluminium chloride as


a catalyst. On this page, we will look at substituting a methyl group, but
any other alkyl group could be used in the same way.
Substituting a methyl group gives methylbenzene.

or:

Note: The methylbenzene formed is more reactive than the


original benzene, and so the reaction doesn't stop there. You
get further methyl groups substituted around the ring. You
can improve your chances of just getting monosubstitution by
using a large excess of benzene.
You won't have to worry about this for UK A level purposes.
You will find the mechanism for this reaction in the
mechanisms section of this site.
Use the BACK button on your browser to return to this page
later.

http://www.chemguide.co.uk/organicprops/arenes/fc.html (4 of 7)30/12/2004 11:25:38

friedel-crafts reactions of benzene and methylbenzene

Friedel-Crafts alkylation of methylbenzene (toluene)


Again, the reaction is just the same with methylbenzene except that you
have to worry about where the alkyl group attaches to the ring relative to
the methyl group.
Unfortunately this time there is a problem! Where the incoming alkyl
group ends up depends to a large extent on the temperature of the
reaction.
At 0C, substituting methyl groups into methylbenzene, you get a mixture
of the 2-.3- and 4- isomers in the proportion 54% / 17% / 29%. That's a
higher proportion of the 3- isomer than you might expect.
At 25C, the proportions change to 3% / 69% / 28%. In other words the
proportion of the 3- isomer has increased even more. Raise the
temperature some more and the trend continues.
The reason for this is again beyond UK A level.

Note: The problem in this case lies in the fact that the methyl
groups attaching to the ring can fall off again and reattach
somewhere else in the presence of the aluminium chloride.
You can get equilibria set up between the various isomers.
The reason for the 2,4- directing effect of the methyl group in
methylbenzene lies in the fact that the 2- and 4- isomers form
faster than the 3- isomer. However, in this case, the 3isomer is the most thermodynamically stable of the three. If
you raise the temperature, or allow more time, the equilibria
set up favour the most stable product.
You do NOT have to worry about this for A level purposes.

http://www.chemguide.co.uk/organicprops/arenes/fc.html (5 of 7)30/12/2004 11:25:38

friedel-crafts reactions of benzene and methylbenzene

Friedel-Crafts alkylation industrially


The manufacture of ethylbenzene
Ethylbenzene is an important industrial chemical used to make styrene
(phenylethene), which in turn is used to make polystyrene - poly
(phenylethene).
It is manufactured from benzene and ethene. There are several ways of
doing this, some of which use a variation on Friedel-Crafts alkylation.
What follows is the method required by the UK A level Exam Board, AQA.
The reaction is done in the liquid state. Ethene is passed through a liquid
mixture of benzene, aluminium chloride and a catalyst promoter which
might be chloroethane or hydrogen chloride. We are going to assume it
is HCl (because that's what AQA want!).
Promoters are used to make catalysts work better.
There are two variants on the process. One (the Union Carbide / Badger
process) uses a temperature no higher than 130C and a pressure just
high enough to keep everything liquid.
The other (the Monsanto process) uses a slightly higher temperature of
160C which needs less catalyst. (Presumably - although I haven't been
able to confirm this - the pressure would also need to be higher to keep
everything liquid at the higher temperature.)

or:

Again, the aluminium chloride and HCl aren't written into these equations
because they are acting as catalysts. If you wanted to include them, you
could write AlCl3 and HCl over the top of the arrow.

http://www.chemguide.co.uk/organicprops/arenes/fc.html (6 of 7)30/12/2004 11:25:38

friedel-crafts reactions of benzene and methylbenzene

Note: You will find the mechanism for this reaction in the
mechanisms section of this site. This will show exactly what
the HCl and aluminium chloride are doing in the reaction.
Use the BACK button on your browser to return to this page
later.

Where would you like to go now?


To the arenes menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/arenes/fc.html (7 of 7)30/12/2004 11:25:38

addition to aldehydes and ketones

SIMPLE ADDITION TO ALDEHYDES AND


KETONES

This page looks at the addition of hydrogen cyanide and sodium


hydrogensulphite (sodium bisulphite) to aldehydes and ketones.

Addition of hydrogen cyanide to aldehydes and ketones


The reactions
Hydrogen cyanide adds across the carbon-oxygen double bond in
aldehydes and ketones to produce compounds known as hydroxynitriles.
These used to be known as cyanohydrins.
For example, with ethanal (an aldehyde) you get 2-hydroxypropanenitrile:

With propanone (a ketone) you get 2-hydroxy-2-methylpropanenitrile:

Note: When you are naming these compounds, don't forget


that the longest carbon chain has to include the carbon in the
-CN group. The carbon with the nitrogen attached is always
counted as the number 1 carbon in the chain.

http://www.chemguide.co.uk/organicprops/carbonyls/addition.html (1 of 5)30/12/2004 11:25:42

addition to aldehydes and ketones

The reaction isn't normally done using hydrogen cyanide itself, because
this is an extremely poisonous gas. Instead, the aldehyde or ketone is
mixed with a solution of sodium or potassium cyanide in water to which a
little sulphuric acid has been added. The pH of the solution is adjusted to
about 4 - 5, because this gives the fastest reaction. The reaction
happens at room temperature.
The solution will contain hydrogen cyanide (from the reaction between
the sodium or potassium cyanide and the sulphuric acid), but still
contains some free cyanide ions. This is important for the mechanism.

Note: If you want the mechanism for this reaction, you will
find it explained if you follow this link.
Use the BACK button on your browser to return to this page.

Uses of the reaction


The product molecules contain two functional groups:

the -OH group which behaves like a simple alcohol and can be
replaced by other things like chlorine, which can in turn be
replaced to give, for example, an -NH2 group;
the -CN group which is easily converted into a carboxylic acid
group -COOH.

For example, starting from a hydroxynitrile made from an aldehyde, you


can quite easily produce relatively complicated molecules like 2-amino
acids - the amino acids which are used to construct proteins.

http://www.chemguide.co.uk/organicprops/carbonyls/addition.html (2 of 5)30/12/2004 11:25:42

addition to aldehydes and ketones

Note: The first step is the replacement of -OH in an alcohol


by chlorine.
The second step is the reaction between a halogenoalkane
and ammonia.
The final step is hydrolysis of a nitrile.
Use the BACK button on your browser to return to this page if
you follow any of these links.

Addition of sodium hydrogensulphite to aldehydes and


ketones
Sodium hydrogensulphite used to be known as sodium bisulphite, and
you might well still come across it in organic textbooks under this name or using the bisulfite spelling.

Note: If you are a UK A level student, you probably won't


need this reaction. Check your syllabus before you bother to
read on. If you haven't got a syllabus, you ought to have!
Follow this link to find out how to get one.

http://www.chemguide.co.uk/organicprops/carbonyls/addition.html (3 of 5)30/12/2004 11:25:42

addition to aldehydes and ketones

The reactions
This reaction only works well for aldehydes. In the case of ketones, one
of the hydrocarbon groups attached to the carbonyl group needs to be a
methyl group. Bulky groups attached to the carbonyl group get in the way
of the reaction happening.
The aldehyde or ketone is shaken with a saturated solution of sodium
hydrogensulphite in water. Where the product is formed, it separates as
white crystals.

Note: In the case of methanal and ethanal, the product is


reasonably soluble. If you start with aqueous solutions of
methanal or ethanal, there may be enough water present that
the product doesn't actually form crystals.

In the case of ethanal, the equation is:

. . . and with propanone, the equation is:

These compounds are rarely named systematically, and are usually


known as "hydrogensulphite (or bisulphite) addition compounds".

Uses of the reaction


The reaction is usually used during the purification of aldehydes (and any
http://www.chemguide.co.uk/organicprops/carbonyls/addition.html (4 of 5)30/12/2004 11:25:42

addition to aldehydes and ketones

ketones that it works for). The addition compound can be split easily to
regenerate the aldehyde or ketone by treating it with either dilute acid or
dilute alkali.
If you have an impure aldehyde, for example, you could shake it with a
saturated solution of sodium hydrogensulphite to produce the crystals.
These crystals could easily be filtered and washed to remove any other
impurities. Addition of dilute acid, for example, would then regenerate the
original aldehyde.
It would, of course, still need to be separated from the excess acid and
assorted inorganic products of the reaction - but that is beyond the scope
of this page!

Where would you like to go now?


To the aldehydes and ketones menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/carbonyls/addition.html (5 of 5)30/12/2004 11:25:42

carboxylic acids as acids

CARBOXYLIC ACIDS AS ACIDS

This page looks at the simple reactions of carboxylic acids as acids,


including their reactions with metals, metal hydroxides, carbonates and
hydrogencarbonates, ammonia and amines.

Note: This page covers all the reactions likely to be asked


by any of the UK A level syllabuses, but no single syllabus is
likely to want all of them. Check your syllabus and past
papers to find out what you need to know. If you are a UK A
level student and haven't got these, follow this link to find out
how to get hold of them.

The acidity of the carboxylic acids


Why are carboxylic acids acidic?
Using the definition of an acid as a "substance which donates protons
(hydrogen ions) to other things", the carboxylic acids are acidic because
of the hydrogen in the -COOH group.
In solution in water, a hydrogen ion is transferred from the -COOH group
to a water molecule. For example, with ethanoic acid, you get an
ethanoate ion formed together with a hydroxonium ion, H3O+.
This reaction is reversible and, in the case of ethanoic acid, no more than
about 1% of the acid has reacted to form ions at any one time. (This is a
rough-and-ready figure and varies with the concentration of the solution.)
These are therefore weak acids.

http://www.chemguide.co.uk/organicprops/acids/acidity.html (1 of 8)30/12/2004 11:25:52

carboxylic acids as acids

This equation is often simplified to:

However, if you are going to use this second equation, you must include
state symbols. They imply that the hydrogen ion is actually attached to a
water molecule.

The pH of carboxylic acid solutions


The pH depends on both the concentration of the acid and how easily it
loses hydrogen ions from the -COOH group.
Ethanoic acid is typical of the acids where the -COOH group is attached
to a simple alkyl group. Typical lab solutions have pH's in the 2 - 3 range,
depending on their concentrations.
Methanoic acid is rather stronger than the other simple acids, and
solutions have pH's about 0.5 pH units less than ethanoic acid of the
same concentration.

Note: You will find factors affecting the acidity of organic


acids discussed in detail if you follow this link.
If you should know how to calculate the pH of weak acids like
ethanoic acid, but aren't happy about it, you might be
interested in my chemistry calculations book.
If you choose to follow either of these links, use the BACK
button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/acids/acidity.html (2 of 8)30/12/2004 11:25:52

carboxylic acids as acids

Reactions of the carboxylic acids


With metals
Carboxylic acids react with the more reactive metals to produce a salt
and hydrogen. The reactions are just the same as with acids like
hydrochloric acid, except they tend to be rather slower.
For example, dilute ethanoic acid reacts with magnesium. The
magnesium reacts to produce a colourless solution of magnesium
ethanoate, and hydrogen is given off. If you use magnesium ribbon, the
reaction is less vigorous than the same reaction with hydrochloric acid,
but with magnesium powder, both are so fast that you probably wouldn't
notice much difference.

Warning! Students have a high tendency to get the formulae


of salts like magnesium ethanoate wrong. Remember that
these simple carboxylic acids form ions with a single negative
charge, and that the magnesium ion has two positive
charges. In essence, this is no different from writing the
formula for magnesium chloride - except that with these salts
the metal is written at the end of the formula rather than at
the beginning.

With metal hydroxides


These are simple neutralisation reactions and are just the same as any
other reaction in which hydrogen ions from an acid react with hydroxide
ions. They are most quickly and easily represented by the equation:

http://www.chemguide.co.uk/organicprops/acids/acidity.html (3 of 8)30/12/2004 11:25:52

carboxylic acids as acids

Note: This assumes that the hydroxide is in solution. (It


probably will be, because almost the only reaction ever done
of this type is with sodium hydroxide solution!) If you were
using a solid hydroxide, however, you would have to swap
the hydroxide state symbol from "(aq)" to "(s)".

If you mix dilute ethanoic acid with sodium hydroxide solution, for
example, you simply get a colourless solution containing sodium
ethanoate. The only sign that a change has happened is that the
temperature of the mixture will have increased.
This change could well be represented by the ionic equation above, but if
you want it, the full equation for this particular reaction is:

With carbonates and hydrogencarbonates


In both of these cases, a salt is formed together with carbon dioxide and
water. Both are most easily represented by ionic equations.
For carbonates:

. . . and for hydrogencarbonates:

http://www.chemguide.co.uk/organicprops/acids/acidity.html (4 of 8)30/12/2004 11:25:52

carboxylic acids as acids

Note: This time, this assumes that the carbonate or


hydrogencarbonate will be a solid - which it often is. If you
used a solution of, say, sodium carbonate or
hydrogencarbonate, you would have to remember to change
the state symbol.

If you pour some dilute ethanoic acid onto some white sodium carbonate
or sodium hydrogencarbonate crystals, there is an immediate fizzing as
carbon dioxide is produced. You end up with a colourless solution of
sodium ethanoate.
With sodium carbonate, the full equation is:

. . . and for sodium hydrogencarbonate:

There is very little obvious difference in the vigour of these reactions


compared with the same reactions with dilute hydrochloric acid.
However, you would notice the difference if you used a slower reaction for example with calcium carbonate in the form of a marble chip. With
ethanoic acid, you would eventually produce a colourless solution of
calcium ethanoate.

In this case, the marble chip would react noticeably more slowly with
ethanoic acid than with hydrochloric acid.

With ammonia
Ethanoic acid reacts with ammonia in exactly the same way as any other
acid does. It transfers a hydrogen ion to the lone pair on the nitrogen of
the ammonia and forms an ammonium ion.
http://www.chemguide.co.uk/organicprops/acids/acidity.html (5 of 8)30/12/2004 11:25:52

carboxylic acids as acids

If you mix together a solution of ethanoic acid and a solution of ammonia,


you will get a colourless solution of ammonium ethanoate.

With amines
Amines are compounds in which one or more of the hydrogen atoms in
an ammonia molecule have been replaced by a hydrocarbon group such
as an alkyl group. For simplicity, we'll just look at compounds where only
one of the hydrogen atoms has been replaced. These are called primary
amines.
Simple primary amines include:

The small amines are very similar indeed to ammonia in many ways. For
example, they smell very much like ammonia and are just as soluble in
water. Because all you have done to an ammonia molecule is swap a
hydrogen for an alkyl group, the lone pair is still there on the nitrogen
atom.
That means that they will react with acids (including carboxylic acids) in
just the same way as ammonia does.

http://www.chemguide.co.uk/organicprops/acids/acidity.html (6 of 8)30/12/2004 11:25:52

carboxylic acids as acids

For example, ethanoic acid reacts with methylamine to produce a


colourless solution of the salt methylammonium ethanoate.

Note: I have deliberately not tried to write this as a one line


equation. It is much easier to see what is happening, to write
it, and to remember it, if you draw the structures.
Don't be put off by the name of the product methylammonium ethanoate. "Methylammonium" just means
an ammonium ion in which one of the hydrogens is replaced
by a methyl group.

However complicated the amine, because all of them have got a lone
pair on the nitrogen atom, you would get the same sort of reaction.

Note: If you want to find out more about amines you could
explore the amines menu by following this link.

Where would you like to go now?


To the carboxylic acids menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

http://www.chemguide.co.uk/organicprops/acids/acidity.html (7 of 8)30/12/2004 11:25:52

carboxylic acids as acids

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/acids/acidity.html (8 of 8)30/12/2004 11:25:52

esterification - alcohols and carboxylic acids

ESTERIFICATION OF CARBOXYLIC ACIDS

This page looks at esterification - the reaction between alcohols and


carboxylic acids to make esters.

What are esters?


Esters have a hydrocarbon group of some sort replacing the hydrogen in
the -COOH group of a carboxylic acid. We shall just be looking at cases
where it is replaced by an alkyl group, but it could equally well be an aryl
group (one based on a benzene ring).
A common ester - ethyl ethanoate
The most commonly discussed ester is ethyl ethanoate. In this case, the
hydrogen in the -COOH group has been replaced by an ethyl group. The
formula for ethyl ethanoate is:

Notice that the ester is named the opposite way around from the way the
formula is written. The "ethanoate" bit comes from ethanoic acid. The
"ethyl" bit comes from the ethyl group on the end.

Note: In my experience, students starting organic chemistry


get more confused about writing names and formulae for
esters than for almost anything else - particularly when it
comes to less frequently met esters like the ones coming up
next. Take time and care to make sure you understand!

http://www.chemguide.co.uk/organicprops/acids/esterification.html (1 of 6)30/12/2004 11:25:55

esterification - alcohols and carboxylic acids

A few more esters


In each case, be sure that you can see how the names and formulae
relate to each other.

Remember that the acid is named by counting up the total number of


carbon atoms in the chain - including the one in the -COOH group. So,
for example, CH3CH2COOH is propanoic acid, and CH3CH2COO is the
propanoate group.

Note: You can find more about naming acids and esters by
following this link to a different part of this site.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/acids/esterification.html (2 of 6)30/12/2004 11:25:55

esterification - alcohols and carboxylic acids

Making esters
The chemistry of the reaction
Esters are produced when carboxylic acids are heated with alcohols in
the presence of an acid catalyst. The catalyst is usually concentrated
sulphuric acid. Dry hydrogen chloride gas is used in some cases, but
these tend to involve aromatic esters (ones containing a benzene ring). If
you are a UK A level student you won't have to worry about these.
The esterification reaction is both slow and reversible. The equation for
the reaction between an acid RCOOH and an alcohol R'OH (where R
and R' can be the same or different) is:

So, for example, if you were making ethyl ethanoate from ethanoic acid
and ethanol, the equation would be:

Note: The mechanism for the esterification reaction is


covered in the catalysis section of this site. It is not required
for any UK A level chemistry syllabus.
If you follow this link, use the BACK button on your browser
to return to this page.

http://www.chemguide.co.uk/organicprops/acids/esterification.html (3 of 6)30/12/2004 11:25:55

esterification - alcohols and carboxylic acids

Doing the reactions


On a test tube scale
Carboxylic acids and alcohols are often warmed together in the presence
of a few drops of concentrated sulphuric acid in order to observe the
smell of the esters formed.
You would normally use small quantities of everything heated in a test
tube stood in a hot water bath for a couple of minutes.
Because the reactions are slow and reversible, you don't get a lot of
ester produced in this time. The smell is often masked or distorted by the
smell of the carboxylic acid. A simple way of detecting the smell of the
ester is to pour the mixture into some water in a small beaker.
Esters are virtually insoluble in water and tend to form a thin layer on the
surface. Excess acid and alcohol both dissolve and are tucked safely
away under the ester layer.
Small esters like ethyl ethanoate smell like typical organic solvents (ethyl
ethanoate is a common solvent in, for example, glues).
As the esters get bigger, the smells tend towards artificial fruit flavouring
- "pear drops", for example.
On a larger scale
If you want to make a reasonably large sample of an ester, the method
used depends to some extent on the size of the ester. Small esters are
formed faster than bigger ones.
To make a small ester like ethyl ethanoate, you can gently heat a mixture
of ethanoic acid and ethanol in the presence of concentrated sulphuric
acid, and distil off the ester as soon as it is formed.
This prevents the reverse reaction happening. It works well because the
ester has the lowest boiling point of anything present. The ester is the
only thing in the mixture which doesn't form hydrogen bonds, and so it
http://www.chemguide.co.uk/organicprops/acids/esterification.html (4 of 6)30/12/2004 11:25:55

esterification - alcohols and carboxylic acids

has the weakest intermolecular forces.

Note: Follow this link if you aren't sure about hydrogen


bonding.
Use the BACK button on your browser to return to this page.

Larger esters tend to form more slowly. In these cases, it may be


necessary to heat the reaction mixture under reflux for some time to
produce an equilibrium mixture. The ester can be separated from the
carboxylic acid, alcohol, water and sulphuric acid in the mixture by
fractional distillation.

Note: Providing full details for organic preparations


(including all the steps necessary in cleaning up the product)
is beyond the scope of this site. If you need this sort of detail,
you should be looking at an organic practical book.

Where would you like to go now?


To the carboxylic acids menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/acids/esterification.html (5 of 6)30/12/2004 11:25:55

esterification - alcohols and carboxylic acids

http://www.chemguide.co.uk/organicprops/acids/esterification.html (6 of 6)30/12/2004 11:25:55

reduction of carboxylic acids

REDUCTION OF CARBOXYLIC ACIDS

This page looks at the reduction of carboxylic acids to primary alcohols


using lithium tetrahydridoaluminate(III) (lithium aluminium hydride), LiAlH4.
The "(III)" is the oxidation state of the aluminium. Since aluminium only
ever shows the +3 oxidation state in its compounds, the "(III)" is actually
unnecessary. I shall leave it out for the rest of this page to make the name
a bit shorter.

Note: If you haven't come across this reducing agent before,


you might find it useful to read a rather more detailed account
of its use on the page about reduction of aldehydes and
ketones. The present page isn't done in the same sort of detail
as that one.
Use the BACK button on your browser to return to this page.

The reaction
The reducing agent
Lithium tetrahydridoaluminate has the structure:

In the negative ion, one of the bonds is a co-ordinate covalent (dative


http://www.chemguide.co.uk/organicprops/acids/reduction.html (1 of 4)30/12/2004 11:25:59

reduction of carboxylic acids

covalent) bond using the lone pair on a hydride ion (H-) to form a bond with
an empty orbital on the aluminium.

Note: If you aren't happy about co-ordinate covalent (dative


covalent) bonding you could follow this link, although it isn't
important to understanding the rest of this page.
Use the BACK button on your browser to return to this page.

The reduction of a carboxylic acid


The reaction happens in two stages - first to form an aldehyde and then a
primary alcohol. Because lithium tetrahydridoaluminate reacts rapidly with
aldehydes, it is impossible to stop at the halfway stage.
Equations for these reactions are usually written in a simplified form for UK
A level purposes. The "[H]" in the equations represents hydrogen from a
reducing agent.
Because of the impossibility of stopping at the aldehyde, there isn't much
point in giving an equation for the two separate stages. The overall
reaction is:

"R" is hydrogen or a hydrocarbon group. For example, ethanoic acid will


reduce to the primary alcohol, ethanol.

http://www.chemguide.co.uk/organicprops/acids/reduction.html (2 of 4)30/12/2004 11:25:59

reduction of carboxylic acids

Note: Follow this link if you aren't sure what a primary alcohol
is.
Use the BACK button on your browser to return to this page.

Sodium tetrahydridoborate (sodium borohydride) won't work!


If you are familiar with the reduction of aldehydes and ketones using
lithium tetrahydridoaluminate, you are probably aware that sodium
tetrahydridoborate is often used as a safer alternative.
It CAN'T be used with carboxylic acids. The sodium tetrahydridoborate isn't
reactive enough to reduce carboxylic acids.

Reaction conditions
Lithium tetrahydridoaluminate reacts violently with water and so the
reactions are carried out in solution in dry ethoxyethane (diethyl ether or
just "ether"). The reaction happens at room temperature.
At the end of the reaction, the product is a complex aluminium salt. This is
converted into the alcohol by treatment with dilute sulphuric acid.

Note: You are unlikely to need this equation for UK A level


purposes, although you may need to know that you have to
finish the reaction off using dilute sulphuric acid.

http://www.chemguide.co.uk/organicprops/acids/reduction.html (3 of 4)30/12/2004 11:25:59

reduction of carboxylic acids

Where would you like to go now?


To the carboxylic acids menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/acids/reduction.html (4 of 4)30/12/2004 11:25:59

converting carboxylic acids into acyl (acid) chlorides

CONVERTING CARBOXYLIC ACIDS INTO ACYL


CHLORIDES (ACID CHLORIDES)

This page looks at ways of swapping the -OH group in the -COOH group
of a carboxylic acid for a chlorine atom. This produces useful compounds
called acyl chlorides (acid chlorides).
It covers the use of phosphorus(V) chloride and phosphorus(III) chloride
as well as sulphur dichloride oxide (thionyl chloride).

Replacing -OH by -Cl


Introduction
We are going to be looking at converting a carboxylic acid, RCOOH, into
an acyl chloride, RCOCl. Acyl chlorides are also known as acid chlorides.

By far the most commonly used example of the conversion of a


carboxylic acid into an acyl chloride is ethanoic acid to ethanoyl chloride.

Acyl chlorides are very reactive, and can be used to make a wide range
of other things. That's why they are important.

http://www.chemguide.co.uk/organicprops/acids/pcl5.html (1 of 3)30/12/2004 11:26:03

converting carboxylic acids into acyl (acid) chlorides

Note: If you are want to find out more about acyl chlorides,
you can explore the acyl chlorides menu by following this link.
Use the BACK button on your browser to return to this page.

Replacing the -OH group using phosphorus(V) chloride, PCl5


Phosphorus(V) chloride is a solid which reacts with carboxylic acids in
the cold to give steamy acidic fumes of hydrogen chloride. It leaves a
liquid mixture of the acyl chloride and a phosphorus compound,
phosphorus trichloride oxide (phosphorus oxychloride) - POCl3.
The acyl chloride can be separated by fractional distillation.
For example:

Replacing the -OH group using phosphorus(III) chloride, PCl3


Phosphorus(III) chloride is a liquid at room temperature. Its reaction with
a carboxylic acid is less dramatic than that of phosphorus(V) chloride
because there is no hydrogen chloride produced.
You end up with a mixture of the acyl chloride and phosphoric(III) acid
(old names: phosphorous acid or orthophosphorous acid), H3PO3.
For example:

Again, the ethanoyl chloride can be separated by fractional distillation.

http://www.chemguide.co.uk/organicprops/acids/pcl5.html (2 of 3)30/12/2004 11:26:03

converting carboxylic acids into acyl (acid) chlorides

Replacing the -OH group using sulphur dichloride oxide (thionyl


chloride)
Sulphur dichloride oxide (thionyl chloride) is a liquid at room temperature
and has the formula SOCl2.
Traditionally, the formula is written as shown, despite the fact that the
modern name writes the chlorine before the oxygen (alphabetical order).
The sulphur dichloride oxide reacts with carboxylic acids to produce an
acyl chloride, and sulphur dioxide and hydrogen chloride gases are given
off.
For example:

The separation is simplified to an extent because the by-products are


both gases. You would obviously still have to fractionally distil the mixture
to separate the acyl chloride from any excess acid or sulphur dichloride
oxide.

Where would you like to go now?


To the carboxylic acids menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/acids/pcl5.html (3 of 3)30/12/2004 11:26:03

Acyl chlorides (acid chlorides) Menu

Understanding Chemistry

ACYL CHLORIDES (acid chlorides) MENU

Background . . .
An introduction to acyl chlorides including a general look at their
reactivity.
Preparation of acyl chlorides . . .
Their preparation from carboxylic acids.
Reactions of acyl chlorides involving oxygen compounds . . .
The reactions of acyl chlorides with water, alcohols and phenols.
Reactions of acyl chlorides involving nitrogen compounds . . .
The reactions of acyl chlorides with ammonia and primary amines.
Using acyl chlorides in Friedel-Crafts reactions . . .
The use of acyl chlorides in substituting an acyl group into a
benzene ring.

Go to menu of other organic compounds . . .


Go to Main Menu . . .

http://www.chemguide.co.uk/organicprops/acylclmenu.html (1 of 2)30/12/2004 11:26:04

Acyl chlorides (acid chlorides) Menu

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/acylclmenu.html (2 of 2)30/12/2004 11:26:04

an introduction to acyl chlorides (acid chlorides)

INTRODUCING ACYL CHLORIDES (acid


chlorides)

This page explains what acyl chlorides are and looks at their simple
physical properties such as boiling points. It introduces their chemical
reactivity in a general way, but details of specific reactions are given on
separate pages - see the acyl chlorides menu (link at the bottom of the
page).

What are acyl chlorides?


Acyl chlorides as "acid derivatives"
A carboxylic acid such as ethanoic acid has the structure:

There are a number of related compounds in which the -OH group in the
acid is replaced by something else. Compounds like this are described
as acid derivatives.
Acyl chlorides (also known as acid chlorides) are one example of an acid
derivative. In this case, the -OH group has been replaced by a chlorine
atom.

http://www.chemguide.co.uk/organicprops/acylchlorides/background.html (1 of 5)30/12/2004 11:26:08

an introduction to acyl chlorides (acid chlorides)

The acyl group


The acyl group is a hydrocarbon group attached to a carbon-oxygen
double bond:

For UK A level purposes, the "R" group is normally restricted to an alkyl


group. It could, however, equally well be a group based on a benzene
ring.

Naming acyl chlorides


The easiest way of thinking about the names is to see the relationship
with the corresponding carboxylic acid:

carboxylic acid name

acyl chloride name

acyl chloride formula

ethanoic acid

ethanoyl chloride

CH3COCl

propanoic acid

propanoyl chloride

CH3CH2COCl

butanoic acid

butanoyl chloride

CH3CH2CH2COCl

If you have something substituted into the hydrocarbon chain, the carbon
in the -COCl group counts as the number 1 carbon.
For example, 2-methylbutanoyl chloride is:

http://www.chemguide.co.uk/organicprops/acylchlorides/background.html (2 of 5)30/12/2004 11:26:08

an introduction to acyl chlorides (acid chlorides)

Note: Hardly anyone ever mentions methanoyl chloride,


HCOCl - derived from methanoic acid. That is because
methanoyl chloride is very unstable, decomposing to give
carbon monoxide and HCl.

Physical properties of acyl chlorides


Appearance
An acyl chloride like ethanoyl chloride is a colourless fuming liquid. The
strong smell of ethanoyl chloride is a mixture of the smell of vinegar
(ethanoic acid) and the acrid smell of hydrogen chloride gas.
The smell and the fumes are because ethanoyl chloride reacts with water
vapour in the air. The reaction with water is given in detail on another
page. (Find it from the acyl chlorides menu - link at the bottom of this
page.)

Solubility in water
Acyl chlorides can't be said to dissolve in water because they react (often
violently) with it. The strong reaction means that it is impossible to get a
simple aqueous solution of an acyl chloride.

Boiling points

http://www.chemguide.co.uk/organicprops/acylchlorides/background.html (3 of 5)30/12/2004 11:26:08

an introduction to acyl chlorides (acid chlorides)

Taking ethanoyl chloride as typical:


Ethanoyl chloride boils at 51C. It is a polar molecule, and so has dipoledipole attractions between its molecules as well as van der Waals
dispersion forces.
However, it doesn't form hydrogen bonds. Its boiling point is therefore
higher than, say, an alkane of similar size (which has no permanent
dipoles), but not as high as a similarly sized alcohol (which forms
hydrogen bonds in addition to everything else.)

Note: If you aren't happy about intermolecular forces


(including van der Waals dispersion forces and hydrogen
bonds) then you really ought to follow this link before you go
on.
Use the BACK button on your browser to return to this page.

Reactivity of acyl chlorides


Substitution of the chlorine atom by other groups
Acyl chlorides are extremely reactive, and in their reactions the chlorine
atom is replaced by other things.
In each case, in the first instance, hydrogen chloride gas is produced as
steamy acidic fumes. However, in some cases the hydrogen chloride
goes on to react with one of the substances in the reaction mixture.
Taking ethanoyl chloride as typical, the initial reaction is of this kind:

http://www.chemguide.co.uk/organicprops/acylchlorides/background.html (4 of 5)30/12/2004 11:26:08

an introduction to acyl chlorides (acid chlorides)

The reactions involve things like water, alcohols and phenols, or


ammonia and amines. All of these particular cases contain a very
electronegative element with an active lone pair of electrons - either
oxygen or nitrogen.

Note: You can find details of all these reactions from the
acyl chlorides menu (link below).
If you are interested in exploring the general mechanism for
these reactions, you will find it by following this link to another
part of the site dealing with nucleophilic addition-elimination
reactions. If you want mechanisms for specific reactions you
could explore other pages from the nucleophilic additionelimination menu as well - but read the general mechanism
first.

Where would you like to go now?


To the acyl chlorides menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/acylchlorides/background.html (5 of 5)30/12/2004 11:26:08

What is nucleophilic addition / elimination?

NUCLEOPHILIC ADDITION / ELIMINATION IN


THE REACTIONS OF ACYL CHLORIDES

Background
What are acyl chlorides?
Acyl chlorides (also known as acid chlorides) are one of a number of
types of compounds known as "acid derivatives". This is ethanoic acid:

If you remove the -OH group and replace it by a -Cl, you have produced
an acyl chloride.

This molecule is known as ethanoyl chloride and for the rest of this topic
will be taken as typical of acyl chlorides in general.
Acyl chlorides are extremely reactive. They are open to attack by
nucleophiles - with the overall result being a replacement of the chlorine
by something else.
Nucleophiles
A nucleophile is a species (an ion or a molecule) which is strongly
attracted to a region of positive charge in something else.
Nucleophiles are either fully negative ions, or else have a strongly charge somewhere on a molecule. The nucleophiles that we shall be
looking at all depend on lone pairs on either an oxygen atom or a
http://www.chemguide.co.uk/mechanisms/addelim/whatis.html (1 of 6)30/12/2004 11:26:15

What is nucleophilic addition / elimination?

nitrogen atom.
Nucleophiles based on oxygen atoms
We shall be talking about water and alcohols, taking ethanol as a typical
alcohol.

Notice how similar these two molecules are around the oxygen atom.
That's what turns out to be important.
Nucleophiles based on nitrogen atoms
We shall be considering ammonia and primary amines, taking ethylamine
as a typical primary amine. A primary amine contains the -NH2 group
attached to either an alkyl group (as it is here) or a benzene ring. As far
as these reactions are concerned, the nature of any hydrocarbon
attached to the nitrogen makes no difference. The nitrogen atom is the
important bit.

Again, notice how similar these two molecules are around the nitrogen
atom - and also how similar they are to the previous ones containing
oxygen. Both types of molecule have an active lone pair of electrons
attached to one of the most electronegative elements. All of these
molecules also have at least one hydrogen atom attached to the oxygen
or nitrogen.
Why are acyl chlorides attacked by nucleophiles?
The carbon atom in the -COCl group has both an oxygen atom and a
chlorine atom attached to it. Both of these are very electronegative. They
both pull electrons towards themselves, leaving the carbon atom quite
positively charged.

http://www.chemguide.co.uk/mechanisms/addelim/whatis.html (2 of 6)30/12/2004 11:26:15

What is nucleophilic addition / elimination?

Note: If you aren't sure about electronegativity and bond


polarity follow this link before you go on. Use the BACK
button on your browser to return to this page.

The overall reaction


We are going to generalise this for the moment by writing the reacting
molecule as "Nu-H". Nu is the bit of the molecule which contains the
nucleophilic oxygen or nitrogen atom. The attached hydrogen turns out to
be essential to the reaction.
The general equation for the reaction is:

In each case, the net effect is that you replace the -Cl by -Nu, and
hydrogen chloride is formed as well.
Since the initial attack is by a nucleophile, and the overall result is
substitution, it would seem reasonable to describe the reaction as
nucleophilic substitution. However, the reaction happens in two distinct
stages. The first involves an addition reaction, which is followed by an
elimination reaction where HCl is produced. So the mechanism is also
known as nucleophilic addition / elimination.

http://www.chemguide.co.uk/mechanisms/addelim/whatis.html (3 of 6)30/12/2004 11:26:15

What is nucleophilic addition / elimination?

Note: You will find both terms in use - and to confuse the
issue still further, these are also examples of condensation
reactions. A condensation reaction is one in which two
molecules join together to make a bigger one, and in the
process shed a little molecule.
The only Exam Board to require these mechanisms (AQA)
calls them addition / elimination reactions.

The general mechanism


The addition stage of the mechanism
As the lone pair on the nucleophile approaches the fairly positive carbon
in the acyl chloride, it moves to form a bond with it. In the process, the
two electrons in one of the carbon-oxygen bonds are repelled entirely
onto the oxygen, leaving that oxygen negatively charged.

Note: If you aren't happy about the use of curly arrows in


mechanisms, follow this link before you go on. Use the BACK
button on your browser to return to this page.

http://www.chemguide.co.uk/mechanisms/addelim/whatis.html (4 of 6)30/12/2004 11:26:15

What is nucleophilic addition / elimination?

Notice the positive charge that forms on the nucleophile. Just accept this
for the moment. It's much easier to explain why that charge must be
there if you have a real example in front of you. This is fully explained in
the pages on the reactions involving water, ammonia and so on.
The elimination stage of the mechanism
This happens in two steps. In the first step, the carbon-oxygen double
bond reforms. To make room for it, the electrons in the carbon-chlorine
bond are repelled until they are entirely on the chlorine atom - forming a
chloride ion.

Finally, the chloride ion plucks the hydrogen off the original nucleophile.
It removes it as a hydrogen ion, leaving the pair of electrons behind on
the oxygen or nitrogen atom in that nucleophile. That cancels the positive
charge.

http://www.chemguide.co.uk/mechanisms/addelim/whatis.html (5 of 6)30/12/2004 11:26:15

What is nucleophilic addition / elimination?

Important! Take some trouble to understand this general


mechanism. When you read the pages on the specific
reactions you should then be able to see all the similarities
which exist - and realise that they are all essentially the same
mechanism. If you try to learn them all as separate
mechanisms, you are in for a lot of tedium!

Where would you like to go now?


To menu of nucleophilic addition / elimination reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/addelim/whatis.html (6 of 6)30/12/2004 11:26:15

Nucleophilic Addition / Eliminination Mechanisms Menu

Understanding Chemistry

NUCLEOPHILIC ADDITION / ELIMINATION


MECHANISMS MENU

What is nucleophilic addition / elimination? . . .


Includes an explanation of all the terms involved, together with a
general mechanism for these reactions. It would be useful to read
this before you look at specific examples.
The reaction of acyl chlorides with water . . .
Facts and mechanism for the formation of acids from acyl
chlorides by reaction with water.
The reaction of acyl chlorides with alcohols . . .
Facts and mechanism for the formation of esters from acyl
chlorides by reaction with alcohols.
The reaction of acyl chlorides with ammonia . . .
Facts and mechanism for the formation of amides from acyl
chlorides by reaction with ammonia.
The reaction of acyl chlorides with primary amines . . .
Facts and mechanism for the formation of N-substituted amides
from acyl chlorides by reaction with primary amines.

http://www.chemguide.co.uk/mechanisms/addelimmenu.html (1 of 2)30/12/2004 11:26:16

Nucleophilic Addition / Eliminination Mechanisms Menu

Go to menu of other types of mechanism. . .


Go to Main Menu . . .

You might also be interested in:


properties and reactions of acyl chlorides . . .
Covers all the physical and chemical properties of acyl chlorides
(acid chlorides) required by UK A level syllabuses.

Jim Clark 2000 (modified 2004)

http://www.chemguide.co.uk/mechanisms/addelimmenu.html (2 of 2)30/12/2004 11:26:16

reaction between acyl chlorides and water - addition / elimination

NUCLEOPHILIC ADDITION / ELIMINATION IN


THE REACTION BETWEEN ACYL CHLORIDES
AND WATER

This page gives you the facts and a simple, uncluttered mechanism for
the nucleophilic addition / elimination reaction between acyl chlorides
(acid chlorides) and water. If you want the mechanism explained to you
in detail, there is a link at the bottom of the page.
Ethanoyl chloride is taken as a typical acyl chloride. Any other acyl
chloride will behave in the same way. Simply replace the CH3 group in
what follows by anything else you want.

The reaction between ethanoyl chloride and water


The facts
Ethanoyl chloride reacts instantly with cold water. There is a very
exothermic reaction in which a steamy acidic gas is given off (hydrogen
chloride) and ethanoic acid is formed.

The mechanism
The first stage (the addition stage of the reaction) involves a nucleophilic
attack on the fairly positive carbon atom by one of the lone pairs on the
oxygen of a water molecule.

http://www.chemguide.co.uk/mechanisms/addelim/water.html (1 of 3)30/12/2004 11:26:20

reaction between acyl chlorides and water - addition / elimination

Note: Only one of the two lone pairs on the oxygen in water
is shown. This is to avoid cluttering an already complicated
diagram with things that aren't relevant.

The second stage (the elimination stage) happens in two steps. In the
first, the carbon-oxygen double bond reforms and a chloride ion is
pushed off.

That is followed by removal of a hydrogen ion by the chloride ion to give


ethanoic acid and hydrogen chloride.

Where would you like to go now?


http://www.chemguide.co.uk/mechanisms/addelim/water.html (2 of 3)30/12/2004 11:26:20

reaction between acyl chlorides and water - addition / elimination

Help! Talk me through this mechanism . . .


To menu of nucleophilic addition / elimination reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/addelim/water.html (3 of 3)30/12/2004 11:26:20

Explaining the reaction between acyl chlorides and water - addition / elimination

EXPLAINING NUCLEOPHILIC ADDITION /


ELIMINATION IN THE REACTION BETWEEN
ACYL CHLORIDES AND WATER

This page guides you through the mechanism for the nucleophilic
addition / elimination reaction between acyl chlorides (acid chlorides) and
water.
Ethanoyl chloride is taken as a typical acyl chloride. Any other acyl
chloride will behave in the same way. Simply replace the CH3 group in
what follows by anything else you want.

Important! If you haven't already done so, it would help if


you first read the page What is nucleophilic addition /
elimination? before you go on.

The reaction between ethanoyl chloride and water


Water as a nucleophile
A nucleophile is a species (an ion or a molecule) which is strongly
attracted to a region of positive charge in something else.
Nucleophiles are either fully negative ions, or else have
a strongly - charge somewhere on a molecule.
Oxygen is much more electronegative than hydrogen
and so drags the electrons in the oxygen-hydrogen
bonds towards itself. That produces a significant
amount of negative charge on the oxygen atom. The oxygen also has
two active lone pairs of electrons. One of these attacks the ethanoyl
chloride.

http://www.chemguide.co.uk/mechanisms/addelim/watertt.html (1 of 4)30/12/2004 11:26:23

Explaining the reaction between acyl chlorides and water - addition / elimination

Note: If you aren't sure about electronegativity and bond


polarity follow this link before you go on. Use the BACK
button on your browser to return to this page.

Why are acyl chlorides attacked by nucleophiles?


The carbon atom in the -COCl group has both an oxygen atom and a
chlorine atom attached to it. Both of these are very electronegative. They
both pull electrons towards themselves, leaving the carbon atom quite
positively charged. It is that carbon atom which is attacked by one of the
lone pairs on the oxygen atom in a water molecule.
The ethanoyl chloride molecule is also planar (flat) around that carbon
atom, and that leaves plenty of room for a nucleophile to attack either
from above or below the plane of the molecule.

The mechanism
The reaction happens in two main stages - an addition stage, followed by
an elimination stage.
In the addition stage, a water molecule becomes attached to the carbon
in the ethanoyl chloride.
As the lone pair on the water approaches the fairly positive carbon in the
ethanoyl chloride, it moves to form a bond with it. In the process, the two
electrons in one of the carbon-oxygen bonds are repelled entirely onto
the oxygen, leaving that oxygen negatively charged.

http://www.chemguide.co.uk/mechanisms/addelim/watertt.html (2 of 4)30/12/2004 11:26:23

Explaining the reaction between acyl chlorides and water - addition / elimination

Note: Only one of the two lone pairs on the oxygen in water
is shown because only one of them is involved in the
reaction. There's no reason why you couldn't write both in if
you want to - it just clutters the diagram.

Notice that the oxygen atom in the water has gained a positive charge.
The underlying reason for this is that when the lone pair forms a bond
with the carbon, electrons are moving away from the oxygen. What
matters, though, is that you remember to show the positive charge in an
exam. Think of it like this:
If you leave out the positive charge, two things are wrong with the
equation:

The charges don't balance. You start with two overall neutral
molecules and, if you forgot the positive charge, you would end up
with a negative ion. There has got to be a positive charge
somewhere to balance the negative one.
The oxygen looks wrong! Oxygen normally forms two bonds, but
here it is forming three. Oxygen can only form three bonds if it
carries a positive charge. (A positively charged oxygen atom has
the same electronic structure as nitrogen - which normally forms
three bonds.)

You can put both things right with a positive charge on the oxygen.

http://www.chemguide.co.uk/mechanisms/addelim/watertt.html (3 of 4)30/12/2004 11:26:23

Explaining the reaction between acyl chlorides and water - addition / elimination

The elimination stage stage of the reaction happens in two steps. In the
first, the carbon-oxygen double bond reforms. As the electron pair moves
back to form a bond with the carbon, the pair of electrons in the carbonchlorine bond are forced entirely onto the chlorine to give a chloride ion.

Finally, the chloride ion forms a bond with one of the hydrogens on the
oxygen - removing it as a hydrogen ion. The electrons in the hydrogenoxgyen bond move back onto the oxygen, cancelling the positive charge.

Where would you like to go now?


To menu of nucleophilic addition / elimination reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/addelim/watertt.html (4 of 4)30/12/2004 11:26:23

reaction between acyl chlorides and alcohols - addition / elimination

NUCLEOPHILIC ADDITION / ELIMINATION IN THE


REACTION BETWEEN ACYL CHLORIDES AND
ALCOHOLS

This page gives you the facts and a simple, uncluttered mechanism for the
nucleophilic addition / elimination reaction between acyl chlorides (acid
chlorides) and alcohols. If you want the mechanism explained to you in
detail, there is a link at the bottom of the page.
Ethanoyl chloride is taken as a typical acyl chloride. Any other acyl chloride
will behave in the same way. Simply replace the CH3 group in what follows
by anything else you want. Similarly, ethanol is taken as a typical alcohol. If
you are interested in another alcohol, you can replace the CH3CH2 group
by any other alkyl group.

The reaction between ethanoyl chloride and ethanol


The facts
Ethanoyl chloride reacts instantly with cold ethanol. There is a very
exothermic reaction in which a steamy acidic gas is given off (hydrogen
chloride). Ethyl ethanoate (an ester) is formed.

The mechanism
The first stage (the addition stage of the reaction) involves a nucleophilic
attack on the fairly positive carbon atom by one of the lone pairs on the
oxygen of an ethanol molecule.

http://www.chemguide.co.uk/mechanisms/addelim/alcohol.html (1 of 3)30/12/2004 11:26:27

reaction between acyl chlorides and alcohols - addition / elimination

Note: Only one of the two lone pairs on the oxygen in the
ethanol is shown. This is to avoid cluttering an already
complicated diagram with things that aren't relevant.

The second stage (the elimination stage) happens in two steps. In the first,
the carbon-oxygen double bond reforms and a chloride ion is pushed off.

That is followed by removal of a hydrogen ion by the chloride ion to give


ethyl ethanoate and hydrogen chloride.

Where would you like to go now?

http://www.chemguide.co.uk/mechanisms/addelim/alcohol.html (2 of 3)30/12/2004 11:26:27

reaction between acyl chlorides and alcohols - addition / elimination

Help! Talk me through this mechanism . . .


To menu of nucleophilic addition / elimination reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/addelim/alcohol.html (3 of 3)30/12/2004 11:26:27

Explaining the reaction between acyl chlorides and alcohols - addition / elimination

EXPLAINING NUCLEOPHILIC ADDITION /


ELIMINATION IN THE REACTION BETWEEN
ACYL CHLORIDES AND ALCOHOLS

This page guides you through the mechanism for the nucleophilic
addition / elimination reaction between acyl chlorides (acid chlorides) and
alcohols.
Ethanoyl chloride is taken as a typical acyl chloride. Any other acyl
chloride will behave in the same way. Ethanol is taken as a typical
alcohol. Again, any other alcohol will behave in the same way.

Important! If you haven't already done so, it would help if


you first read the page What is nucleophilic addition /
elimination? before you go on.
The reaction between ethanoyl chloride and ethanol has a
mechanism which is identical to that between ethanoyl
chloride and water, but it looks more complicated. It would be
a good idea to be sure about the water mechanism before
you attack this one!

The reaction between ethanoyl chloride and ethanol


Ethanol as a nucleophile
A nucleophile is a species (an ion or a molecule) which is strongly
attracted to a region of positive charge in something else.

http://www.chemguide.co.uk/mechanisms/addelim/alcoholtt.html (1 of 5)30/12/2004 11:26:32

Explaining the reaction between acyl chlorides and alcohols - addition / elimination

Nucleophiles are either fully negative ions, or


else have a strongly - charge somewhere on a
molecule.
Oxygen is much more electronegative than
hydrogen and carbon and so drags bonding electrons towards itself. That
produces a significant amount of negative charge on the oxygen atom.
The oxygen also has two active lone pairs of electrons. One of these
attacks the ethanoyl chloride.
Why are acyl chlorides attacked by nucleophiles?
The carbon atom in the -COCl group has both an oxygen atom and a
chlorine atom attached to it. Both of these are very electronegative. They
both pull electrons towards themselves, leaving the carbon atom quite
positively charged. It is that carbon atom which is attacked by one of the
lone pairs on the oxygen atom in an ethanol molecule.
The ethanoyl chloride molecule is also planar (flat) around that carbon
atom, and that leaves plenty of room for a nucleophile to attack either
from above or below the plane of the molecule.

The mechanism
The reaction happens in two main stages - an addition stage, followed by
an elimination stage.
In the addition stage, an ethanol molecule becomes attached to the
carbon in the ethanoyl chloride.
As the lone pair on the oxygen approaches the fairly positive carbon in
the ethanoyl chloride, it moves to form a bond with it. In the process, the
two electrons in one of the carbon-oxygen bonds are repelled entirely
http://www.chemguide.co.uk/mechanisms/addelim/alcoholtt.html (2 of 5)30/12/2004 11:26:32

Explaining the reaction between acyl chlorides and alcohols - addition / elimination

onto that oxygen, leaving it negatively charged.

Note: Only one of the two lone pairs on the oxygen in the
ethanol is shown because only one of them is involved in the
reaction. There's no reason why you couldn't write both in if
you want to - it just clutters the diagram.

Notice that the oxygen atom in the ethanol has gained a positive charge.
The underlying reason for this is that when the lone pair forms a bond
with the carbon, electrons are moving away from the oxygen. What
matters, though, is that you remember to show the positive charge in an
exam. Think of it like this:
If you leave out the positive charge, two things are wrong with the
equation:

The charges don't balance. You start with two overall neutral
molecules and, if you forgot the positive charge, you would end up
with a negative ion. There has got to be a positive charge
somewhere to balance the negative one.
The oxygen looks wrong! Oxygen normally forms two bonds, but
here it is forming three. Oxygen can only form three bonds if it
carries a positive charge. (A positively charged oxygen atom has
the same electronic structure as nitrogen - which normally forms
three bonds.)

You can put both things right with a positive charge on the oxygen.

http://www.chemguide.co.uk/mechanisms/addelim/alcoholtt.html (3 of 5)30/12/2004 11:26:32

Explaining the reaction between acyl chlorides and alcohols - addition / elimination

The elimination stage stage of the reaction happens in two steps. In the
first, the carbon-oxygen double bond reforms. As the electron pair moves
back to form a bond with the carbon, the pair of electrons in the carbonchlorine bond are forced entirely onto the chlorine to give a chloride ion.

Finally, the chloride ion forms a bond with the hydrogen on the oxygen removing it as a hydrogen ion. The electrons in the hydrogen-oxgyen
bond move back onto the oxygen, cancelling the positive charge.

Where would you like to go now?


To menu of nucleophilic addition / elimination reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

http://www.chemguide.co.uk/mechanisms/addelim/alcoholtt.html (4 of 5)30/12/2004 11:26:32

Explaining the reaction between acyl chlorides and alcohols - addition / elimination

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/addelim/alcoholtt.html (5 of 5)30/12/2004 11:26:32

reaction between acyl chlorides and ammonia - addition / elimination

NUCLEOPHILIC ADDITION / ELIMINATION IN


THE REACTION BETWEEN ACYL CHLORIDES
AND AMMONIA

This page gives you the facts and a simple, uncluttered mechanism for
the nucleophilic addition / elimination reaction between acyl chlorides
(acid chlorides) and ammonia. If you want the mechanism explained to
you in detail, there is a link at the bottom of the page.
Ethanoyl chloride is taken as a typical acyl chloride. Any other acyl
chloride will behave in the same way. Simply replace the CH3 group in
what follows by anything else you want.

The reaction between ethanoyl chloride and ammonia


The facts
Ethanoyl chloride reacts violently with a cold concentrated solution of
ammonia. A white solid product is formed which is a mixture of
ethanamide (an amide) and ammonium chloride.

Notice that, unlike the reactions between ethanoyl chloride and water or
ethanol, hydrogen chloride isn't produced - at least, not in any quantity.
Any hydrogen chloride formed would immediately react with excess
ammonia to give ammonium chloride.

The mechanism
The first stage (the addition stage of the reaction) involves a nucleophilic
attack on the fairly positive carbon atom by the lone pair on the nitrogen
http://www.chemguide.co.uk/mechanisms/addelim/ammonia.html (1 of 3)30/12/2004 11:26:45

reaction between acyl chlorides and ammonia - addition / elimination

atom in the ammonia.

The second stage (the elimination stage) happens in two steps. In the
first, the carbon-oxygen double bond reforms and a chloride ion is
pushed off.

That is followed by removal of a hydrogen ion from the nitrogen. This


might happen in one of two ways:
It might be removed by a chloride ion, producing HCl (which would
immediately react with excess ammonia to give ammonium chloride as
above) . . .

and

http://www.chemguide.co.uk/mechanisms/addelim/ammonia.html (2 of 3)30/12/2004 11:26:45

reaction between acyl chlorides and ammonia - addition / elimination

. . . or it might be removed directly by an ammonia molecule.

The ammonium ion, together with the chloride ion already there, makes
up the ammonium chloride formed in the reaction.

Where would you like to go now?


Help! Talk me through this mechanism . . .
To menu of nucleophilic addition / elimination reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/addelim/ammonia.html (3 of 3)30/12/2004 11:26:45

Explaining the reaction between acyl chlorides and ammonia - addition / elimination

EXPLAINING NUCLEOPHILIC ADDITION /


ELIMINATION IN THE REACTION BETWEEN
ACYL CHLORIDES AND AMMONIA

This page guides you through the mechanism for the nucleophilic
addition / elimination reaction between acyl chlorides (acid chlorides) and
ammonia.
Ethanoyl chloride is taken as a typical acyl chloride. Any other acyl
chloride will behave in the same way.

Important! If you haven't already done so, it would help if


you first read the page What is nucleophilic addition /
elimination? before you go on.

The reaction between ethanoyl chloride and ammonia


Ammonia as a nucleophile
A nucleophile is a species (an ion or a molecule) which is strongly
attracted to a region of positive charge in something else.
Nucleophiles are either fully negative ions, or else
have a strongly - charge somewhere on a
molecule.
Nitrogen is more electronegative than hydrogen
and so drags the bonding electrons towards itself.
That produces a significant amount of negative
charge on the nitrogen atom. The nitrogen also has an active lone pair of
electrons. It is this which attacks the ethanoyl chloride.
Why are acyl chlorides attacked by nucleophiles?

http://www.chemguide.co.uk/mechanisms/addelim/ammoniatt.html (1 of 5)30/12/2004 11:26:48

Explaining the reaction between acyl chlorides and ammonia - addition / elimination

The carbon atom in the -COCl group has both an oxygen atom and a
chlorine atom attached to it. Both of these are very electronegative. They
both pull electrons towards themselves, leaving the carbon atom quite
positively charged. It is that carbon atom which is attacked by the lone
pair on the nitrogen atom in an ammonia molecule.
The ethanoyl chloride molecule is also planar (flat) around that carbon
atom, and that leaves plenty of room for a nucleophile to attack either
from above or below the plane of the molecule.

The mechanism
The reaction happens in two main stages - an addition stage, followed by
an elimination stage.
In the addition stage, an ammonia molecule becomes attached to the
carbon in the ethanoyl chloride.
As the lone pair on the nitrogen approaches the fairly positive carbon in
the ethanoyl chloride, it moves to form a bond with it. In the process, the
two electrons in one of the carbon-oxygen bonds are repelled entirely
onto the oxygen, leaving it negatively charged.

Notice that the nitrogen atom has gained a positive charge. The
http://www.chemguide.co.uk/mechanisms/addelim/ammoniatt.html (2 of 5)30/12/2004 11:26:48

Explaining the reaction between acyl chlorides and ammonia - addition / elimination

underlying reason for this is that when the lone pair forms a bond with
the carbon, electrons are moving away from the nitrogen. What matters,
though, is that you remember to show the positive charge in an exam.
Think of it like this:
If you leave out the positive charge, two things are wrong with the
equation:

The charges don't balance. You start with two overall neutral
molecules and, if you forgot the positive charge, you would end up
with a negative ion. There has got to be a positive charge
somewhere to balance the negative one.
The nitrogen looks wrong! Nitrogen normally forms three bonds,
but here it is forming four. Nitrogen can only form four bonds if it
carries a positive charge. (A positively charged nitrogen atom has
the same electronic structure as carbon - which normally forms
four bonds.)

You can put both things right with a positive charge on the nitrogen.

The elimination stage stage of the reaction happens in two steps. In the
first, the carbon-oxygen double bond reforms. As the electron pair moves
back to form a bond with the carbon, the pair of electrons in the carbonchlorine bond are forced entirely onto the chlorine to give a chloride ion.

Finally, one of the hydrogens attached to the nitrogen is removed as a


hydrogen ion. The electrons in the hydrogen-nitrogen bond move back
onto the nitrogen, cancelling the positive charge.
The hydrogen ion might be removed in one of two ways. The first way is

http://www.chemguide.co.uk/mechanisms/addelim/ammoniatt.html (3 of 5)30/12/2004 11:26:48

Explaining the reaction between acyl chlorides and ammonia - addition / elimination

entirely consistent with what happens in the reactions between water or


ethanol and acyl chlorides. The hydrogen is removed by the chloride ion.

The hydrogen chloride produced would at once react with any excess
ammonia present to form ammonium chloride.

The other (more likely) possibility is that the hydrogen ion gets removed
directly by an ammonia molecule. This forms an ammonium ion. The
ammonium ion together with the chloride ion formed in the previous
stage makes up the ammonium chloride produced in the reaction.

http://www.chemguide.co.uk/mechanisms/addelim/ammoniatt.html (4 of 5)30/12/2004 11:26:48

Explaining the reaction between acyl chlorides and ammonia - addition / elimination

Note: You probably don't really need to know both of these


routes for the removal of the hydrogen ion. Choose for
yourself. It doesn't matter very much, because both will
happen.
The route involving the formation of hydrogen chloride is
exactly in line with the water or alcohol mechanisms, but you
do need to remember to react the HCl with excess ammonia.
The removal by another ammonia molecule is probably the
major route as long as the ammonia is in excess - which it
almost certainly will be. The reaction is normally done in the
lab by dropping the ethanoyl chloride into concentrated
ammonia solution.
Don't worry too much about this. Most sources opt out of the
problem altogether by quoting the final step simply as loss of
hydrogen ion: " - H+ ", without making any attempt to explain
what removes it.

Where would you like to go now?


To menu of nucleophilic addition / elimination reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/addelim/ammoniatt.html (5 of 5)30/12/2004 11:26:48

reaction between acyl chlorides and amines - addition / elimination

NUCLEOPHILIC ADDITION / ELIMINATION IN


THE REACTION BETWEEN ACYL CHLORIDES
AND AMINES

This page gives you the facts and a simple, uncluttered mechanism for
the nucleophilic addition / elimination reaction between acyl chlorides
(acid chlorides) and amines. If you want the mechanism explained to you
in detail, there is a link at the bottom of the page.
Ethanoyl chloride is taken as a typical acyl chloride. Any other acyl
chloride will behave in the same way. Simply replace the CH3 group in
what follows by anything else you want.
Similarly, ethylamine is taken as a typical amine. Any other amine will
behave in the same way. Replacing the CH3CH2 group by any other
hydrocarbon group won't affect the mechanism in any way.

The reaction between ethanoyl chloride and ethylamine


The facts
Ethanoyl chloride reacts violently with a cold concentrated solution of
ethylamine. A white solid product is formed which is a mixture of Nethylethanamide (an N-substituted amide) and ethylammonium chloride.

Notice that, unlike the reactions between ethanoyl chloride and water or
ethanol, hydrogen chloride isn't produced - at least, not in any quantity.
Any hydrogen chloride formed would immediately react with excess

http://www.chemguide.co.uk/mechanisms/addelim/amines.html (1 of 4)30/12/2004 11:26:53

reaction between acyl chlorides and amines - addition / elimination

ethylamine to give ethylammonium chloride.

The mechanism
The first stage (the addition stage of the reaction) involves a nucleophilic
attack on the fairly positive carbon atom by the lone pair on the nitrogen
atom in the ethylamine.

The second stage (the elimination stage) happens in two steps. In the
first, the carbon-oxygen double bond reforms and a chloride ion is
pushed off.

That is followed by removal of a hydrogen ion from the nitrogen. This


might happen in one of two ways:
It might be removed by a chloride ion, producing HCl (which would
immediately react with excess ethylamine to give ethylammonium
chloride as above) . . .

http://www.chemguide.co.uk/mechanisms/addelim/amines.html (2 of 4)30/12/2004 11:26:53

reaction between acyl chlorides and amines - addition / elimination

and

. . . or it might be removed directly by an ethylamine molecule.

The ethylammonium ion, together with the chloride ion already there,
makes up the ethylammonium chloride formed in the reaction.

Where would you like to go now?


Help! Talk me through this mechanism . . .
To menu of nucleophilic addition / elimination reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

http://www.chemguide.co.uk/mechanisms/addelim/amines.html (3 of 4)30/12/2004 11:26:53

reaction between acyl chlorides and amines - addition / elimination

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/addelim/amines.html (4 of 4)30/12/2004 11:26:53

Explaining the reaction between acyl chlorides and amines - addition / elimination

EXPLAINING NUCLEOPHILIC ADDITION /


ELIMINATION IN THE REACTION BETWEEN
ACYL CHLORIDES AND AMINES

This page guides you through the mechanism for the nucleophilic
addition / elimination reaction between acyl chlorides (acid chlorides) and
amines.
Ethanoyl chloride is taken as a typical acyl chloride. Similarly, ethylamine
is taken as a typical amine. Changing either the acyl choride or the
amine won't affect the mechanism in any way.

Important! If you haven't already done so, it would help if


you first read the page What is nucleophilic addition /
elimination? before you go on.
This mechanism also looks extremely complicated! In fact, it
isn't any more difficult than the mechanism for the reaction
between acyl chlorides and ammonia. Make sure you
understand that mechanism before you go any further with
this one.

The reaction between ethanoyl chloride and ethylamine


The products of the reaction
Remember that the reaction produces N-ethylethanamide and
ethylammonium chloride.

http://www.chemguide.co.uk/mechanisms/addelim/aminestt.html (1 of 6)30/12/2004 11:26:58

Explaining the reaction between acyl chlorides and amines - addition / elimination

N-ethylethanamide is an N-substituted amide. Ethanamide (a simple


amide) has the formula CH3CONH2. In an N-substituted amide, one of
the hydrogens on the nitrogen has been substituted by a hydrocarbon
group - which may be either an alkyl group (as here), or a benzene ring.
In the case we are discussing, an ethyl group has replaced one of the
hydrogens on the nitrogen - hence N-ethylethanamide.
The other product is ethylammonium chloride. Ammonium chloride is
NH4Cl. In ethylammonium chloride, one of the hydrogens on the nitrogen
has been substituted by an ethyl group.
Ethylamine as a nucleophile
A nucleophile is a species (an ion or a molecule) which is strongly
attracted to a region of positive charge in something else.
Nucleophiles are either fully negative ions,
or else have a strongly - charge
somewhere on a molecule.
Nitrogen is more electronegative than both
hydrogen and carbon and so drags the
bonding electrons towards itself. That
produces a significant amount of negative charge on the nitrogen atom.
The nitrogen also has an active lone pair of electrons. It is this which
attacks the ethanoyl chloride.
Why are acyl chlorides attacked by nucleophiles?
The carbon atom in the -COCl group has both an oxygen atom and a
chlorine atom attached to it. Both of these are very electronegative. They
both pull electrons towards themselves, leaving the carbon atom quite
positively charged. It is that carbon atom which is attacked by the lone

http://www.chemguide.co.uk/mechanisms/addelim/aminestt.html (2 of 6)30/12/2004 11:26:58

Explaining the reaction between acyl chlorides and amines - addition / elimination

pair on the nitrogen atom in an ethylamine molecule.


The ethanoyl chloride molecule is also planar (flat) around that carbon
atom, and that leaves plenty of room for a nucleophile to attack either
from above or below the plane of the molecule.

The mechanism
The reaction happens in two main stages - an addition stage, followed by
an elimination stage.
In the addition stage, an ethylamine molecule becomes attached to the
carbon in the ethanoyl chloride.
As the lone pair on the nitrogen approaches the fairly positive carbon in
the ethanoyl chloride, it moves to form a bond with it. In the process, the
two electrons in one of the carbon-oxygen bonds are repelled entirely
onto the oxygen, leaving it negatively charged.

Notice that the nitrogen atom has gained a positive charge. The
underlying reason for this is that when the lone pair forms a bond with
the carbon, electrons are moving away from the nitrogen. What matters,
though, is that you remember to show the positive charge in an exam.
Think of it like this:
http://www.chemguide.co.uk/mechanisms/addelim/aminestt.html (3 of 6)30/12/2004 11:26:58

Explaining the reaction between acyl chlorides and amines - addition / elimination

If you leave out the positive charge, two things are wrong with the
equation:

The charges don't balance. You start with two overall neutral
molecules and, if you forgot the positive charge, you would end up
with a negative ion. There has got to be a positive charge
somewhere to balance the negative one.
The nitrogen looks wrong! Nitrogen normally forms three bonds,
but here it is forming four. Nitrogen can only form four bonds if it
carries a positive charge. (A positively charged nitrogen atom has
the same electronic structure as carbon - which normally forms
four bonds.)

You can put both things right with a positive charge on the nitrogen.

The elimination stage stage of the reaction happens in two steps. In the
first, the carbon-oxygen double bond reforms. As the electron pair moves
back to form a bond with the carbon, the pair of electrons in the carbonchlorine bond are forced entirely onto the chlorine to give a chloride ion.

Note: There's nothing sophisticated going on in the switch of


the ethyl group from left to right in the diagram. It just makes
it look better in the final stages!

http://www.chemguide.co.uk/mechanisms/addelim/aminestt.html (4 of 6)30/12/2004 11:26:58

Explaining the reaction between acyl chlorides and amines - addition / elimination

Finally, one of the hydrogens attached to the nitrogen is removed as a


hydrogen ion. The electrons in the hydrogen-nitrogen bond move back
onto the nitrogen, cancelling the positive charge.
The hydrogen ion might be removed in one of two ways. The first way is
entirely consistent with what happens in the reactions between water or
ethanol and acyl chlorides. The hydrogen is removed by the chloride ion.

The hydrogen chloride produced would at once react with any excess
ethylamine present to form ethylammonium chloride.

The other (more likely) possibility is that the hydrogen ion gets removed
directly by an ethylamine molecule. This forms an ethylammonium ion.
The ethylammonium ion together with the chloride ion formed in the
previous stage makes up the ethylammonium chloride produced in the
reaction.

http://www.chemguide.co.uk/mechanisms/addelim/aminestt.html (5 of 6)30/12/2004 11:26:58

Explaining the reaction between acyl chlorides and amines - addition / elimination

Note: You probably don't really need to know both of these


routes for the removal of the hydrogen ion. Choose for
yourself. It doesn't matter very much, because both will
happen.
The route involving the formation of hydrogen chloride is
exactly in line with the water or alcohol mechanisms, but you
do need to remember to react the HCl with excess
ethylamine. You may well feel that this is less daunting than
the other route.
The removal by another ethylamine molecule is probably the
major route as long as the ethylamine is in excess - which it
almost certainly will be. The reaction is normally done in the
lab by dropping the ethanoyl chloride into concentrated
ethylamine solution.
Don't worry too much about this. Most sources opt out of the
problem altogether by quoting the final step simply as loss of
hydrogen ion: " - H+ ", without making any attempt to explain
what removes it.

Where would you like to go now?


To menu of nucleophilic addition / elimination reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/addelim/aminestt.html (6 of 6)30/12/2004 11:26:58

preparation of acyl chlorides (acid chlorides)

MAKING ACYL CHLORIDES (ACID CHLORIDES)

This page looks at ways of swapping the -OH group in the -COOH group
of a carboxylic acid for a chlorine atom to make acyl chlorides (acid
chlorides), and is a very slightly modified version of a page in the section
on reactions of carboxylic acids.
It covers the use of phosphorus(V) chloride, phosphorus(III) chloride and
sulphur dichloride oxide (thionyl chloride).

Replacing the -OH in a carboxylic acid by -Cl


We'll take the conversion of ethanoic acid to ethanoyl chloride as typical.

Note: If you haven't already done so, it would probably pay


you to have a quick look at the beginning of the introduction
to acyl chlorides before you go on.
Use the BACK button on your browser to return to this page.

Replacing the -OH group using phosphorus(V) chloride, PCl5


Phosphorus(V) chloride is a solid which reacts with carboxylic acids in
the cold to give steamy acidic fumes of hydrogen chloride. It leaves a
liquid mixture of the acyl chloride and a phosphorus compound,
phosphorus trichloride oxide (phosphorus oxychloride) - POCl3.
The acyl chloride can be separated by fractional distillation.
For example:

http://www.chemguide.co.uk/organicprops/acylchlorides/preparation.html (1 of 3)30/12/2004 11:27:01

preparation of acyl chlorides (acid chlorides)

Replacing the -OH group using phosphorus(III) chloride, PCl3


Phosphorus(III) chloride is a liquid at room temperature. Its reaction with
a carboxylic acid is less dramatic than that of phosphorus(V) chloride
because there is no hydrogen chloride produced.
You end up with a mixture of the acyl chloride and phosphoric(III) acid
(old names: phosphorous acid or orthophosphorous acid), H3PO3.
For example:

Again, the ethanoyl chloride can be separated by fractional distillation.

Replacing the -OH group using sulphur dichloride oxide (thionyl


chloride)
Sulphur dichloride oxide (thionyl chloride) is a liquid at room temperature
and has the formula SOCl2.
Traditionally, the formula is written as shown, despite the fact that the
modern name writes the chlorine before the oxygen (alphabetical order).
The sulphur dichloride oxide reacts with carboxylic acids to produce an
acyl chloride, and sulphur dioxide and hydrogen chloride gases are given
off.
For example:

The separation is simplified to an extent because the by-products are


both gases. You would obviously still have to fractionally distil the mixture
to separate the acyl chloride from any excess acid or sulphur dichloride
http://www.chemguide.co.uk/organicprops/acylchlorides/preparation.html (2 of 3)30/12/2004 11:27:01

preparation of acyl chlorides (acid chlorides)

oxide.

Where would you like to go now?


To the acyl chlorides menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/acylchlorides/preparation.html (3 of 3)30/12/2004 11:27:01

acyl chlorides and water, alcohols or phenol

THE REACTION OF ACYL CHLORIDES WITH


WATER, ALCOHOLS AND PHENOL

This page looks at the reactions of acyl chlorides (acid chlorides) with
water, alcohols and phenol. These reactions are all considered together
because their chemistry is so similar.

Similarities between the reactions


Comparing the structures of water, ethanol and phenol
Each substance contains an -OH group. In water, this is attached to a
hydrogen atom. In an alcohol, it is attached to an alkyl group - shown in
the diagrams below as "R". In phenol, it is attached to a benzene ring.
Phenol is C6H5OH.

Note: If you aren't sure about using this symbol for a


benzene ring, you could follow this link to find out all about it.
It is likely to take you some time, though, and you may have
to visit several other pages as well.
It isn't particularly important in the context of the current
page. All you need to know is that at each corner of the
hexagon there is a carbon atom, together with a hydrogen
atom apart from where the -OH group is attached.
If you choose to follow the link, use the BACK button (or the
HISTORY file or GO menu) on your browser to return to this
http://www.chemguide.co.uk/organicprops/acylchlorides/oxygen.html (1 of 8)30/12/2004 11:27:09

acyl chlorides and water, alcohols or phenol

page.

What happens when these react with an acyl chloride?


We'll take ethanoyl chloride as typical of the acyl chlorides. For UK A
level purposes, it is the one you are most likely to be asked about
anyway.
Taking a general case of a reaction between ethanoyl chloride and a
compound X-O-H (where X is hydrogen, or an alkyl group or a benzene
ring):

So . . . in each case, hydrogen chloride gas is produced - the hydrogen


coming from the -OH group, and the chlorine from the ethanoyl chloride.
Everything left over just gets joined together.

Note: If you are interested in exploring the general


mechanism for these reactions, you will find it by following
this link to another part of the site dealing with nucleophilic
addition-elimination reactions.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/acylchlorides/oxygen.html (2 of 8)30/12/2004 11:27:09

acyl chlorides and water, alcohols or phenol

The individual reactions


The reaction with water
Modifying the general equation we've just looked at, you will see that you
will get ethanoic acid produced together with hydrogen chloride gas.

This is more usually (and more easily!) written as:

Adding an acyl chloride to water produces the corresponding carboxylic


acid together with steamy acidic fumes of hydrogen chloride. The
reaction is usually extremely vigorous at room temperature.

Note: If you are want the mechanism for this reaction, you
will find it by following this link to another part of the site
dealing with nucleophilic addition-elimination reactions.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/acylchlorides/oxygen.html (3 of 8)30/12/2004 11:27:09

acyl chlorides and water, alcohols or phenol

The reaction with alcohols


We'll start by taking the general case of any alcohol reacting with
ethanoyl chloride. The equation would be:

The organic product this time is an ester. For example, with ethanol you
would get the ester ethyl ethanoate:

. . . or, more commonly:

Note: There is no "right" way of writing these equations. I am


using the colour-coded structural versions to try to show the
link between the various reactions. You can choose
whichever version you want to use.
If you are want the mechanism for this reaction, you will find it
by following this link to another part of the site dealing with
nucleophilic addition-elimination reactions.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/acylchlorides/oxygen.html (4 of 8)30/12/2004 11:27:09

acyl chlorides and water, alcohols or phenol

This is an easy way of producing an ester from an alcohol because it


happens at room temperature, and is irreversible. Making an ester from
an alcohol and a carboxylic acid (the usual alternative method) needs
heat, a catalyst and is reversible - so that it is difficult to get a 100%
conversion.

The reaction with phenols


The reaction with phenol itself
Phenols have an -OH group attached directly to a benzene ring. In the
substance normally called "phenol", there isn't anything else attached to
the ring as well. We'll look at that first.
The reaction between phenol and ethanoyl chloride isn't quite as
vigorous as that between alcohols and ethanoyl chloride. The reactivity
of the -OH group is modified by the benzene ring.
That apart, the reaction is just the same as with an alcohol.

Or, more simply:

Especially if you write the equation in this second way, it is obvious that
you have just produced another ester - in this case, called phenyl
ethanoate.
But beware! You may come across the structure drawn in a variety of
other ways which make it look much more as if it was a derivative of
phenol (which of course it is!).
For example:
http://www.chemguide.co.uk/organicprops/acylchlorides/oxygen.html (5 of 8)30/12/2004 11:27:09

acyl chlorides and water, alcohols or phenol

This would tend to happen if you were concentrating on the reactions of


the phenol rather than the acyl chloride.
Looking at it this way, notice that the hydrogen of the phenol -OH group
has been replaced by an acyl group - an alkyl group attached to a carbonoxygen double bond.
You can say that the phenol has been acylated or has undergone
acylation.
Because of the nature of this particular acyl group, it is also described as
ethanoylation. The hydrogen is being replaced by an ethanoyl group,
CH3CO-.

Using a similar reaction to make aspirin


The reaction with phenol itself isn't very important, but you can make
aspirin by a very similar reaction.
Here is 2-hydroxybenzoic acid (also known as 2hydroxybenzenecarboxylic acid). The old name for this is salicylic acid.
You might find it written in either of these two ways. They are the same
structure with the molecule just flipped over in space.

http://www.chemguide.co.uk/organicprops/acylchlorides/oxygen.html (6 of 8)30/12/2004 11:27:09

acyl chlorides and water, alcohols or phenol

You might also find it with the -OH group at the top and the -COOH
group next door and either to the left or right of it. Life can get very
confusing!
When this reacts with ethanoyl chloride, it is ethanoylated (or acylated, if
you want to use the more general term) to give:

You might find all sorts of other variants on drawing this as well.
This molecule is aspirin.

Note: Although you can make aspirin this way, the 2hydroxybenzoic acid is normally ethanoylated commercially
using ethanoic anhydride rather than ethanoyl chloride. There
are various reasons for this, which are explored at the bottom
of the corresponding page about the reactions of acid
anhydrides with oxygen compounds.
Use the BACK button on your browser if you want to return to
this page later.

http://www.chemguide.co.uk/organicprops/acylchlorides/oxygen.html (7 of 8)30/12/2004 11:27:09

acyl chlorides and water, alcohols or phenol

Where would you like to go now?


To the acyl chlorides menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/acylchlorides/oxygen.html (8 of 8)30/12/2004 11:27:09

acid anhydrides and water, alcohols or phenol

THE REACTION OF ACID ANHYDRIDES WITH


WATER, ALCOHOLS AND PHENOL

This page looks at the reactions of acid anhydrides with water, alcohols
and phenols (including the manufacture of aspirin). These reactions are all
considered together because their chemistry is so similar.
There is also a great similarity between acid anhydrides and acyl
chlorides (acid chlorides) as far as these reactions are concerned.
Concentrate on these similarities as you go through this page because it
should help you to remember.

Similarities between the reactions


Comparing the structures of water, ethanol and phenol
Each substance contains an -OH group. In water, this is attached to a
hydrogen atom. In an alcohol, it is attached to an alkyl group - shown in
the diagrams below as "R". In phenols, it is attached to a benzene ring.
Phenol (the simplest member of the family of phenols) is C6H5OH.

http://www.chemguide.co.uk/organicprops/anhydrides/oxygen.html (1 of 8)30/12/2004 11:27:20

acid anhydrides and water, alcohols or phenol

Note: If you aren't sure about using this symbol for a benzene
ring, you could follow this link to find out all about it. It is likely
to take you some time, though, and you may have to visit
several other pages as well.
It isn't particularly important in the context of the current page.
All you need to know is that at each corner of the hexagon
there is a carbon atom, together with a hydrogen atom apart
from where the -OH group is attached.
If you choose to follow the link, use the BACK button (or the
HISTORY file or GO menu) on your browser to return to this
page.

Comparing the reactions of acyl chlorides and acid anhydrides with


these compounds
Because the formula is much easier, it helps to start with the acyl
chlorides.
The reactions with acyl chlorides
We'll take ethanoyl chloride as typical of the acyl chlorides.
Taking a general case of a reaction between ethanoyl chloride and a
compound X-O-H (where X is hydrogen, or an alkyl group or a benzene
ring):

So . . . in each case, hydrogen chloride gas is produced - the hydrogen


coming from the -OH group, and the chlorine from the ethanoyl chloride.
Everything left over just gets joined together.
The same reactions with acid anhydrides
http://www.chemguide.co.uk/organicprops/anhydrides/oxygen.html (2 of 8)30/12/2004 11:27:20

acid anhydrides and water, alcohols or phenol

Ethanoic anhydride is the only one you are likely to come across for UK A
level purposes.

If you compare this with the acyl chloride equation, you can see that the
only difference is that ethanoic acid is produced as the second product of
the reaction rather than hydrogen chloride.

Note: The colour coding in these equations is to try to help


you to see where everything ends up and where it came from,
and to enable you to compare the two reactions more easily.
You can think of the entire bit of the ethanoic anhydride shown
in red as being exactly the equivalent of the chlorine atom in
the acyl chloride as far as these reactions are concerned.

These reactions are just the same as the corresponding acyl chloride
reactions except:

Ethanoic acid is formed as the second product rather than


hydrogen chloride gas.
The reactions are slower. Acid anhydrides aren't so violently
reactive as acyl chlorides.

The individual reactions


The reaction with water
Modifying the general equation we've just looked at, you will see that you
just get two molecules of ethanoic acid produced.
http://www.chemguide.co.uk/organicprops/anhydrides/oxygen.html (3 of 8)30/12/2004 11:27:20

acid anhydrides and water, alcohols or phenol

This is more usually (and more easily!) written as:

The reaction happens slowly at room temperature (faster on gentle


warming) without a great deal exciting to observe - unlike in the acyl
chloride case where hydrogen chloride fumes are produced. You mix two
colourless liquids and get another colourless liquid!
The equivalent acyl chloride reaction is:

The reaction with alcohols


We'll start by taking the general case of any alcohol reacting with ethanoic
anhydride. The equation would be:

or, more simply:

The product this time (apart from the ethanoic acid always produced) is an

http://www.chemguide.co.uk/organicprops/anhydrides/oxygen.html (4 of 8)30/12/2004 11:27:20

acid anhydrides and water, alcohols or phenol

ester. For example, with ethanol you would get the ester ethyl ethanoate:

This reaction also needs gentle heating for it to happen at a reasonable


rate, and again there isn't anything visually dramatic.
The equivalent acyl chloride reaction is:

Note: If you are exceptionally wide awake, you might wonder


why the reaction between ethanoic anhydride and an alcohol
doesn't result in two molecules of ester. Carboxylic acids like
ethanoic acid react with alcohols to give esters - so why
doesn't this happen with the acid molecule that is formed
reacting with alcohol in the mixture?
The reason is simply one of reaction conditions. This reaction
is happening in the absence of any catalysts. To get
carboxylic acids and alcohols to react at any sort of
reasonable rate, you need heat and a catalyst such as
concentrated sulphuric acid.

The reaction with phenols


The reaction with phenol itself
Phenols have an -OH group attached directly to a benzene ring. In the
substance normally called "phenol", there isn't anything else attached to
the ring as well. We'll look at that first.
The reaction between phenol and ethanoic anhydride isn't particularly
important, but you would get an ester just as you do with an alcohol.

http://www.chemguide.co.uk/organicprops/anhydrides/oxygen.html (5 of 8)30/12/2004 11:27:20

acid anhydrides and water, alcohols or phenol

Or, more simply:

Especially if you write the equation in this second way, it is obvious that
you have just produced another ester - in this case, called phenyl
ethanoate.
The equivalent acyl chloride reaction is:

But beware! You may come across the structure of the ester drawn in a
variety of other ways which make it look much more as if it was a
derivative of phenol (which of course it is!).
For example:

Looking at it this way, notice that the hydrogen of the phenol -OH group
has been replaced by an acyl group - an alkyl group attached to a carbonoxygen double bond.
You can say that the phenol has been acylated or has undergone
acylation.
Because of the nature of this particular acyl group, it is also described as
ethanoylation. The hydrogen is being replaced by an ethanoyl group,

http://www.chemguide.co.uk/organicprops/anhydrides/oxygen.html (6 of 8)30/12/2004 11:27:20

acid anhydrides and water, alcohols or phenol

CH3CO-.

Using a similar reaction to make aspirin


The reaction with phenol itself isn't very important, but you can make
aspirin by a very similar reaction.
The molecule below is 2-hydroxybenzoic acid (also known as 2hydroxybenzenecarboxylic acid). The old name for this is salicylic acid.
You might find it written in either of these two ways. They are the same
structure with the molecule just flipped over in space.

You might also find it with the -OH group at the top and the -COOH group
next door and either to the left or right of it. Life can get very confusing!
When this reacts with ethanoic anhydride, it is ethanoylated (or acylated, if
you want to use the more general term) to give:

You might find all sorts of other variants on drawing this as well.
This molecule is aspirin.
Although this reaction can also be done with ethanoyl chloride, aspirin is
manufactured by reacting 2-hydroxybenzoic acid with ethanoic anhydride
at 90C.
http://www.chemguide.co.uk/organicprops/anhydrides/oxygen.html (7 of 8)30/12/2004 11:27:20

acid anhydrides and water, alcohols or phenol

The reasons for using ethanoic anhydride rather than ethanoyl chloride
include:

Ethanoic anhydride is cheaper than ethanoyl chloride.


Ethanoic anhydride is safer to use than ethanoyl chloride. It is less
corrosive and not so readily hydrolysed (its reaction with water is
slower).
Ethanoic anhydride doesn't produce dangerous (corrosive and
poisonous) fumes of hydrogen chloride.

Where would you like to go now?


To the acid anhydrides menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/anhydrides/oxygen.html (8 of 8)30/12/2004 11:27:20

Acid Anhydrides Menu

Understanding Chemistry

ACID ANHYDRIDES MENU

Background . . .
An introduction to acid anhydrides including a general look at their
reactivity.
Reactions of acid anhydrides with oxygen compounds . . .
The reactions of acid anhydrides with water, alcohols and phenols
(including the manufacture of aspirin).
Reactions of acid anhydrides with nitrogen compounds . . .
The reactions of acid anhydrides with ammonia and primary
amines.

Go to menu of other organic compounds . . .


Go to Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/anhydridemenu.html30/12/2004 11:27:23

an introduction to acid anhydrides

INTRODUCING ACID ANHYDRIDES

This page explains what acid anhydrides are and looks at their simple
physical properties such as boiling points. It introduces their chemical
reactivity in a general way, but details of specific reactions are given on
separate pages - see the acid anhydrides menu (link at the bottom of the
page).

What are acid anhydrides?


The structure of acid anhydrides
A carboxylic acid such as ethanoic acid has the structure:

If you took two ethanoic acid molecules and removed a molecule of


water between them you would get the acid anhydride, ethanoic
anhydride (old name: acetic anhydride).

You can actually make ethanoic anhydride by dehydrating ethanoic acid,


but it is normally made in a more efficient, round-about way. (Making
ethanoic anhydride is beyond the scope of UK A level.)
http://www.chemguide.co.uk/organicprops/anhydrides/background.html (1 of 5)30/12/2004 11:27:27

an introduction to acid anhydrides

Naming acid anhydrides


This is really easy. You just take the name of the parent acid, and
replace the word "acid" by "anhydride". "Anhydride" simply means
"without water".
So . . . ethanoic acid forms ethanoic anhydride; propanoic acid forms
propanoic anhydride, and so on.
For UK A level purposes, the only one you are at all likely to come across
is ethanoic anhydride.

Physical properties of acid anhydrides


We will take ethanoic anhydride as typical.
Appearance
Ethanoic anhydride is a colourless liquid, smelling strongly of vinegar
(ethanoic acid).
The smell is because ethanoic anhydride reacts with water vapour in the
air (and moisture in your nose) to produce ethanoic acid again. This
reaction with water is given in detail on another page. (Find it from the
acid anhydrides menu - link at the bottom of this page.)

Solubility in water
Ethanoic anhydride can't be said to dissolve in water because it reacts
with it to give ethanoic acid. There is no such thing as an aqueous
solution of ethanoic anhydride.

http://www.chemguide.co.uk/organicprops/anhydrides/background.html (2 of 5)30/12/2004 11:27:27

an introduction to acid anhydrides

Boiling point
Ethanoic anhydride boils at 140C. This is because it is a fairly big polar
molecule and so has both van der Waals dispersion forces and dipoledipole attractions.
It doesn't, however, form hydrogen bonds. That means that its boiling
point isn't as high as a carboxylic acid of similar size. For example,
pentanoic acid (the most similarly sized acid) boils at 186C.

Note: If you aren't happy about intermolecular forces


(including van der Waals dispersion forces and hydrogen
bonds) then you really ought to follow this link before you go
on.
Use the BACK button on your browser to return to this page.

Reactivity of acid anhydrides


Comparing acid anhydrides with acyl chlorides (acid chlorides)
You have almost certainly come across acid anhydrides for the first time
just after looking at acyl chlorides, or you may be studying them at the
same time as acyl chlorides.
It is much, much easier to think of acid anhydrides as if they were a sort
of modified acyl chloride than to try to learn about them from scratch.
That is the line I intend to take throughout all this section.
Compare the structure of an acid anhydride with that of an acyl chloride looking carefully at the way it is colour-coded in the diagram.

http://www.chemguide.co.uk/organicprops/anhydrides/background.html (3 of 5)30/12/2004 11:27:27

an introduction to acid anhydrides

In the reactions of ethanoic anhydride, the red group at the bottom


always stays intact. It is behaving in many ways as if it was a single atom
- just like the chlorine atom in the acyl chloride.
The usual reaction of an acyl chloride is replacement of the chlorine by
something else.
Taking ethanoyl chloride as typical, the initial reaction is of this kind:

Hydrogen chloride gas is given off, although that might go on to react


with other components of the mixture.
With an acid anhydride, the reaction is slower, but the only essential
difference is that instead of hydrogen chloride being produced as the
other product, you get ethanoic acid instead.

Just like the hydrogen chloride, this might afterwards go on to react with
other things present.

http://www.chemguide.co.uk/organicprops/anhydrides/background.html (4 of 5)30/12/2004 11:27:27

an introduction to acid anhydrides

The reactions (of both acyl chlorides and acid anhydrides) involve things
like water, alcohols and phenols, or ammonia and amines. All of these
particular cases contain a very electronegative element with an active
lone pair of electrons - either oxygen or nitrogen.

Note: You can find details of all these reactions from the
acid anhydrides menu (link below).

Where would you like to go now?


To the acid anhydrides menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/anhydrides/background.html (5 of 5)30/12/2004 11:27:27

acid anhydrides with ammonia or primary amines

THE REACTIONS OF ACID ANHYDRIDES WITH


AMMONIA AND PRIMARY AMINES

This page looks at the reactions of acid anhydrides with ammonia and with
primary amines. These reactions are considered together because their
chemistry is so similar.
There is also a great similarity between acid anhydrides and acyl chlorides
(acid chlorides) as far as these reactions are concerned. Concentrate on
these similarities as you go through this page because it should help you
to remember.

Similarities between the reactions


Comparing the structures of ammonia and primary amines
Each substance contains an -NH2 group. In ammonia, this is attached to a
hydrogen atom. In a primary amine, it is attached to an alkyl group (shown
by "R" in the diagram below) or a benzene ring.

http://www.chemguide.co.uk/organicprops/anhydrides/nitrogen.html (1 of 9)30/12/2004 11:27:39

acid anhydrides with ammonia or primary amines

Note: If you aren't sure about using this symbol for a benzene
ring, you could follow this link to find out all about it. It is likely
to take you some time, though, and you may have to visit
several other pages as well.
It isn't particularly important in the context of the current page.
All you need to know is that at each corner of the hexagon
there is a carbon atom, together with a hydrogen atom apart
from where the -NH2 group is attached.
If you choose to follow the link, use the BACK button (or the
HISTORY file or GO menu) on your browser to return to this
page.

Comparing the reactions of acyl chlorides and acid anhydrides with


these compounds
Because the formula is much easier, it helps to start with the acyl chlorides.
The reactions with acyl chlorides
We'll take ethanoyl chloride as typical of the acyl chlorides.
Taking a general case of a reaction between ethanoyl chloride and a
compound XNH2 (where X is hydrogen, or an alkyl group, or a benzene
ring). The reaction happens in two stages:
First:

So . . . in each case, you initially get hydrogen chloride gas - the hydrogen
coming from the -NH2 group, and the chlorine from the ethanoyl chloride.
Everything left over just gets joined together.
http://www.chemguide.co.uk/organicprops/anhydrides/nitrogen.html (2 of 9)30/12/2004 11:27:39

acid anhydrides with ammonia or primary amines

But ammonia and amines are basic, and react with the hydrogen chloride
to produce a salt. So the second stage of the reaction is:

The same reactions with acid anhydrides


Ethanoic anhydride is the only one you are likely to come across for UK A
level purposes.
Again, the reaction happens in two stages. In the first:

If you compare this with the acyl chloride equation, you can see that the
only difference is that ethanoic acid is produced as the second product of
the reaction rather than hydrogen chloride.

Note: The colour coding in these equations is to try to help


you to see where everything ends up and where it came from,
and to enable you to compare the two reactions more easily.
You can think of the entire bit of the ethanoic anhydride shown
in red as being exactly the equivalent of the chlorine atom in
the acyl chloride as far as these reactions are concerned.

http://www.chemguide.co.uk/organicprops/anhydrides/nitrogen.html (3 of 9)30/12/2004 11:27:39

acid anhydrides with ammonia or primary amines

Then the ethanoic acid reacts with excess ammonia or amine to give a salt
- this time an ethanoate.

This looks more difficult than the acyl chloride case because of the way the
salt is written. You get an ethanoate ion and a positive ion with this
structure:

This is easier to understand with real compounds - as you will see below.
In summary:
These reactions are just the same as the corresponding acyl chloride
reactions except:

Initially, ethanoic acid is formed as the second product rather than


hydrogen chloride gas.
The second stage of the reaction involves the formation of an
ethanoate rather than a chloride.
The reactions are slower. Acid anhydrides aren't so violently
reactive as acyl chlorides, and the reactions normally need heating.

The individual reactions


The reaction with ammonia
In this case, the "X" in the equations above is a hydrogen atom. So in the
first instance you get ethanoic acid and an organic compound called an
amide.

http://www.chemguide.co.uk/organicprops/anhydrides/nitrogen.html (4 of 9)30/12/2004 11:27:39

acid anhydrides with ammonia or primary amines

Amides contain the group -CONH2. In the reaction between ethanoic


anhydride and ammonia, the amide formed is called ethanamide.

This is more usually (and more easily!) written as:

The ethanoic acid produced reacts with excess ammonia to give


ammonium ethanoate.

. . . and you can combine all this together to give one overall equation:

You need to follow this through really carefully, because the two products
of the reaction overall can look confusingly similar.
The corresponding reaction with an acyl chloride is:

http://www.chemguide.co.uk/organicprops/anhydrides/nitrogen.html (5 of 9)30/12/2004 11:27:39

acid anhydrides with ammonia or primary amines

Note: There isn't anything difficult involved in the fact that I


have moved the position of the positive charge in the
ammonium ion in this equation compared with the previous
one. In each case I have placed it as close as possible to the
negative charge in the ethanoate ion or chloride ion.
In the ethanoate case, that makes it easier to draw the
structure of the similar positive ion in the amine equations
below.

The reaction with primary amines


The reaction with methylamine
We'll take methylamine as typical of primary amines where the -NH2 is
attached to an alkyl group.
The initial equation would be:

The first product this time is called an N-substituted amide.


If you compare the structure with the amide produced in the reaction with
ammonia, the only difference is that one of the hydrogens on the nitrogen
has been substituted for a methyl group.
This particular compound is N-methylethanamide. The "N" simply shows
that the substitution is on the nitrogen atom, and not elsewhere in the
molecule.

http://www.chemguide.co.uk/organicprops/anhydrides/nitrogen.html (6 of 9)30/12/2004 11:27:39

acid anhydrides with ammonia or primary amines

The equation would normally be written:

You can think of primary amines as just being modified ammonia. If


ammonia is basic and forms a salt with the ethanoic acid, excess
methylamine will do exactly the same thing.

The salt is called methylammonium ethanoate. It is just like ammonium


ethanoate, except that one of the hydrogens has been replaced by a
methyl group.
You would usually combine these equations into one overall equation for
the reaction:

The corresponding reaction with an acyl chloride is:

The reaction with phenylamine (aniline)


Phenylamine is the simplest primary amine where the -NH2 group is
attached directly to a benzene ring. Its old name is aniline.
In phenylamine, there isn't anything else attached to the ring as well. You
can write the formula of phenylamine as C6H5NH2.
There is no essential difference between this reaction and the reaction with
methylamine, but I just want to look at the structure of the N-substituted
amide formed.
The overall equation for the reaction is:

http://www.chemguide.co.uk/organicprops/anhydrides/nitrogen.html (7 of 9)30/12/2004 11:27:39

acid anhydrides with ammonia or primary amines

The products are N-phenylethanamide and phenylammonium ethanoate.

This reaction can sometimes look (even more!) confusing if the


phenylamine is drawn showing the benzene ring, and especially if the
reaction is looked at from the point of view of the phenylamine.
For example, the product molecule might be drawn looking like this:

If you stop and think about it, this is obviously the same molecule as in the
equation above, but it stresses the phenylamine part of it much more.
Looking at it this way, notice that one of the hydrogens of the -NH2 group
has been replaced by an acyl group - an alkyl group attached to a carbonoxygen double bond.
You can say that the phenylamine has been acylated or has undergone
acylation.
Because of the nature of this particular acyl group, it is also described as
ethanoylation. The hydrogen is being replaced by an ethanoyl group,
CH3CO-.

http://www.chemguide.co.uk/organicprops/anhydrides/nitrogen.html (8 of 9)30/12/2004 11:27:39

acid anhydrides with ammonia or primary amines

Note: For UK A level purposes, this is probably the most likely


context that you will meet this particular reaction in. You would
need to know that phenylamine reacted with ethanoic
anhydride to produce an N-substituted amide, and that this was
an example of acylation (or ethanoylation). You would probably
be expected to draw the structure, but would be pretty unlucky
if you had to write equations as above.
I have taught a lot of very clever people in my time, but there
aren't many where I would have confidence that they could
produce those equations under exam conditions!

Where would you like to go now?


To the acid anhydrides menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/anhydrides/nitrogen.html (9 of 9)30/12/2004 11:27:39

acyl chlorides with ammonia or primary amines

THE REACTION OF ACYL CHLORIDES WITH


AMMONIA AND PRIMARY AMINES

This page looks at the reactions of acyl chlorides (acid chlorides) with
ammonia and with primary amines. These reactions are considered
together because their chemistry is so similar.

Similarities between the reactions


Comparing the structures of ammonia and primary amines
Each substance contains an -NH2 group. In ammonia, this is attached to a
hydrogen atom. In a primary amine, it is attached to an alkyl group (shown
by "R" in the diagram below) or a benzene ring.

Note: If you aren't sure about using this symbol for a benzene
ring, you could follow this link to find out all about it. It is likely
to take you some time, though, and you may have to visit
several other pages as well.
It isn't particularly important in the context of the current page.
All you need to know is that at each corner of the hexagon
there is a carbon atom, together with a hydrogen atom apart
from where the -NH2 group is attached.
If you want to know more about what primary amines are, you
http://www.chemguide.co.uk/organicprops/acylchlorides/nitrogen.html (1 of 7)30/12/2004 11:27:47

acyl chlorides with ammonia or primary amines

could follow this link to a page introducing amines, and just


read the beginning of it.
If you choose to follow either of these links, use the BACK
button (or the HISTORY file or GO menu) on your browser to
return to this page.

What happens when these react with an acyl chloride?


We'll take ethanoyl chloride as typical of the acyl chlorides. For UK A level
purposes, it is the one you are most likely to be asked about anyway.
Taking a general case of a reaction between ethanoyl chloride and a
compound XNH2 (where X is hydrogen, or an alkyl group, or a benzene
ring). The reaction happens in two stages:
First:

So . . . in each case, you initially get hydrogen chloride gas - the hydrogen
coming from the -NH2 group, and the chlorine from the ethanoyl chloride.
Everything left over just gets joined together.
But ammonia and amines are basic, and react with the hydrogen chloride
to produce a salt. So the second stage of the reaction is:

This is easier to understand with real compounds - as you will see below.

http://www.chemguide.co.uk/organicprops/acylchlorides/nitrogen.html (2 of 7)30/12/2004 11:27:47

acyl chlorides with ammonia or primary amines

Note: If you are interested in exploring the general mechanism


for these reactions, you will find it by following this link to
another part of the site dealing with nucleophilic additionelimination reactions.
Use the BACK button on your browser to return to this page.

The individual reactions


The reaction with ammonia
In this case, the "X" in the equations above is a hydrogen atom. So in the
first instance you get hydrogen chloride gas and an organic compound
called an amide.
Amides contain the group -CONH2. In the reaction between ethanoyl
chloride and ammonia, the amide formed is called ethanamide.

This is more usually (and more easily!) written as:

The hydrogen chloride produced reacts with excess ammonia to give


ammonium chloride.

. . . and you can combine all this together to give one overall equation:

http://www.chemguide.co.uk/organicprops/acylchlorides/nitrogen.html (3 of 7)30/12/2004 11:27:47

acyl chlorides with ammonia or primary amines

You normally add the ethanoyl chloride to a concentrated solution of


ammonia in water. There is a very violent reaction producing lots of white
smoke - a mixture of solid ammonium chloride and ethanamide. Some of
the mixture remains dissolved in water as a colourless solution.

Note: If you are want the mechanism for this reaction, you will
find it by following this link to another part of the site dealing
with nucleophilic addition-elimination reactions.
Use the BACK button on your browser to return to this page.

The reaction with primary amines


The reaction with methylamine
We'll take methylamine as typical of primary amines where the -NH2 is
attached to an alkyl group.
The initial equation would be:

The organic product this time is called an N-substituted amide.


If you compare the structure with the amide produced in the reaction with
ammonia, the only difference is that one of the hydrogens on the nitrogen
has been substituted for a methyl group.
http://www.chemguide.co.uk/organicprops/acylchlorides/nitrogen.html (4 of 7)30/12/2004 11:27:47

acyl chlorides with ammonia or primary amines

This particular compound is N-methylethanamide. The "N" simply shows


that the substitution is on the nitrogen atom, and not elsewhere in the
molecule.
The equation would normally be written:

You can think of primary amines as just being modified ammonia. If


ammonia is basic and forms a salt with the hydrogen chloride, excess
methylamine will do exactly the same thing.

The salt is called methylammonium chloride. It is just like ammonium


chloride, except that one of the hydrogens has been replaced by a methyl
group.
You would usually combine these equations into one overall equation for
the reaction:

Note: If you are want the mechanism for this reaction


(although using ethylamine rather than methylamine as the
example), you will find it by following this link to another part of
the site dealing with nucleophilic addition-elimination reactions.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/acylchlorides/nitrogen.html (5 of 7)30/12/2004 11:27:47

acyl chlorides with ammonia or primary amines

The reaction looks exactly the same as the one with ammonia. The
methylamine is again used as a concentrated solution in water. There is a
violent reaction producing a white solid mixture of N-methylethanamide
and methylammonium chloride.

The reaction with phenylamine (aniline)


Phenylamine is the simplest primary amine where the -NH2 group is
attached directly to a benzene ring. Its old name is aniline.
In phenylamine, there isn't anything else attached to the ring as well. You
can write the formula of phenylamine as C6H5NH2.
There is no essential difference between this reaction and the reaction with
methylamine, except that the phenylamine is a brownish liquid, and the
solid products tend to be stained brownish.
The overall equation for the reaction is:

The products are N-phenylethanamide and phenylammonium chloride.

This reaction can sometimes look confusing if the phenylamine is drawn


showing the benzene ring, and especially if the reaction is looked at from
the point of view of the phenylamine.
For example, the product molecule might be drawn looking like this:

http://www.chemguide.co.uk/organicprops/acylchlorides/nitrogen.html (6 of 7)30/12/2004 11:27:47

acyl chlorides with ammonia or primary amines

If you stop and think about it, this is obviously the same molecule as in the
equation above, but it stresses the phenylamine part of it much more.
Looking at it this way, notice that one of the hydrogens of the -NH2 group
has been replaced by an acyl group - an alkyl group attached to a carbonoxygen double bond.
You can say that the phenylamine has been acylated or has undergone
acylation.
Because of the nature of this particular acyl group, it is also described as
ethanoylation. The hydrogen is being replaced by an ethanoyl group,
CH3CO-.

Where would you like to go now?


To the acyl chlorides menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/acylchlorides/nitrogen.html (7 of 7)30/12/2004 11:27:47

friedel-crafts acylation of benzene

THE REACTION OF ACYL CHLORIDES WITH


BENZENE

This page looks at the reaction of acyl chlorides (acid chlorides) with
benzene in the presence of an aluminium chloride catalyst. This is known
as a Friedel-Crafts acylation.

Friedel-Crafts acylation of benzene


Background and equations
Acylation is the term given to substituting an acyl group such as
CH3CO- into another molecule. An acyl group is a hydrocarbon group
attached to a carbon-oxygen double bond.
For UK A level purposes, the most commonly used example of an acyl
group is the ethanoyl group, CH3CO-, and so that's the one we will stick
with throughout.
So, if you react benzene with ethanoyl chloride in the presence of an
aluminium chloride catalyst, the equation for the reaction is:

Or, simplifying it without drawing the benzene ring:

In the simplified formula for the product, the phenyl group is usually

http://www.chemguide.co.uk/organicprops/acylchlorides/fc.html (1 of 4)30/12/2004 11:27:52

friedel-crafts acylation of benzene

written on the left-hand side and the alkyl group to the right of the carbonoxygen double bond - but I doubt if it really matters!
The aluminium chloride isn't written into these equations because it is
acting as a catalyst. If you wanted to include it, you could write AlCl3 over
the top of the arrow.
The product is called phenylethanone (old name, acetophenone).

Note: If you are want the mechanism for this reaction, you
will find it by following this link to another part of the site
dealing with electrophilic substitution reactions.
Use the BACK button on your browser to return to this page.

Doing the reaction


Ethanoyl chloride is added carefully to a mixture of benzene and solid
aluminium chloride in the cold. Hydrogen chloride gas is given off.
When all the ethanoyl chloride has been added, the mixture is heated
under reflux at a temperature of 60C for about 30 minutes to complete
the reaction.

Heating under reflux: Heating the flask with a condenser


placed vertically in it to prevent loss of volatile liquids.
Separating the product from the reaction mixture is fairly
long-winded and beyond the scope of this site.
You may find slight variations on the temperature and time
for this reaction.

http://www.chemguide.co.uk/organicprops/acylchlorides/fc.html (2 of 4)30/12/2004 11:27:52

friedel-crafts acylation of benzene

Why the reaction is important


Friedel-Crafts acylation is a very effective way of attaching a
hydrocarbon-based group to a benzene ring. Although the product is a
ketone (a compound containing a carbon-oxygen double bond with a
hydrocarbon group either side), it is easily converted into other things.
For example:
The carbon-oxygen double bond can be reduced to give a secondary
alcohol, which in turn can undergo a whole lot of other reactions.

Note: This reduction is explored on a page about reduction


of aldehydes and ketones. That page doesn't deal specifically
with this particular ketone, but the same principles apply.
Use the BACK button on your browser to return to this page.

Phenylethanone can also be reduced to produce ethylbenzene.


This is known as the Clemmensen reduction and involves heating the
ketone with amalgamated zinc (a mixture of zinc and mercury) and
concentrated hydrochloric acid for a long time.

This indirect route is the best way of getting an alkyl group attached to a
benzene ring.
It is possible to attach an alkyl group directly to the ring, but it is

http://www.chemguide.co.uk/organicprops/acylchlorides/fc.html (3 of 4)30/12/2004 11:27:52

friedel-crafts acylation of benzene

impossible to stop at substituting just one. An alkyl group attached to the


ring makes the ring more reactive than the original benzene. That means
that something like ethylbenzene reacts faster than benzene itself.
The result is that you get several ethyl groups substituted around the ring
rather than just one.
Attaching an acyl group to the ring makes the ring so unreactive that it
won't substitute a second one.

Note: You can find more about from Friedel-Crafts reactions


on a page in the arenes section, although quite of lot of it
repeats what is above.
Use the BACK button on your browser if you want to return to
this page.

Where would you like to go now?


To the acyl chlorides menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/acylchlorides/fc.html (4 of 4)30/12/2004 11:27:52

decarboxylation of carboxylic acids and their salts

THE DECARBOXYLATION OF CARBOXYLIC


ACIDS AND THEIR SALTS

This page looks at the formation of hydrocarbons by the decarboxylation


of the salts of carboxylic acids (and of certain acids themselves) by
heating them with soda lime. It does NOT cover the decarboxylation of
some acids by simply heating them.

Decarboxylation using soda lime


What does "decarboxylation" mean?
A carboxylic acid has the formula RCOOH where R can be hydrogen or a
hydrocarbon group such as an alkyl group. The hydrocarbon group could
equally well be based on a benzene ring.
The sodium salt of a carboxylic acid will have the formula RCOONa.
In decarboxylation, the -COOH or -COONa group is removed and
replaced with a hydrogen atom.

What is soda lime?


Soda lime is manufactured by adding sodium hydroxide solution to solid
calcium oxide (quicklime). It is essentially a mixture of sodium hydroxide,
calcium oxide and calcium hydroxide. It comes as white granules.
http://www.chemguide.co.uk/organicprops/acids/decarbox.html (1 of 3)30/12/2004 11:27:59

decarboxylation of carboxylic acids and their salts

In equations, it is almost always written as if it were simply sodium


hydroxide.
It is an easier material to handle than solid sodium hydroxide. Solid
sodium hydroxide absorbs water from the atmosphere and you tend to
end up with puddles of extremely concentrated (and corrosive) sodium
hydroxide solution if you leave it exposed to the air.
Soda lime has much less tendency to absorb water.

The reaction
The solid sodium salt of a carboxylic acid is mixed with solid soda lime,
and the mixture is heated.
For example, if you heat sodium ethanoate with soda lime, you get
methane gas formed:

This reaction can be done with certain carboxylic acids themselves, but
the acid would have to be solid. For example, benzene can be made by
heating soda lime with solid benzoic acid (benzenecarboxylic acid),
C6H5COOH.

You can think of this as first a reaction between the acid and the soda
lime to make sodium benzoate, and then a decarboxylation as in the first
example.

Where would you like to go now?

http://www.chemguide.co.uk/organicprops/acids/decarbox.html (2 of 3)30/12/2004 11:27:59

decarboxylation of carboxylic acids and their salts

To the carboxylic acids menu . . .


To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/acids/decarbox.html (3 of 3)30/12/2004 11:27:59

Esters Menu

Understanding Chemistry

ESTERS MENU

Background . . .
An introduction to esters, including large esters like animal or
vegetable fats and oils.
Preparation of esters . . .
Their preparation from alcohols (and phenols) reacting with
carboxylic acids, acyl chlorides and acid anhydrides as appropriate.
The hydrolysis of esters . . .
The hydrolysis of esters using water, acids or alkalis (including
soap making).
Polyesters . . .
A summary of the chemistry of simple polyesters like Terylene,
including their formation, structure, hydrolysis and uses.

Go to menu of other organic compounds . . .


Go to Main Menu . . .

http://www.chemguide.co.uk/organicprops/estermenu.html (1 of 2)30/12/2004 11:28:00

Esters Menu

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/estermenu.html (2 of 2)30/12/2004 11:28:00

an introduction to esters

INTRODUCING ESTERS

This page explains what esters are and looks at their simple physical
properties such as solubility and boiling points. It includes an introduction
to more complicated naturally-occurring esters like animal and vegetable
fats and oils.

What are esters?


Esters are derived from carboxylic acids. A carboxylic acid contains the COOH group, and in an ester the hydrogen in this group is replaced by a
hydrocarbon group of some kind. This could be an alkyl group like methyl
or ethyl, or one containing a benzene ring like phenyl.

A common ester - ethyl ethanoate


The most commonly discussed ester is ethyl ethanoate. In this case, the
hydrogen in the -COOH group has been replaced by an ethyl group. The
formula for ethyl ethanoate is:

Notice that the ester is named the opposite way around from the way the
formula is written. The "ethanoate" bit comes from ethanoic acid. The
"ethyl" bit comes from the ethyl group on the end.

http://www.chemguide.co.uk/organicprops/esters/background.html (1 of 12)30/12/2004 11:28:10

an introduction to esters

Note: In my experience, students starting organic chemistry


get more confused about writing names and formulae for
esters than for almost anything else - particularly when it
comes to less frequently met esters like the ones coming up
next. Take time and care to make sure you understand!

A few more esters


In each case, be sure that you can see how the names and formulae
relate to each other.

Notice that the acid is named by counting up the total number of carbon
atoms in the chain - including the one in the -COOH group. So, for
example, CH3CH2COOH is propanoic acid, and CH3CH2COO is the
propanoate group.

http://www.chemguide.co.uk/organicprops/esters/background.html (2 of 12)30/12/2004 11:28:10

an introduction to esters

Note: You can find more about naming acids and esters by
following this link to a different part of this site.
Use the BACK button on your browser to return to this page.

Fats and oils


Differences between fats and oils
Animal and vegetable fats and oils are just big complicated esters. The
difference between a fat (like butter) and an oil (like sunflower oil) is
simply in the melting points of the mixture of esters they contain.
If the melting points are below room temperature, it will be a liquid - an
oil. If the melting points are above room temperature, it will be a solid - a
fat.
The causes of the differences in melting points will be discussed further
down the page under physical properties.

A simple introduction to their structures


Fats and oils as big esters
Esters can be made from carboxylic acids and alcohols. This is
discussed in detail on another page, but in general terms, the two
combine together losing a molecule of water in the process.
We'll start with a very, very simple ester like ethyl ethanoate - not
something complicated like a fat or oil!
The diagram shows the relationship between the ethanoic acid, the
ethanol and the ester.
http://www.chemguide.co.uk/organicprops/esters/background.html (3 of 12)30/12/2004 11:28:10

an introduction to esters

This isn't intended to be a full equation. Water, of course, is also


produced.

Note: The colour coding refers to the name of the ester and
not strictly to the structure. When the ester is made, the water
that is lost comes from the whole -OH group of the acid and
just a single hydrogen from the alcohol. That means that as
far as the structure is concerned the oxygen attached to the
ethyl group actually ought to be coloured red. Don't worry
about this at this level.

Now lets make the alcohol a bit more complicated by having more than
one -OH group. The diagram below shows the structure of propane-1,2,3triol (old name: glycerol).

Just as with the ethanol in the previous equation, I've drawn this back-tofront to make the next diagrams clearer. Normally, it is drawn with the OH groups on the right-hand side.
If you make an ester of this with ethanoic acid, you could attach three
ethanoate groups.

http://www.chemguide.co.uk/organicprops/esters/background.html (4 of 12)30/12/2004 11:28:10

an introduction to esters

Now, make the acid chains much longer, and you finally have a fat.

Note: The colour coding is still there only to help you to see
how the formulae are built up. In each case, if you want to be
strictly accurate, the final oxygen in each row should actually
be colour-coded red because it comes from the propane1,2,3-triol.
Although I have shown all the chains in the last structure as
the same for simplicity, there is no reason why the three
chains in any particular fat or oil molecule have to be the
same.

The acid CH3(CH2)16COOH is called octadecanoic acid, but the old


name is still commonly used. This is stearic acid.
The full name for the ester of this with propane-1,2,3-triol is propane1,2,3-triyl trioctadecanoate. But the truth is that almost everybody calls it
(not surprisingly!) by its old name of glyceryl tristearate.

Saturated and unsaturated fats and oils


If the fat or oil is saturated, it means that the acid that it was derived from
http://www.chemguide.co.uk/organicprops/esters/background.html (5 of 12)30/12/2004 11:28:10

an introduction to esters

has no carbon-carbon double bonds in its chain. Stearic acid is a


saturated acid, and so glyceryl tristearate is a saturated fat.
If the acid has just one carbon-carbon double bond somewhere in the
chain, it is called mono-unsaturated. If it has more than one carboncarbon double bond, it is polyunsaturated.
Those same terms will then apply to the esters that are formed.
All of these are saturated acids, and so will form saturated fats and oils:

Oleic acid is a typical mono-unsaturated acid:

. . . and linoleic and linolenic acids are typical polyunsaturated acids.

You might possibly have come across the terms "omega 6" and "omega
3" in the context of fats and oils.
Linoleic acid is an omega 6 acid. It just means that the first carboncarbon double bond starts on the sixth carbon from the CH3 end.
Linolenic acid is an omega 3 acid for the same reason.

http://www.chemguide.co.uk/organicprops/esters/background.html (6 of 12)30/12/2004 11:28:10

an introduction to esters

Note: This is quite confusing because it is exactly the


opposite of the way the acids are named systematically. The
"9" or "12" and so on in the systematic names count from the
other end! In a carboxylic acid, the carbon in the -COOH
group is counted as the number 1 carbon.

Because of their relationship with fats and oils, all of the acids above are
sometimes described as fatty acids.

Physical properties
Simple esters
I am thinking here about things like ethyl ethanoate.
Boiling points
The small esters have boiling points which are similar to those of
aldehydes and ketones with the same number of carbon atoms.
Like aldehydes and ketones, they are polar molecules and so have
dipole-dipole interactions as well as van der Waals dispersion forces.
However, they don't form hydrogen bonds, and so their boiling points
aren't anything like as high as an acid with the same number of carbon
atoms.
For example:

molecule

type

boiling point (C)

CH3COOCH2CH3

ester

77.1

CH3CH2CH2COOH

carboxylic acid

164

http://www.chemguide.co.uk/organicprops/esters/background.html (7 of 12)30/12/2004 11:28:10

an introduction to esters

Note: If you aren't happy about intermolecular forces


(including hydrogen bonds) then you really ought to follow
this link before you go on.
Use the BACK button on your browser to return to this page.

Solubility in water
The small esters are fairly soluble in water but solubility falls with chain
length.
For example:

ester

formula

solubility (g per 100


g of water)

ethyl methanoate

HCOOCH2CH3

10.5

ethyl ethanoate

CH3COOCH2CH3

8.7

ethyl propanoate

CH3CH2COOCH2CH3

1.7

The reason for the solubility is that although esters can't hydrogen bond
with themselves, they can hydrogen bond with water molecules.
One of the slightly positive hydrogen atoms in a water molecule can be
sufficiently attracted to one of the lone pairs on one of the oxygen atoms
in an ester for a hydrogen bond to be formed.
There will also, of course, be dispersion forces and dipole-dipole
attractions between the ester and the water molecules.

http://www.chemguide.co.uk/organicprops/esters/background.html (8 of 12)30/12/2004 11:28:10

an introduction to esters

Forming these attractions releases energy. This helps to supply the


energy needed to separate water molecule from water molecule and
ester molecule from ester molecule before they can mix together.
As chain lengths increase, the hydrocarbon parts of the ester molecules
start to get in the way.
By forcing themselves between water molecules, they break the
relatively strong hydrogen bonds between water molecules without
replacing them by anything as good. This makes the process
energetically less profitable, and so solubility decreases.

The physical properties of fats and oils


Solubility in water
None of these molecules are water soluble. The chain lengths are now
so great that far too many hydrogen bonds between water molecules
would have to be broken - so it isn't energetically profitable.
Melting points
The melting points determine whether the substance is a fat (a solid at
room temperature) or an oil (a liquid at room temperature).
Fats normally contain saturated chains. These allow more effective van
der Waals dispersion forces between the molecules. That means you
need more energy to separate them, and so increases the melting points.
The greater the extent of the unsaturation in the molecules, the lower the
melting points tend to be because the van der Waals dispersion forces
are less effective.
Why should this be? We are talking about molecules of very similar sizes
and so the potential for temporary dipoles should be much the same in
all of them. What matters, though, is how close together the molecules
can get.
van der Waals dispersion forces need the molecules to be able to pack
closely together to be really effective. The presence of carbon-carbon
http://www.chemguide.co.uk/organicprops/esters/background.html (9 of 12)30/12/2004 11:28:10

an introduction to esters

double bonds in the chains gets in the way of tidy packing.


Here is a simplified diagram of a saturated fat:

The hydrocarbon chains are, of course, in constant motion in the liquid,


but it is possible for them to lie tidily when the substance solidifies. If the
chains in one molecule can lie tidily, that means that neighbouring
molecules can get close.
That increases the attractions between one molecule and its neighbours
and so increases the melting point.

Unsaturated fats and oils have at least one carbon-carbon double bond
in at least one chain.
There isn't any rotation about a carbon-carbon double bond and so that
locks a permanent kink into the chain. That makes packing molecules
close together more difficult. If they don't pack so well, the van der Waals
forces won't work as well.
This effect is much worse for molecules where the hydrocarbon chains
either end of the double bond are arranged cis to each other - in other
words, both of them on the same side of the double bond:

http://www.chemguide.co.uk/organicprops/esters/background.html (10 of 12)30/12/2004 11:28:10

an introduction to esters

If they are on opposite sides of the double bond (the trans form) the
effect isn't as marked. It is, however, rather more than the diagram below
suggests because of the changes in bond angles around the double
bond compared with the rest of the chain.

Trans fats and oils have higher melting points than cis ones because the
packing isn't affected quite as much. Naturally occurring unsaturated fats
and oils tend to be the cis form.

http://www.chemguide.co.uk/organicprops/esters/background.html (11 of 12)30/12/2004 11:28:10

an introduction to esters

Note: Follow this link if you aren't sure about cis and trans
forms around a carbon-carbon double bond.
Use the BACK button on your browser to return to this page.

Where would you like to go now?


To the esters menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/esters/background.html (12 of 12)30/12/2004 11:28:10

preparation of esters

MAKING ESTERS

This page describes ways of making esters in the lab from alcohols and
phenols using carboxylic acids, acyl chlorides (acid chlorides) or acid
anhydrides as appropriate.

Making esters using carboxylic acids


This method can be used for converting alcohols into esters, but it doesn't
work with phenols - compounds where the -OH group is attached directly
to a benzene ring. Phenols react with carboxylic acids so slowly that the
reaction is unusable for preparation purposes.
The chemistry of the reaction
Esters are produced when carboxylic acids are heated with alcohols in the
presence of an acid catalyst. The catalyst is usually concentrated
sulphuric acid. Dry hydrogen chloride gas is used in some cases, but
these tend to involve aromatic esters (ones where the carboxylic acid
contains a benzene ring). If you are a UK A level student you won't have
to worry about these.
The esterification reaction is both slow and reversible. The equation for
the reaction between an acid RCOOH and an alcohol R'OH (where R and
R' can be the same or different) is:

So, for example, if you were making ethyl ethanoate from ethanoic acid
and ethanol, the equation would be:

http://www.chemguide.co.uk/organicprops/esters/preparation.html (1 of 8)30/12/2004 11:28:16

preparation of esters

Note: The mechanism for the esterification reaction is


covered in the catalysis section of this site. It is not required
for any UK A level chemistry syllabus.
If you follow this link, use the BACK button on your browser to
return to this page.

Doing the reactions


On a test tube scale
Carboxylic acids and alcohols are often warmed together in the presence
of a few drops of concentrated sulphuric acid in order to observe the smell
of the esters formed.
You would normally use small quantities of everything heated in a test
tube stood in a hot water bath for a couple of minutes.
Because the reactions are slow and reversible, you don't get a lot of ester
produced in this time. The smell is often masked or distorted by the smell
of the carboxylic acid. A simple way of detecting the smell of the ester is
to pour the mixture into some water in a small beaker.
Apart from the very small ones, esters are fairly insoluble in water and
tend to form a thin layer on the surface. Excess acid and alcohol both
dissolve and are tucked safely away under the ester layer.
Small esters like ethyl ethanoate smell like typical organic solvents (ethyl
ethanoate is a common solvent in, for example, glues).
As the esters get bigger, the smells tend towards artificial fruit flavouring "pear drops", for example.
On a larger scale

http://www.chemguide.co.uk/organicprops/esters/preparation.html (2 of 8)30/12/2004 11:28:16

preparation of esters

If you want to make a reasonably large sample of an ester, the method


used depends to some extent on the size of the ester. Small esters are
formed faster than bigger ones.
To make a small ester like ethyl ethanoate, you can gently heat a mixture
of ethanoic acid and ethanol in the presence of concentrated sulphuric
acid, and distil off the ester as soon as it is formed.
This prevents the reverse reaction happening. It works well because the
ester has the lowest boiling point of anything present. The ester is the only
thing in the mixture which doesn't form hydrogen bonds, and so it has the
weakest intermolecular forces.

Note: Follow this link if you aren't sure about hydrogen


bonding.
Use the BACK button on your browser to return to this page.

Larger esters tend to form more slowly. In these cases, it may be


necessary to heat the reaction mixture under reflux for some time to
produce an equilibrium mixture. The ester can be separated from the
carboxylic acid, alcohol, water and sulphuric acid in the mixture by
fractional distillation.

Note: Providing full details for organic preparations (including


all the steps necessary in cleaning up the product) is beyond
the scope of this site. If you need this sort of detail, you should
be looking at an organic practical book.

http://www.chemguide.co.uk/organicprops/esters/preparation.html (3 of 8)30/12/2004 11:28:16

preparation of esters

Making esters using acyl chlorides (acid chlorides)


This method will work for alcohols and phenols. In the case of phenols,
the reaction is sometimes improved by first converting the phenol into a
more reactive form.
The basic reaction
If you add an acyl chloride to an alcohol, you get a vigorous (even violent)
reaction at room temperature producing an ester and clouds of steamy
acidic fumes of hydrogen chloride.
For example, if you add the liquid ethanoyl chloride to ethanol, you get a
burst of hydrogen chloride produced together with the liquid ester ethyl
ethanoate.

The substance normally called "phenol" is the simplest of the family of


phenols. Phenol has an -OH group attached to a benzene ring - and
nothing else.
The reaction between ethanoyl chloride and phenol is similar to the
ethanol reaction although not so vigorous. Phenyl ethanoate is formed
together with hydrogen chloride gas.

http://www.chemguide.co.uk/organicprops/esters/preparation.html (4 of 8)30/12/2004 11:28:16

preparation of esters

Note: These reactions are discussed in rather more detail on


a page about reactions of acyl chlorides.
If you want the mechanism for the reaction involving alcohols
you can find it by following this link. The phenol mechanism is
similar, although hindered by the interaction between the lone
pair on the oxygen of the -OH group and the ring electrons.
If you aren't sure about using this symbol for a benzene ring,
you could follow this link to find out all about it. It is likely to
take you some time, though, and you may have to visit several
other pages as well.
It isn't particularly important in the context of the current page.
All you need to know is that at each corner of the hexagon
there is a carbon atom, together with a hydrogen atom apart
from where the oxygen is attached.
If you choose to follow any of these links, use the BACK
button (or GO menu or HISTORY file) on your browser to
return to this page.

Improving the reactions between phenols and some less reactive


acyl chlorides
Benzoyl chloride has the formula C6H5COCl. The -COCl group is attached
directly to a benzene ring. It is much less reactive than simple acyl
chlorides like ethanoyl chloride.
The phenol is first converted into the ionic compound sodium phenoxide
(sodium phenate) by dissolving it in sodium hydroxide solution.

The phenoxide ion reacts more rapidly with benzoyl chloride than the
http://www.chemguide.co.uk/organicprops/esters/preparation.html (5 of 8)30/12/2004 11:28:16

preparation of esters

original phenol does, but even so you have to shake it with benzoyl
chloride for about 15 minutes. Solid phenyl benzoate is formed.

Making esters using acid anhydrides


This reaction can again be used to make esters from both alcohols and
phenols. The reactions are slower than the corresponding reactions with
acyl chlorides, and you usually need to warm the mixture.
In the case of a phenol, you can react the phenol with sodium hydroxide
solution first, producing the more reactive phenoxide ion.

Taking ethanol reacting with ethanoic anhydride as a typical reaction


involving an alcohol:
There is a slow reaction at room temperature (or faster on warming).
There is no visible change in the colourless liquids, but a mixture of ethyl
ethanoate and ethanoic acid is formed.

The reaction with phenol is similar, but will be slower. Phenyl ethanoate is
formed together with ethanoic acid.

http://www.chemguide.co.uk/organicprops/esters/preparation.html (6 of 8)30/12/2004 11:28:16

preparation of esters

This reaction isn't important itself, but a very similar reaction is involved in
the manufacture of aspirin (covered in detail on another page - link below).
If the phenol is first converted into sodium phenoxide by adding sodium
hydroxide solution, the reaction is faster. Phenyl ethanoate is again
formed, but this time the other product is sodium ethanoate rather than
ethanoic acid.

Note: These reactions (including the formation of aspirin) are


discussed in more detail on a page about reactions of acid
anhydrides.
Unless you are already very familiar with acid anhydrides, it
would definitely be a good idea to follow this link. Use the
BACK button on your browser to return to this page later if you
want to.

Where would you like to go now?


To the esters menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

http://www.chemguide.co.uk/organicprops/esters/preparation.html (7 of 8)30/12/2004 11:28:16

preparation of esters

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/esters/preparation.html (8 of 8)30/12/2004 11:28:16

hydrolysis of esters

HYDROLYSING ESTERS

This page describes ways of hydrolysing esters - splitting them into


carboxylic acids (or their salts) and alcohols by the action of water, dilute
acid or dilute alkali. It starts by looking at the hydrolysis of simple esters
like ethyl ethanoate, and goes on to look at hydrolysing bigger, more
complicated ones to make soap.

Hydrolysing simple esters


What is hydrolysis?
Technically, hydrolysis is a reaction with water. That is exactly what
happens when esters are hydrolysed by water or by dilute acids such as
dilute hydrochloric acid.
The alkaline hydrolysis of esters actually involves reaction with hydroxide
ions, but the overall result is so similar that it is lumped together with the
other two.

Hydrolysis using water or dilute acid


The reaction with pure water is so slow that it is never used. The reaction
is catalysed by dilute acid, and so the ester is heated under reflux with a
dilute acid like dilute hydrochloric acid or dilute sulphuric acid.
Here are two simple examples of hydrolysis using an acid catalyst.
First, hydrolysing ethyl ethanoate:

http://www.chemguide.co.uk/organicprops/esters/hydrolysis.html (1 of 5)30/12/2004 11:28:21

hydrolysis of esters

. . . and then hydrolysing methyl propanoate:

Notice that the reactions are reversible. To make the hydrolysis as


complete as possible, you would have to use an excess of water. The
water comes from the dilute acid, and so you would mix the ester with an
excess of dilute acid.

Note: These reactions are exactly the reverse of those used


to make an ester from a carboxylic acid and an alcohol. The
only difference in that case is that you use a concentrated acid
as the catalyst. To get as much ester as possible, you
wouldn't add any water otherwise you would favour the
hydrolysis reaction.
The mechanism for the acid hydrolysis of esters is covered in
the catalysis section of this site. It is not required for any UK A
level chemistry syllabus.
If you follow this link, use the BACK button on your browser to
return to this page.

http://www.chemguide.co.uk/organicprops/esters/hydrolysis.html (2 of 5)30/12/2004 11:28:21

hydrolysis of esters

Hydrolysis using dilute alkali


This is the usual way of hydrolysing esters. The ester is heated under
reflux with a dilute alkali like sodium hydroxide solution.
There are two big advantages of doing this rather than using a dilute acid.
The reactions are one-way rather than reversible, and the products are
easier to separate.
Taking the same esters as above, but using sodium hydroxide solution
rather than a dilute acid:
First, hydrolysing ethyl ethanoate using sodium hydroxide solution:

. . . and then hydrolysing methyl propanoate in the same way:

Notice that you get the sodium salt formed rather than the carboxylic acid
itself.
This mixture is relatively easy to separate. Provided you use an excess of
sodium hydroxide solution, there won't be any ester left - so you don't
have to worry about that.
The alcohol formed can be distilled off. That's easy!
If you want the acid rather than its salt, all you have to do is to add an
excess of a strong acid like dilute hydrochloric acid or dilute sulphuric acid
http://www.chemguide.co.uk/organicprops/esters/hydrolysis.html (3 of 5)30/12/2004 11:28:21

hydrolysis of esters

to the solution left after the first distillation.


If you do this, the mixture is flooded with hydrogen ions. These are picked
up by the ethanoate ions (or propanoate ions or whatever) present in the
salts to make ethanoic acid (or propanoic acid, etc). Because these are
weak acids, once they combine with the hydrogen ions, they tend to stay
combined.
The carboxylic acid can now be distilled off.

Hydrolysing complicated esters to make soap


This next bit deals with the alkaline hydrolysis (using sodium hydroxide
solution) of the big esters found in animal and vegetable fats and oils.

Important: If you haven't already read that page, you should


read the introduction to esters so that you understand the
nature of the fats and oils that are coming up next.
Use the BACK button on your browser to return to this page.

If the large esters present in animal or vegetable fats and oils are heated
with concentrated sodium hydroxide solution exactly the same reaction
happens as with the simple esters.
A salt of a carboxylic acid is formed - in this case, the sodium salt of a big
acid such as octadecanoic acid (stearic acid). These salts are the
important ingredients of soap - the ones that do the cleaning.
An alcohol is also produced - in this case, the more complicated alcohol,
propane-1,2,3-triol (glycerol).

http://www.chemguide.co.uk/organicprops/esters/hydrolysis.html (4 of 5)30/12/2004 11:28:21

hydrolysis of esters

Because of its relationship with soap making, the alkaline hydrolysis of


esters is sometimes known as saponification.

Where would you like to go now?


To the esters menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/esters/hydrolysis.html (5 of 5)30/12/2004 11:28:21

polyesters - terylene and PET

POLYESTERS

This page looks at the formation, structure and uses of a common


polyester sometimes known as Terylene if it is used as a fibre, or PET if
it used in, for example, plastic drinks bottles

Poly(ethylene terephthalate)
What is a polyester?
A polyester is a polymer (a chain of repeating units) where the individual
units are held together by ester linkages.

The diagram shows a very small bit of the polymer chain and looks pretty
complicated. But it isn't very difficult to work out - and that's the best thing
to do: work it out, not try to remember it. You will see how to do that in a
moment.
The usual name of this common polyester is poly(ethylene
terephthalate). The everyday name depends on whether it is being used
as a fibre or as a material for making things like bottles for soft drinks.
When it is being used as a fibre to make clothes, it is often just called
polyester. It may sometimes be known by a brand name like Terylene.
When it is being used to make bottles, for example, it is usually called
PET.

http://www.chemguide.co.uk/organicprops/esters/polyesters.html (1 of 6)30/12/2004 11:28:26

polyesters - terylene and PET

Making polyesters as an example of condensation polymerisation


In condensation polymerisation, when the monomers join together a
small molecule gets lost. That's different from addition polymerisation
which produces polymers like poly(ethene) - in that case, nothing is lost
when the monomers join together.
A polyester is made by a reaction involving an acid with two -COOH
groups, and an alcohol with two -OH groups.
In the common polyester drawn above:
The acid is benzene-1,4-dicarboxylic acid (old name: terephthalic acid).
The alcohol is ethane-1,2-diol (old name: ethylene glycol).

Now imagine lining these up alternately and making esters with each
acid group and each alcohol group, losing a molecule of water every time
an ester linkage is made.

That would produce the chain shown above (although this time written
without separating out the carbon-oxygen double bond - write it
whichever way you like).

http://www.chemguide.co.uk/organicprops/esters/polyesters.html (2 of 6)30/12/2004 11:28:26

polyesters - terylene and PET

Note: This does NOT describe the way the actual


reaction happens - it is a way of working out the
structure of the polymer. The chemistry of the reaction is
more complicated than this (see below).
The diagram shows a slightly shorter bit of chain than the
corresponding one at the top of the page. However, it is
exactly consistent with the loss of water from the last
diagram. It was impossible to include another ethane-1,2-diol
in that diagram for space reasons. If any of this offends you,
draw it again yourself so that everything matches! In fact, it
would be good practice to draw a bit of chain starting from a
few more monomers.
This is what I meant further up the page by working the
structure out rather than remembering it. The structures of
both monomers are easy to remember. If you line them up
and remove water as I have shown, the structure follows
automatically.

Manufacturing poly(ethylene terephthalate)


The reaction takes place in two main stages: a pre-polymerisation stage
and the actual polymerisation.

Warning! This manufacturing process is only currently


required by one UK A level Exam Board (WJEC). If you don't
need to know about this, skip over the next bit. It is really
confusing, because it doesn't relate easily to the way we
have used to work out the structure of the polyester. The
overall result is the same, but it happens by a much more
complicated process.

http://www.chemguide.co.uk/organicprops/esters/polyesters.html (3 of 6)30/12/2004 11:28:26

polyesters - terylene and PET

In the first stage, before polymerisation happens, you get a fairly simple
ester formed between the acid and two molecules of ethane-1,2-diol.

In the polymerisation stage, this is heated to a temperature of about 260


C and at a low pressure. A catalyst is needed - there are several
possibilities including antimony compounds like antimony(III) oxide.
The polyester forms and half of the ethane-1,2-diol is regenerated. This
is removed and recycled.

Note: Notice the way the polymer is drawn. This is the


minimum amount of chain that you can draw to show the
repeating unit.

http://www.chemguide.co.uk/organicprops/esters/polyesters.html (4 of 6)30/12/2004 11:28:26

polyesters - terylene and PET

Hydrolysis of polyesters
Simple esters are easily hydrolysed by reaction with dilute acids or
alkalis.
Polyesters are attacked readily by alkalis, but much more slowly by dilute
acids. Hydrolysis by water alone is so slow as to be completely
unimportant. (You wouldn't expect your polyester fleece to fall to pieces if
you went out in the rain!)
If you spill dilute alkali on a fabric made from polyester, the ester linkages
are broken. Ethane-1,2-diol is formed together with the salt of the
carboxylic acid.
Because you produce small molecules rather than the original polymer,
the fibres are destroyed, and you end up with a hole!
For example, if you react the polyester with sodium hydroxide solution:

Note: Hydrolysis of esters is covered in detail on another


page in this section.

http://www.chemguide.co.uk/organicprops/esters/polyesters.html (5 of 6)30/12/2004 11:28:26

polyesters - terylene and PET

Where would you like to go now?


To the esters menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/esters/polyesters.html (6 of 6)30/12/2004 11:28:26

triiodomethane (iodoform) reaction with alcohols

THE TRIIODOMETHANE (IODOFORM) REACTION


WITH ALCOHOLS

This page looks at how the triiodomethane (iodoform) reaction can be


used to identify the presence of a CH3CH(OH) group in alcohols.

Note: This reaction can also be used in testing for the


CH3CO group in aldehydes and ketones. You will find a link
to this at the bottom of the page.

Doing the triiodomethane (iodoform) reaction


There are two apparently quite different mixtures of reagents that can be
used to do this reaction. They are, in fact, chemically equivalent.

Note: It would be silly to learn both of these methods. Use


whichever one your examiners want - find out by looking at
past papers and mark schemes. If you haven't got these, go
to the syllabuses page to find out how to get hold of them.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/alcohols/iodoform.html (1 of 5)30/12/2004 11:28:30

triiodomethane (iodoform) reaction with alcohols

Using iodine and sodium hydroxide solution


This is chemically the more obvious method.
Iodine solution is added to a small amount of an alcohol, followed by just
enough sodium hydroxide solution to remove the colour of the iodine. If
nothing happens in the cold, it may be necessary to warm the mixture
very gently.
A positive result is the appearance of a very pale yellow precipitate of
triiodomethane (previously known as iodoform) - CHI3.
Apart from its colour, this can be recognised by its faintly "medical" smell.
It is used as an antiseptic on the sort of sticky plasters you put on minor
cuts, for example.
Using potassium iodide and sodium chlorate(I) solutions
Sodium chlorate(I) is also known as sodium hypochlorite.
Potassium iodide solution is added to a small amount of an alcohol,
followed by sodium chlorate(I) solution. Again, if no precipitate is formed
in the cold, it may be necessary to warm the mixture very gently.
The positive result is the same pale yellow precipitate as before.

Why the two reactions are equivalent: This reaction


happens in three stages. In the first, the alcohol is oxidised to
an aldehyde or ketone. In the first mixture, the iodine reacts
with the sodium hydroxide solution to produce some sodium
iodate(I). This is an oxidising agent.
In the second mixture, the sodium chlorate(I) already present
is an oxidising agent.
After that the reaction happens in two further stages: first the
aldehyde or ketone formed reacts with iodine, and the
product of that reaction reacts with hydroxide ions. Iodine and
sodium hydroxide is exactly what you are adding in the first
method above.
http://www.chemguide.co.uk/organicprops/alcohols/iodoform.html (2 of 5)30/12/2004 11:28:30

triiodomethane (iodoform) reaction with alcohols

In the second method, the sodium chlorate(I) solution is an


oxidising agent, and oxidises the iodide ions in the potassium
iodide to iodine. As well as any possible precipitate, you will
also see the typical reddish-brown colour of iodine solution
being formed during the reaction.
So although you didn't put any iodine into the mixture, it is
made in situ. What about the hydroxide ions?
Sodium chlorate(I) solution is alkaline and contains enough
hydroxide ions to carry out the second half of the reaction.
Sodium chlorate(I) is alkaline because it reacts reversibly
with water to form the weak acid chloric(I) acid together with
hydroxide ions.

The chemistry of the triiodomethane (iodoform) reaction


What the triiodomethane (iodoform) reaction shows
A positive result - the pale yellow precipitate of triiodomethane (iodoform)
- is given by an alcohol containing the grouping:

"R" can be a hydrogen atom or a hydrocarbon group (for example, an


alkyl group).
If "R" is hydrogen, then you have the primary alcohol ethanol,
CH3CH2OH.

Ethanol is the only primary alcohol to give the triiodomethane


(iodoform) reaction.
If "R"is a hydrocarbon group, then you have a secondary alcohol.

http://www.chemguide.co.uk/organicprops/alcohols/iodoform.html (3 of 5)30/12/2004 11:28:30

triiodomethane (iodoform) reaction with alcohols

Lots of secondary alcohols give this reaction, but those that do all
have a methyl group attached to the carbon with the -OH group.

No tertiary alcohols can contain this group because no tertiary


alcohols can have a hydrogen atom attached to the carbon with
the -OH group. No tertiary alcohols give the triiodomethane
(iodoform) reaction.

Summary of the reactions during the triiodomethane (iodoform)


reaction
We will take the reagents as being iodine and sodium hydroxide solution.

This is being given as a flow scheme rather than full equations. You
aren't likely to need the equation for the oxidation stage for UK A level
purposes. The equations for the other two steps are given on a page
about reactions of aldehydes and ketones. Follow the first link below if
http://www.chemguide.co.uk/organicprops/alcohols/iodoform.html (4 of 5)30/12/2004 11:28:30

triiodomethane (iodoform) reaction with alcohols

you are interested.

Where would you like to go now?


Using the same reaction with aldehydes and ketones . . .
To the alcohols menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/alcohols/iodoform.html (5 of 5)30/12/2004 11:28:30

triiodomethane (iodoform) reaction with aldehydes and ketones

THE TRIIODOMETHANE (IODOFORM) REACTION


WITH ALDEHYDES AND KETONES

This page looks at how the triiodomethane (iodoform) reaction can be


used to identify the presence of a CH3CO group in aldehydes and
ketones.

Note: This reaction can also be used in testing for the


CH3CH(OH) group in alcohols. You will find a link to this at
the bottom of the page.

Doing the triiodomethane (iodoform) reaction


There are two apparently quite different mixtures of reagents that can be
used to do this reaction. They are, in fact, chemically equivalent.

Note: It would be silly to learn both of these methods. Use


whichever one your examiners want - find out by looking at
past papers and mark schemes. If you haven't got these, go
to the syllabuses page to find out how to get hold of them.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/carbonyls/iodoform.html (1 of 5)30/12/2004 11:28:36

triiodomethane (iodoform) reaction with aldehydes and ketones

Using iodine and sodium hydroxide solution


This is chemically the more obvious method.
Iodine solution is added to a small amount of aldehyde or ketone,
followed by just enough sodium hydroxide solution to remove the colour
of the iodine. If nothing happens in the cold, it may be necessary to warm
the mixture very gently.
A positive result is the appearance of a very pale yellow precipitate of
triiodomethane (previously known as iodoform) - CHI3.
Apart from its colour, this can be recognised by its faintly "medical" smell.
It is used as an antiseptic on the sort of sticky plasters you put on minor
cuts, for example.
Using potassium iodide and sodium chlorate(I) solutions
Sodium chlorate(I) is also known as sodium hypochlorite.
Potassium iodide solution is added to a small amount of aldehyde or
ketone, followed by sodium chlorate(I) solution. Again, if no precipitate is
formed in the cold, it may be necessary to warm the mixture very gently.
The positive result is the same pale yellow precipitate as before.

Why the two reactions are equivalent: The reaction


happens in two stages: first the aldehyde or ketone reacts
with iodine, and the product of that reaction reacts with
hydroxide ions. That is obviously the mixture you are adding
in the first method above.
In the second method, the sodium chlorate(I) solution is an
oxidising agent, and oxidises the iodide ions in the potassium
iodide to iodine. As well as any possible precipitate, you will
also see the typical reddish-brown colour of iodine solution
being formed during the reaction.
So although you didn't put any iodine into the mixture, it is
made in situ. What about the hydroxide ions?
http://www.chemguide.co.uk/organicprops/carbonyls/iodoform.html (2 of 5)30/12/2004 11:28:36

triiodomethane (iodoform) reaction with aldehydes and ketones

Sodium chlorate(I) solution is alkaline and contains enough


hydroxide ions to carry out the second half of the reaction.
The reason that sodium chlorate(I) is alkaline is that it reacts
reversibly with water to form the weak acid chloric(I) acid
together with hydroxide ions.

The chemistry of the triiodomethane (iodoform) reaction


What the triiodomethane (iodoform) reaction shows
A positive result - the pale yellow precipitate of triiodomethane (iodoform)
- is given by an aldehyde or ketone containing the grouping:

"R" can be a hydrogen atom or a hydrocarbon group (for example, an


alkyl group).
If "R" is hydrogen, then you have the aldehyde ethanal, CH3CHO.

Ethanal is the only aldehyde to give the triiodomethane (iodoform)


reaction.
If "R"is a hydrocarbon group, then you have a ketone. Lots of
ketones give this reaction, but those that do all have a methyl
group on one side of the carbon-oxygen double bond. These are
known as methyl ketones.

Equations for the triiodomethane (iodoform) reaction


We will take the reagents as being iodine and sodium hydroxide solution.

http://www.chemguide.co.uk/organicprops/carbonyls/iodoform.html (3 of 5)30/12/2004 11:28:36

triiodomethane (iodoform) reaction with aldehydes and ketones

The first stage involves substitution of all three hydrogens in the methyl
group by iodine atoms. The presence of hydroxide ions is important for
the reaction to happen - they take part in the mechanism for the reaction
(not required for UK A level).

In the second stage, the bond between the C I3 and the rest of the
molecule is broken to produce triiodomethane (iodoform) and the salt of
an acid.

Putting all this together gives the overall equation for the reaction:

Where would you like to go now?


Using the same reaction with alcohols . . .
To the aldehydes and ketones menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

http://www.chemguide.co.uk/organicprops/carbonyls/iodoform.html (4 of 5)30/12/2004 11:28:36

triiodomethane (iodoform) reaction with aldehydes and ketones

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/carbonyls/iodoform.html (5 of 5)30/12/2004 11:28:36

uses of alcohols

USES OF ALCOHOLS

This page has a brief look at some of the more important uses of the
simple alcohols like methanol, ethanol and propan-2-ol.

Uses of ethanol
Drinks
The "alcohol" in alcoholic drinks is simply ethanol.
Industrial methylated spirits (meths)
Ethanol is usually sold as industrial methylated spirits which is ethanol
with a small quantity of methanol added and possibly some colour.
Methanol is poisonous, and so the industrial methylated spirits is unfit to
drink. This avoids the high taxes which are levied on alcoholic drinks
(certainly in the UK!).
As a fuel
Ethanol burns to give carbon dioxide and water and can be used as a
fuel in its own right, or in mixtures with petrol (gasoline). "Gasohol" is a
petrol / ethanol mixture containing about 10 - 20% ethanol.
Because ethanol can be produced by fermentation, this is a useful way
for countries without an oil industry to reduce imports of petrol.

As a solvent
Ethanol is widely used as a solvent. It is relatively safe, and can be used
to dissolve many organic compounds which are insoluble in water. It is
used, for example, in many perfumes and cosmetics.

http://www.chemguide.co.uk/organicprops/alcohols/uses.html (1 of 3)30/12/2004 11:28:39

uses of alcohols

Uses of methanol
As a fuel
Methanol again burns to form carbon dioxide and water.

It can be used a a petrol additive to improve combustion, or work is


currently being done on its use as a fuel in its own right.
As an industrial feedstock
Most methanol is used to make other things - for example, methanal
(formaldehyde), ethanoic acid, and methyl esters of various acids. In
most cases, these are in turn converted into further products.

Uses of propan-2-ol
Propan-2-ol is widely used in an amazing number of different situations
as a solvent. Details on this are probably not required by UK A level
syllabuses, but if you need them, an internet search on propan-2-ol
solvent uses will give you more examples than you can cope with! (There
is a Google search box at the bottom of the Main Menu.)

Where would you like to go now?


To the alcohols menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

http://www.chemguide.co.uk/organicprops/alcohols/uses.html (2 of 3)30/12/2004 11:28:39

uses of alcohols

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/alcohols/uses.html (3 of 3)30/12/2004 11:28:39

more examples of catalysis in organic chemistry

MORE EXAMPLES OF CATALYSIS IN ORGANIC


CHEMISTRY

This page looks a few odds and ends of examples of catalysts used in
organic chemistry. It includes the formation of epoxyethane from ethene,
and several reactions from benzene chemistry - Friedel-Crafts reactions
and halogenation. You will find links to other parts of the site for the
mechanisms of the benzene reactions.
Other examples of catalysts in organic chemistry can be found from the
catalysis menu. There is a link to that menu at the bottom of the page if
you have come direct to this page via a search engine.

The manufacture of epoxyethane from ethene


Epoxyethane is manufactured by reacting ethene with a limited amount
of oxygen in the presence of a silver catalyst at a temperature of about
250 - 300C and a pressure of less than 15 atmospheres. Because the
solid silver is catalysing a gas reaction, this is an example of
heterogeneous catalysis.

The reaction is exothermic and the temperature has to be carefully


controlled to prevent further oxidation of the ethene to carbon dioxide
and water.

http://www.chemguide.co.uk/physical/catalysis/organic.html (1 of 6)30/12/2004 11:28:51

more examples of catalysis in organic chemistry

Note: If you have read the introductory page on catalysis,


you might remember that silver is quoted as a metal which
isn't a good catalyst because it doesn't adsorb reactant
molecules strongly enough, and yet here it is being used as a
catalyst.
If anyone reading this knows of an interesting reason for this
discrepancy (other than the fact that it just happens to work in
this case!), could you please let me know via the address on
the about this site page.

The halogenation of benzene


Benzene reacts with chlorine or bromine in the presence of a catalyst.
The catalyst is either aluminium chloride (or aluminium bromide if you are
reacting benzene with bromine) or iron.
Strictly speaking iron isn't a catalyst, because it gets permanently
changed during the reaction. It reacts with some of the chlorine or
bromine to form iron(III) chloride, FeCl3, or iron(III) bromide, FeBr3.

These compounds act as the catalyst and behave exactly like aluminium
chloride in these reactions.
The reaction with chlorine
The reaction between benzene and chlorine in the presence of either
aluminium chloride or iron gives chlorobenzene.

or:

http://www.chemguide.co.uk/physical/catalysis/organic.html (2 of 6)30/12/2004 11:28:51

more examples of catalysis in organic chemistry

The reaction with bromine


The reaction between benzene and bromine in the presence of either
aluminium bromide or iron gives bromobenzene. Iron is usually used
because it is cheaper and more readily available.

or:

Note: If you have want the mechanism for the halogenation


reaction, you should follow this link.
Use the BACK button on your browser to return quickly to this
page if you want to.

The Friedel-Crafts alkylation of benzene


Alkylation involves replacing a hydrogen atom on a benzene ring by an
alkyl group like methyl or ethyl. This is another example of the use of
aluminium chloride as a catalyst.
Benzene is treated with a chloroalkane (for example, chloromethane or
chloroethane) in the presence of aluminium chloride as a catalyst. The
equation shows the reaction using a methyl group, but any other alkyl
group could be used in the same way.

http://www.chemguide.co.uk/physical/catalysis/organic.html (3 of 6)30/12/2004 11:28:51

more examples of catalysis in organic chemistry

Substituting a methyl group gives methylbenzene - once known as


toluene.

or:

Note: If you have want the mechanism for the alkylation


reaction, you should follow this link.
Use the BACK button on your browser to return quickly to this
page if you want to.

The Friedel-Crafts acylation of benzene


An acyl group is an alkyl group attached to a carbon-oxygen double
bond. Acylation means substituting an acyl group into something - in this
case, into a benzene ring.
The most commonly used acyl group is CH3CO-. This is called the
ethanoyl group. In the example which follows we are substituting a
CH3CO- group into the ring, but you could equally well use any other
alkyl group instead of the CH3.

The most reactive substance containing an acyl group is an acyl chloride


(also known as an acid chloride).

http://www.chemguide.co.uk/physical/catalysis/organic.html (4 of 6)30/12/2004 11:28:51

more examples of catalysis in organic chemistry

Benzene is treated with a mixture of ethanoyl chloride, CH3COCl, and


aluminium chloride as the catalyst. A ketone called phenylethanone is
formed.

Note: Ketones: A family of compounds containing a carbonoxygen double bond with a hydrocarbon group either side of
it. In this case there is a methyl group on one side and a
benzene ring on the other.
Don't worry too much about the name "phenylethanone" - all
that matters is that you can draw the structure.

or:

Note: If you have want the mechanism for the acylation


reaction, you should follow this link.
Use the BACK button on your browser to return quickly to this
page if you want to.

http://www.chemguide.co.uk/physical/catalysis/organic.html (5 of 6)30/12/2004 11:28:51

more examples of catalysis in organic chemistry

Where would you like to go now?


To the catalysis menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

You might also be interested in:


properties and reactions of alkenes . . .
Covers all the physical and chemical properties of alkenes required
by UK A level syllabuses.
properties and reactions of arenes (aromatic hydrocarbons) . . .
A similar survey of aromatic hydrocarbons such as benzene and
methylbenzene.
properties and reactions of acyl chlorides (acid chlorides) . . .
. . . and of acyl chlorides.

Jim Clark 2002 (modified 2004)

http://www.chemguide.co.uk/physical/catalysis/organic.html (6 of 6)30/12/2004 11:28:51

Redox equilibria menu

Understanding Chemistry

REDOX EQUILIBRIA MENU

These pages are designed to be read in sequence. In order to


understand the topic fully, you really ought to read them through in this
order. At the bottom of each page you will find a link to the next one.
Standard electrode potentials (redox potentials) . . .
Explains how electrode potentials of metal / metal ion systems
arise, and how they are measured. Introduces the important
conventions used.
The electrochemical series . . .
Shows how ranking electrode potentials in order leads to the
electrochemical series, and the relationship between this and
oxidation and reduction.
Redox potentials for other systems . . .
Explains how the electrode potentials (redox potentials) of more
complicated systems can be measured and related to the
electrochemical series.
Combinations of half cells . . .
Shows the relationship between combinations of half cells and
simple redox reactions carried out in test tubes.
Using redox potentials to predict the feasibility of reactions . . .
Explains how to use values of redox potentials to predict whether
http://www.chemguide.co.uk/physical/redoxeqiamenu.html (1 of 2)30/12/2004 11:28:52

Redox equilibria menu

particular redox reactions are feasible.

Go to physical chemistry menu . . .


Go to Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/redoxeqiamenu.html (2 of 2)30/12/2004 11:28:52

an introduction to redox equilibria and electrode potentials

AN INTRODUCTION TO REDOX EQUILIBRIA and


ELECTRODE POTENTIALS

This page explains the background to standard electrode potentials


(redox potentials), showing how they arise from simple equilibria, and
how they are measured.
There are as many ways of teaching this as there are teachers and
writers, and too many people make the fundamental mistake of forgetting
that these are just simple equilibria. Too often, you will find the equations
involved written as one-way rather than reversible. That small mistake
makes the whole topic quite unnecessarily difficult to understand.
The approach you will find on this page (and whenever redox potentials
are discussed on this site) avoids this problem completely by always
talking in terms of equilibria.

Important: If you aren't too happy about simple equilibria


(particularly about Le Chatelier's Principle), you should
explore the equilibrium section of this site before you go any
further.
The whole of this topic would also be a nightmare if you didn't
understand about redox reactions.
If you need to explore these sections in any detail, return to
this page via the site menus, or use the BACK button or the
History or Go menus on your browser.

http://www.chemguide.co.uk/physical/redoxeqia/introduction.html (1 of 19)30/12/2004 11:29:27

an introduction to redox equilibria and electrode potentials

Background
The differing reactivities of metals
When metals react, they give away electrons and form positive ions. This
particular topic sets about comparing the ease with which a metal does
this to form hydrated ions in solution - for example, Mg2+(aq) or Cu2+(aq).
We might want to compare the ease with which these two changes take
place:

Everybody who has done chemistry for more than a few months knows
that magnesium is more reactive than copper. The first reaction happens
much more readily than the second one. What this topic does is to try to
express this with some numbers.
Looking at this from an equilibrium point of view
Suppose you have a piece of magnesium in a beaker of water. There will
be some tendency for the magnesium atoms to shed electrons and go
into solution as magnesium ions. The electrons will be left behind on the
magnesium.

http://www.chemguide.co.uk/physical/redoxeqia/introduction.html (2 of 19)30/12/2004 11:29:27

an introduction to redox equilibria and electrode potentials

In a very short time, there will be a build-up of electrons on the


magnesium, and it will be surrounded in the solution by a layer of positive
ions. These will tend to stay close because they are attracted to the
negative charge on the piece of metal.
Some of them will be attracted enough that they will reclaim their
electrons and stick back on to the piece of metal.

http://www.chemguide.co.uk/physical/redoxeqia/introduction.html (3 of 19)30/12/2004 11:29:27

an introduction to redox equilibria and electrode potentials

A dynamic equilibrium will be established when the rate at which ions are
leaving the surface is exactly equal to the rate at which they are joining it
again. At that point there will be a constant negative charge on the
magnesium, and a constant number of magnesium ions present in the
solution around it.
Simplifying the diagram to get rid of the "bites" out of the magnesium,
you would be left with a situation like this:

http://www.chemguide.co.uk/physical/redoxeqia/introduction.html (4 of 19)30/12/2004 11:29:27

an introduction to redox equilibria and electrode potentials

Don't forget that this is just a snapshot of a dynamic equilibrium. Ions are
continually leaving and rejoining the surface.
How would this be different if you used a piece of copper instead of a
piece of magnesium?
Copper is less reactive and so forms its ions less readily. Any ions which
do break away are more likely to reclaim their electrons and stick back
on to the metal again. You will still reach an equilibrium position, but
there will be less charge on the metal, and fewer ions in solution.

http://www.chemguide.co.uk/physical/redoxeqia/introduction.html (5 of 19)30/12/2004 11:29:27

an introduction to redox equilibria and electrode potentials

If we write the two reactions as equilibria, then what we are doing is


comparing the two positions of equilibrium.
The position of the magnesium equilibrium . . .

. . . lies further to the left than that of the copper equilibrium.

Notice the way that the two equilibria are written. By convention, all these
equilibria are written with the electrons on the left-hand side of the
equation. If you stick with this convention without fail, you will find that it
makes the rest of this topic much easier to visualise.
Everything else concerning electrode potentials is simply an attempt to
attach some numbers to these differing positions of equilibrium.
In principle, that is quite easy to do. In the magnesium case, there is a lot
of difference between the negativeness of the metal and the positiveness
of the solution around it. In the copper case, the difference is much less.
This potential difference could be recorded as a voltage - the bigger the
difference between the positiveness and the negativeness, the bigger the
voltage. Unfortunately, that voltage is impossible to measure!

http://www.chemguide.co.uk/physical/redoxeqia/introduction.html (6 of 19)30/12/2004 11:29:27

an introduction to redox equilibria and electrode potentials

It would be easy to connect a voltmeter to the piece of metal, but how


would you make a connection to the solution? By putting a probe into the
solution near the metal? No - it wouldn't work!
Any probe you put in is going to have a similar sort of equilibrium
happening around it. The best you could measure would be some sort of
combination of the effects at the probe and the piece of metal you are
testing.

Understanding the ideas behind a reference electrode


Suppose you had an optical device for measuring heights some distance
away, and wanted to use it to find out how tall a particular person was.
Unfortunately, you can't see their feet because they are standing in long
grass.

Although you can't measure their absolute height, what you can do is to
measure their height relative to the convenient post. Suppose that in this
case, the person turned out to be 15 cm taller than the post.
You could repeat this for a range of people . . .

http://www.chemguide.co.uk/physical/redoxeqia/introduction.html (7 of 19)30/12/2004 11:29:28

an introduction to redox equilibria and electrode potentials

. . . and come up with a set of results like this:

person

height relative to post (cm)

+20

+15

-15

Although you don't know any of their absolute heights, you can usefully
rank them in order, and do some very simple sums to work out exactly
how much taller one is than another. For example, C is 5 cm taller than A.

This turns out to be exactly what we need to do with the equilibria we


started talking about. We don't actually need to know the absolute
position of any of these equilibria. Going back to the magnesium and
copper equilibria:

All we need to know is that the magnesium equilibrium lies further to the
left than the copper one. We need to know that magnesium sheds
electrons and forms ions more readily than copper does.
That means that we don't need to be able to measure the absolute
voltage between the metal and the solution. It is enough to compare the
voltage with a standardised system called a reference electrode.
The system used is called a standard hydrogen electrode.

http://www.chemguide.co.uk/physical/redoxeqia/introduction.html (8 of 19)30/12/2004 11:29:28

an introduction to redox equilibria and electrode potentials

Measuring standard electrode potentials (standard redox


potentials)

Note: It is going to take a while before these terms get


defined. Be patient! It is more important to fully understand
what is going on first.

The standard hydrogen electrode


The standard hydrogen electrode looks like this:

What is happening?
As the hydrogen gas flows over the porous platinum, an equilibrium is set
up between hydrogen molecules and hydrogen ions in solution. The
reaction is catalysed by the platinum.

This is the equilibrium that we are going to compare all the others with.
Standard conditions

http://www.chemguide.co.uk/physical/redoxeqia/introduction.html (9 of 19)30/12/2004 11:29:28

an introduction to redox equilibria and electrode potentials

The position of any equilibrium can be changed by changing conditions.


That means that the conditions must be standardised so that you can
make fair comparisons.
The hydrogen pressure is 1 bar (100 kPa). (You may find 1 atmosphere
quoted in older sources.) The temperature is 298 K (25C).
The concentration of the hydrogen ions in solution is also important.
Changing concentrations is one of the ways of changing the position of
an equilibrium. Throughout this topic, all ions concentrations are taken as
being 1 mol dm-3.
Using the standard hydrogen electrode
The standard hydrogen electrode is attached to the electrode system you
are investigating - for example, a piece of magnesium in a solution
containing magnesium ions.

Cells and half cells


The whole of this set-up is described as a cell. It is a simple system
http://www.chemguide.co.uk/physical/redoxeqia/introduction.html (10 of 19)30/12/2004 11:29:28

an introduction to redox equilibria and electrode potentials

which generates a voltage. Each of the two beakers and their contents
are described as half cells.
The salt bridge
The salt bridge is included to complete the electrical circuit but without
introducing any more bits of metal into the system. It is just a glass tube
filled with an electrolyte like potassium nitrate solution. The ends are
"stoppered" by bits of cotton wool. This stops too much mixing of the
contents of the salt bridge with the contents of the two beakers.
The electrolyte in the salt bridge is chosen so that it doesn't react with the
contents of either beaker.
What happens?
These two equilibria are set up on the two electrodes (the magnesium
and the porous platinum):

Magnesium has a much greater tendency to form its ions than hydrogen
does. The position of the magnesium equilibrium will be well to the left of
that of the hydrogen equilibrium.
That means that there will be a much greater build-up of electrons on the
piece of magnesium than on the platinum. Stripping all the rest of the
diagram out, apart from the essential bits:

http://www.chemguide.co.uk/physical/redoxeqia/introduction.html (11 of 19)30/12/2004 11:29:28

an introduction to redox equilibria and electrode potentials

There is a major difference between the charge on the two electrodes - a


potential difference which can be measured with a voltmeter. The voltage
measured would be 2.37 volts and the voltmeter would show the
magnesium as the negative electrode and the hydrogen electrode as
being positive.
This sometimes confuses people! Obviously, the platinum in the
hydrogen electrode isn't positive in real terms - there is a slight excess of
electrons built up on it. But voltmeters don't deal in absolute terms - they
simply measure a difference.
The magnesium has the greater amount of negativeness - the voltmeter
records that as negative. The platinum of the hydrogen electrode isn't as
negative - it is relatively more positive. The voltmeter records it as
positive.
Throughout the whole of this redox potential work, you have to think in
relative terms. For example, +0.4 is relatively more negative than +1.2.
Or, another example, -0.3 is relatively more positive than -0.9.
What if you replace the magnesium half cell by a copper one?
This means replacing the magnesium half cell by one with a piece of
copper suspended in a solution containing Cu2+ ions with a concentration
of 1 mol dm-3. You would probably choose to use copper(II) sulphate
http://www.chemguide.co.uk/physical/redoxeqia/introduction.html (12 of 19)30/12/2004 11:29:28

an introduction to redox equilibria and electrode potentials

solution.
Copper forms its ions less readily than hydrogen does. Of the two
equilibria . . .

. . . the hydrogen one lies further to the left. That means that there will be
less build-up of electrons on the copper than there is on the platinum of
the hydrogen electrode.

There is less difference between the electrical charges on the two


electrodes, so the voltage measured will be less. This time it is only 0.34
volts.
The other major change is that this time the copper is the more positive
(less negative) electrode. The voltmeter will show the hydrogen electrode
as the negative one and the copper electrode as positive.
The voltmeter
You may have noticed that the voltmeter was described further up the
page as having a "high resistance". Ideally, it wants to have an infinitely
http://www.chemguide.co.uk/physical/redoxeqia/introduction.html (13 of 19)30/12/2004 11:29:28

an introduction to redox equilibria and electrode potentials

high resistance.
This is to avoid any flow of current through the circuit. If there was a low
resistance in the circuit, electrons would flow from where there are a lot
of them (around the magnesium, for example) to where there are less
(on the hydrogen electrode).
If any current flows, the voltage measured drops. In order to make proper
comparisons, it is important to measure the maximum possible voltage in
any situation. This is called the electromotive force or emf.
The emf of a cell measured under standard conditions is given the
symbol Ecell.

Note: You read E as "E nought" or "E standard".


For technical reasons to do with the way that people use
different browsers which may be set to display text differently,
it has been difficult to think of a way of showing the
"standard" symbol in the text that will display reliably in all
browsers. The symbol should actually be a circle with a
horizontal line through it.

Cell conventions
A quick way of drawing a cell
Drawing a full diagram to represent a cell takes too long. Instead, the cell
in which a magnesium electrode is coupled to a hydrogen electrode is
represented like this:

http://www.chemguide.co.uk/physical/redoxeqia/introduction.html (14 of 19)30/12/2004 11:29:28

an introduction to redox equilibria and electrode potentials

You will often find variants on the way the hydrogen electrode is
represented, such as:

Attaching a sign to the cell voltage


The convention is that you show the sign of the right-hand electrode
(as you have drawn it) when you quote the Ecell value. For example:

In the copper case:

http://www.chemguide.co.uk/physical/redoxeqia/introduction.html (15 of 19)30/12/2004 11:29:28

an introduction to redox equilibria and electrode potentials

Defining standard electrode potential (standard redox potential)


The values that we have just quoted for the two cells are actually the
standard electrode potentials of the Mg2+ / Mg and Cu2+ / Cu systems.
The emf measured when a metal / metal ion electrode is coupled to a
hydrogen electrode under standard conditions is known as the standard
electrode potential of that metal / metal ion combination.
By convention, the hydrogen electrode is always written as the left-hand
electrode of the cell. That means that the sign of the voltage quoted
always gives you the sign of the metal electrode.
Standard electrode potential is given the symbol E.

Note: In case you are wondering about the alternative name


(standard redox potential), this comes from the fact that loss
or gain of electrons is a redox reaction. This will be explored
in later pages in this series of linked pages.

http://www.chemguide.co.uk/physical/redoxeqia/introduction.html (16 of 19)30/12/2004 11:29:28

an introduction to redox equilibria and electrode potentials

Summarizing what standard electrode potentials tell you


Remember that the standard electrode potential of a metal / metal ion
combination is the emf measured when that metal / metal ion electrode is
coupled to a hydrogen electrode under standard conditions.
What you are doing is comparing the position of the metal / metal ion
equilibrium with the equilibrium involving hydrogen.
Here are a few typical standard electrode potentials:

metal / metal ion combination

E (volts)

Mg2+ / Mg

-2.37

Zn2+ / Zn

-0.76

Cu2+ / Cu

+0.34

Ag+ / Ag

+0.80

Remember that each of these is comparing the position of the metal /


metal ion equilibrium with the equilibrium involving hydrogen.
Here are the five equilibria (including the hydrogen one):

If you compare these with the E values, you can see that the ones
http://www.chemguide.co.uk/physical/redoxeqia/introduction.html (17 of 19)30/12/2004 11:29:28

an introduction to redox equilibria and electrode potentials

whose positions of equilibrium lie furthest to the left have the most
negative E values. That is because they form ions more readily - and
leave more electrons behind on the metal, making it more negative.
Those which don't shed electrons as readily have positions of equilibrium
further to the right. Their E values get progressively more positive.

Note: Remember that, in each case, we are comparing the


position of equilibrium with the hydrogen equilibrium. For
example, we aren't saying that an equilibrium lies to the left in
absolute terms - just that it is further to the left than the
hydrogen equilibrium.

A final summary
E values give you a way of comparing the positions of equilibrium when
these elements lose electrons to form ions in solution.

The more negative the E value, the further the equilibrium lies to
the left - the more readily the element loses electrons and forms
ions.
The more positive (or less negative) the E value, the further the
equilibrium lies to the right - the less readily the element loses
electrons and forms ions.

Where would you like to go now?


To the next page on electrode potentials . . .
To the redox equilibria menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .
http://www.chemguide.co.uk/physical/redoxeqia/introduction.html (18 of 19)30/12/2004 11:29:28

an introduction to redox equilibria and electrode potentials

Jim Clark 2002

http://www.chemguide.co.uk/physical/redoxeqia/introduction.html (19 of 19)30/12/2004 11:29:28

the electrochemical series

THE ELECTROCHEMICAL SERIES

This page explains the origin of the electrochemical series, and shows
how it can be used to work out the ability of the various substances
included in it to act as oxidising or reducing agents.

Important: If you have come straight to this page via a


search engine (including the Google site search on the main
menu), you should be aware that this is the second page of a
linked series of pages about redox potentials. You will find it
much easier to understand if you first read the introduction to
redox equilibria before you go any further. A link at the
bottom of that page will bring you back here again.
It is also important that you understand about redox
reactions. Follow this link if you aren't confident about
oxidation and reduction in terms of electron transfer, and use
the BACK button on your browser to return to this page.

Building the electrochemical series


Arranging redox equilibria in order of their E values
The electrochemical series is built up by arranging various redox
equilibria in order of their standard electrode potentials (redox potentials).
The most negative E values are placed at the top of the electrochemical
series, and the most positive at the bottom.
For this introductory look at the electrochemical series we are going to
list the sort of metal / metal ion equilibria that we looked at on the
previous page (plus the hydrogen equilibrium) in order of their E values.
This will be extended to other systems on the next page.

http://www.chemguide.co.uk/physical/redoxeqia/ecs.html (1 of 5)30/12/2004 11:29:37

the electrochemical series

The electrochemical series


equilibrium

E (volts)
-3.03
-2.92
-2.87
-2.71
-2.37
-1.66
-0.76
-0.44
-0.13
0
+0.34
+0.80
+1.50

A note on the hydrogen value


Remember that each E value shows whether the position of the
equilibrium lies to the left or right of the hydrogen equilibrium.
That difference in the positions of equilibrium causes the number of
electrons which build up on the metal electrode and the platinum of the
hydrogen electrode to be different. That produces a potential difference
which is measured as a voltage.
Obviously if you connect one standard hydrogen electrode to another
http://www.chemguide.co.uk/physical/redoxeqia/ecs.html (2 of 5)30/12/2004 11:29:37

the electrochemical series

one, there will be no difference whatsoever between the positions of the


two equilibria. The number of electrons built up on each electrode will be
identical and so there will be a potential difference of zero between them.

Oxidation / reduction and the electrochemical series


Reminders about oxidation and reduction
Oxidation and reduction in terms of electron transfer
Remember that in terms of electrons:

Now apply this to one of the redox equilibria:

When solid magnesium forms its ions, it loses electrons. The magnesium
is being oxidised.
Taking another example . . .

When the copper(II) ions gain electrons to form copper, they are being
reduced.
Reducing agents and oxidising agents
A reducing agent reduces something else. That must mean that it gives
electrons to it.
Magnesium is good at giving away electrons to form its ions. Magnesium
must be a good reducing agent.

http://www.chemguide.co.uk/physical/redoxeqia/ecs.html (3 of 5)30/12/2004 11:29:37

the electrochemical series

An oxidising agent oxidises something else. That must mean that it


takes electrons from it.
Copper doesn't form its ions very readily, and its ions easily pick up
electrons from somewhere to revert to metallic copper. Copper(II) ions
must be good oxidising agents.
Summarizing this on the electrochemical series
Metals at the top of the series are good at giving away electrons. They
are good reducing agents. The reducing ability of the metal increases as
you go up the series.
Metal ions at the bottom of the series are good at picking up electrons.
They are good oxidising agents. The oxidising ability of the metal ions
increases as you go down the series.

Judging the oxidising or reducing ability from E values


The more negative the E value, the more the position of equilibrium lies
http://www.chemguide.co.uk/physical/redoxeqia/ecs.html (4 of 5)30/12/2004 11:29:37

the electrochemical series

to the left - the more readily the metal loses electrons. The more negative
the value, the stronger reducing agent the metal is.
The more positive the E value, the more the position of equilibrium lies
to the right - the less readily the metal loses electrons, and the more
readily its ions pick them up again. The more positive the value, the
stronger oxidising agent the metal ion is.

Where would you like to go now?


To the next page on electrode potentials . . .
To the redox equilibria menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/redoxeqia/ecs.html (5 of 5)30/12/2004 11:29:37

redox potentials for non-metallic systems

REDOX POTENTIALS FOR NON-METAL AND


OTHER SYSTEMS

This page explains how non-metals like chlorine can be included in the
electrochemical series, and how other oxidising and reducing agents can
have their standard electrode potentials (redox potentials) measured and
fitted into the series.

Important: If you have come straight to this page via a search


engine (including the Google site search on the main menu), you
should be aware that this is just one page in a linked series of
pages about redox potentials. You will find it much easier to
understand if you start from the beginning. Links at the bottom of
each page will bring you back here again.
It is also important that you understand about redox reactions.
Follow this link if you aren't confident about oxidation and
reduction in terms of electron transfer, and use the BACK button
on your browser to return to this page.

Measuring redox potentials for more complicated systems


Systems involving gases
The obvious example here is chlorine. Chlorine is well known as an oxidising
agent. Since the electrochemical series is about ranking substances
according to their oxidising or reducing ability, it makes sense to include
things like chlorine.
This time we are measuring the position of this equilibium relative to the
hydrogen equilibrium.

http://www.chemguide.co.uk/physical/redoxeqia/nonmetal.html (1 of 7)30/12/2004 11:29:47

redox potentials for non-metallic systems

Notice that the equilibrium is still written with the electrons on the left-hand
side of the equation. That's why the chlorine gas has to appear on the lefthand side rather than on the right (which is where the metals and hydrogen
appeared).
How can this equilibrium be connected into a circuit? The half-cell is built just
the same as a hydrogen electrode. Chlorine gas is bubbled over a platinum
electrode, which is immersed in a solution containing chloride ions with a
concentration of 1 mol dm-3.
The conventional way of writing the whole cell looks like this.

Notice the way that the chlorine half cell is written. The convention is that the
substance losing electrons is written closest to the electrode. In this case, the
chloride ions are losing electrons.

Note: Assuming that you know about oxidation states (oxidation


numbers), that is exactly the same as saying that the substance
with the lowest oxidation state is written closest to the electrode.

If you had the chlorine half cell on the left-hand side in a different situation,
then the convention still has to hold. The half cell would then be written:

What does the E value show in the Cl2 / Cl - case?

http://www.chemguide.co.uk/physical/redoxeqia/nonmetal.html (2 of 7)30/12/2004 11:29:47

redox potentials for non-metallic systems

The value is positive and moderately high as E values go. That means that
the position of the Cl2 / Cl - equilibrium lies more to the right than the
hydrogen equilibrium. Chlorine is much more likely to pick up electrons than
hydrogen ions are.

Chlorine is therefore quite good a removing electrons from other things. It is


a good oxidising agent.

Measuring redox potentials for other systems


The Fe2+ / Fe3+ system
Iron(II) ions are easily oxidised to iron(III) ions, and iron(III) ions are fairly
easily reduced to iron(II) ions. The equilibrium we are interested in this time
is:

To measure the redox potential of this, you would simply insert a platinum
electrode into a beaker containing a solution containing both iron(II) and iron
(III) ions (1 mol dm-3 with respect to both), and couple this to a hydrogen
electrode.
The cell diagram would look like this:

http://www.chemguide.co.uk/physical/redoxeqia/nonmetal.html (3 of 7)30/12/2004 11:29:47

redox potentials for non-metallic systems

Notice that the E value isn't as positive as the chlorine one. The position of
the iron(III) / iron(II) equilibrium isn't as far to the right as the chlorine
equilibrium. That means that Fe3+ ions don't pick up electrons as easily as
chlorine does. Chlorine is a stronger oxidising agent than Fe3+ ions.
Potassium dichromate(VI) as an oxidising agent
A commonly used oxidising agent, especially in organic chemistry, is
potassium dichromate(VI) solution acidified with dilute sulphuric acid. The
potassium ions are just spectator ions and aren't involved in the equilibrium
in any way.
The equilibrium is more complicated this time because it contains more
things:

The half cell would have a piece of platinum dipping into a solution containing
all the ions (dichromate(VI) ions, hydrogen ions and chromium(III) ions) all at
1 mol dm-3.
There is yet another convention when it comes to writing these more
complicated cell diagrams. Where there is more than one thing on either side
of the equilibrium, square brackets are written around them to keep them
tidy. The substances losing electrons are written next to the electrode, just as
before.

Including these new redox potentials in the electrochemical


series
These values can be slotted seamlessly into the electrochemical series that
so far has only included metals and hydrogen.
http://www.chemguide.co.uk/physical/redoxeqia/nonmetal.html (4 of 7)30/12/2004 11:29:47

redox potentials for non-metallic systems

An updated electrochemical series

equilibrium

E
(volts)
-3.03
-2.92
-2.87
-2.71
-2.37
-1.66
-0.76
-0.44
-0.13
0
+0.34
+0.77
+0.80
+1.33
+1.36
+1.50

By coincidence, all the new equilibria we've looked at have positive E


values. It so happens that most of the equilibria with negative E values that
you meet at this level are ones involving simple metal / metal ion

http://www.chemguide.co.uk/physical/redoxeqia/nonmetal.html (5 of 7)30/12/2004 11:29:47

redox potentials for non-metallic systems

combinations.
An update on oxidising agents
Remember:

The more positive the E value, the further the position of equilibrium
lies to the right.
That means that the more positive the E value, the more likely the
substances on the left-hand side of the equations are to pick up
electrons.
A substance which picks up electrons from something else is an
oxidising agent.
The more positive the E value, the stronger the substances on the lefthand side of the equation are as oxidising agents.

Of the new ones we've added to the electrochemical series:

Chlorine gas is the strongest oxidising agent (E = +1.36 v).


A solution containing dichromate(VI) ions in acid is almost as strong an
oxidising agent (E = +1.33 v).
Iron(III) ions are the weakest of the three new ones (E = +0.77 v).
None of these three are as strong an oxidising agent as Au3+ ions (E
= +1.50 v).

There will be more to say about this later in this series of pages.

Where would you like to go now?


To the next page on electrode potentials . . .
To the redox equilibria menu . . .
http://www.chemguide.co.uk/physical/redoxeqia/nonmetal.html (6 of 7)30/12/2004 11:29:47

redox potentials for non-metallic systems

To the Physical Chemistry menu . . .


To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/redoxeqia/nonmetal.html (7 of 7)30/12/2004 11:29:47

redox (electrode) potentials and test tube reactions

REDOX POTENTIALS AND SIMPLE TEST TUBE


REACTIONS

This page explains how standard electrode potentials (redox potentials)


relate to simple and familiar reactions that can be done in test tubes.

Important: If you have come straight to this page via a search


engine (including the Google site search on the main menu), you
should be aware that this is just one page in a linked series of
pages about redox potentials. You will find it much easier to
understand if you start from the beginning. Links at the bottom of
each page will bring you back here again.
It is also important that you understand about redox reactions particularly building ionic equations from electron-half-equations.
Follow this link if you aren't confident about this, and use the
BACK button on your browser to return to this page.

Combining a zinc with a copper half cell


So far in this series of pages, we have looked at combinations of a
hydrogen electrode with the half cell we have been interested in. However,
there isn't any reason why you can't couple any two half cells together.
This next bit looks at what happens if you combine a zinc half cell with a
copper half cell.
In the presence of a high resistance voltmeter

http://www.chemguide.co.uk/physical/redoxeqia/combinations.html (1 of 12)30/12/2004 11:30:02

redox (electrode) potentials and test tube reactions

The two equilibria which are set up in the half cells are:

The negative sign of the zinc E value shows that it releases electrons more
readily than hydrogen does. The positive sign of the copper E shows that it
releases electrons less readily than hydrogen.
That means that you can compare any two equilibria directly. For example,
in this case you can see that the zinc releases electrons more readily than
the copper does - the position of the zinc equilibrium lies further to the left
than the copper equilibrium.
Stripping everything else out of the diagram, and looking only at the build up
of electrons on the two pieces of metal:

http://www.chemguide.co.uk/physical/redoxeqia/combinations.html (2 of 12)30/12/2004 11:30:02

redox (electrode) potentials and test tube reactions

Obviously, the voltmeter will show that the zinc is the negative electrode,
and copper is the (relatively) positive one. It will register a voltage showing
the difference between them.

Note: It is very simple to calculate this voltage from the E


values. If you are interested in doing this, then you may like to
look at my chemistry calculations book. It is impossible for me to
include these calculations on this site without upsetting my
publishers!

Removing the voltmeter


The high resistance of the voltmeter is deliberately designed to stop any
current flow in the circuit. What happens if you remove the voltmeter and
replace it with a bit of wire?
Electrons will flow from where there are a lot of them (on the zinc) to where
there are fewer (on the copper). The movement of the electrons is an
electrical current.

http://www.chemguide.co.uk/physical/redoxeqia/combinations.html (3 of 12)30/12/2004 11:30:02

redox (electrode) potentials and test tube reactions

Important: In chemistry, you should always think in terms of


electron flow, and never in terms of current flow. The problem is
that conventionally in physics current flows in the opposite
direction to the electrons. That's totally silly! Avoid the problem.

The effect of this on the equilibria


These are just simple equilibria, and you can apply Le Chatelier's Principle
to them.

Note: If you are unsure about Le Chatelier's Principle, then you


must follow this link. You don't need to read the whole page just the beginning.
Use the BACK button on your browser to return quickly to this
page.

http://www.chemguide.co.uk/physical/redoxeqia/combinations.html (4 of 12)30/12/2004 11:30:02

redox (electrode) potentials and test tube reactions

Electrons are flowing away from the zinc equilibrium. According to Le


Chatelier's Principle, the position of equilibrium will move to replace the lost
electrons.
Electrons are being dumped onto the piece of copper in the copper
equilibrium. According to Le Chatelier's Principle, the position of equilibrium
will move to remove these extra electrons.

If electrons continue to flow, the positions of equilibrium keep on shifting.


The two equilibria essentially turn into two one-way reactions. The zinc
continues to ionise, and the copper(II) ions keep on picking up electrons.

Taking the apparatus as a whole, there is a chemical reaction going on in


which zinc is going into solution as zinc ions, and is giving electrons to
copper(II) ions to turn them into metallic copper.
Relating this to a test tube reaction

http://www.chemguide.co.uk/physical/redoxeqia/combinations.html (5 of 12)30/12/2004 11:30:02

redox (electrode) potentials and test tube reactions

This is exactly the same reaction that occurs when you drop a piece of zinc
into some copper(II) sulphate solution. The blue colour of the solution fades
as the copper(II) ions are converted into brown copper metal. The final
solution contains zinc sulphate. (The sulphate ions are spectator ions.)
You can add the two electron-half-equations above to give the overall ionic
equation for the reaction.

The only difference in this case is that the zinc gives the electrons directly to
the copper(II) ions rather than the electrons having to travel along a bit of
wire first.
The test tube reaction happens because of the relative tendency of the zinc
and copper to lose electrons to form ions. You can find out this relative
tendency by looking at the E values. That means that any redox reaction
could be discussed in a similar way.
The rest of the examples on this page illustrate this.

The reaction between copper and silver nitrate solution


The reaction in a test tube
If you hang a coil of copper wire in some colourless silver nitrate solution,
the copper gets covered in silver - partly as a grey fur, and partly as delicate
crystals. The solution turns blue.
Building the ionic equation from the two half-equations:

http://www.chemguide.co.uk/physical/redoxeqia/combinations.html (6 of 12)30/12/2004 11:30:02

redox (electrode) potentials and test tube reactions

Note: If you don't understand why the second equation is


multiplied by 2, you really must read about building ionic
equations from electron-half-equations. Use the BACK button on
your browser to return to this page.

Working from the redox potentials


How does this relate to the E values for copper and silver?

You can see that both of these E values are positive. Neither copper nor
silver produce ions and release electrons as easily as hydrogen does.
However, of the two, copper releases electrons more readily. In a cell, the
copper would have the greater build up of electrons, and be the negative
electrode. If the copper and zinc were connected by a bit of wire, electrons
would flow from the copper to the silver.
That, of course, will upset the two equilibria:

http://www.chemguide.co.uk/physical/redoxeqia/combinations.html (7 of 12)30/12/2004 11:30:02

redox (electrode) potentials and test tube reactions

Electrons will continue to flow and the two equilibria will again turn into oneway reactions to give the electron-half-equations we've just used to build the
ionic equation. Showing that again:

Note: People sometimes worry whether the fact that one of the
equations has to be multiplied by two affects the argument in any
way. It doesn't! It makes no difference whatsoever to any stage
of the explanation in terms of shifts in the positions of the two
equilibria.

http://www.chemguide.co.uk/physical/redoxeqia/combinations.html (8 of 12)30/12/2004 11:30:02

redox (electrode) potentials and test tube reactions

A useful "rule of thumb"


Whenever you link two of these equilibria together (either via a bit of wire, or
by allowing one of the substances to give electrons directly to another one
in a test tube):

The equilibrium with the more negative (or less positive) E value will
move to the left.
The equilibrium with the more positive (or less negative) E value will
move to the right.

It sounds simple, but that is the key to using E values in most of the
situations you will come across.

The reaction between magnesium and dilute sulphuric acid


The reaction in a test tube
Magnesium reacts with dilute sulphuric acid to give hydrogen and a
colourless solution containing magnesium sulphate.
Is this what you would expect from the E values?
Working from the redox potentials
Putting all the argument onto one diagram:

http://www.chemguide.co.uk/physical/redoxeqia/combinations.html (9 of 12)30/12/2004 11:30:02

redox (electrode) potentials and test tube reactions

The two equilibria become one-way reactions which you can use to build the
ionic equation:

Using potassium dichromate(VI) as an oxidising agent


The reaction in a test tube
Potassium dichromate(VI) acidified with dilute sulphuric acid oxidises iron(II)
ions to iron(III) ions. The orange solution containing the dichromate(VI) ions
turns green as chromium(III) ions are formed.
How does this relate to the E values?

http://www.chemguide.co.uk/physical/redoxeqia/combinations.html (10 of 12)30/12/2004 11:30:02

redox (electrode) potentials and test tube reactions

Warning! Those of you who have come across other ways of


working with E values may have learnt rules which force you to
write the equilibria down in a particular order. I have deliberately
written this example down so that it disobeys these rules!
This is to show that these rules are completely unnecessary. All
you have to do is remember that the more negative (less
positive) equilibrium will shift to the left, and the other one to the
right.

Building the ionic equation then works like this:

http://www.chemguide.co.uk/physical/redoxeqia/combinations.html (11 of 12)30/12/2004 11:30:02

redox (electrode) potentials and test tube reactions

Coming up next
The final page in the sequence simply expands on this one, and looks at
how you can use E values to predict whether or not redox reactions are
feasible. It would be a good idea to be fairly confident about the present
page before you went on to the final one.

Where would you like to go now?


To the final page on electrode potentials . . .
To the redox equilibria menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/redoxeqia/combinations.html (12 of 12)30/12/2004 11:30:02

making predictions using redox (electrode) potentials

MAKING PREDICTIONS USING REDOX


POTENTIALS (ELECTRODE POTENTIALS)

This page explains how to use redox potentials (electrode potentials) to


predict the feasibility of redox reactions. It also looks at how you go about
choosing a suitable oxidising agent or reducing agent for a particular
reaction.

Important: This is the final page in a sequence of five pages


about redox potentials. You will find it much easier to
understand if you start from the beginning. Links at the bottom
of each page will bring you back here again.
Don't try to short-cut this. Redox potentials are absolutely
simple to work with if you understand what they are about. If
you don't, the whole topic can be a complete nightmare!

Predicting the feasibility of a possible redox reaction


A reminder of what you need to know
Standard electrode potentials (redox potentials) are one way of measuring
how easily a substance loses electrons. In particular, they give a measure
of relative positions of equilibrium in reactions such as:

The more negative the E value, the further the position of equilibrium lies
to the left. Remember that this is always relative to the hydrogen
equilibrium - and not in absolute terms.
The negative sign of the zinc E value shows that it releases electrons
http://www.chemguide.co.uk/physical/redoxeqia/predict.html (1 of 14)30/12/2004 11:30:19

making predictions using redox (electrode) potentials

more readily than hydrogen does. The positive sign of the copper E value
shows that it releases electrons less readily than hydrogen.
Whenever you link two of these equilibria together (either via a bit of wire,
or by allowing one of the substances to give electrons directly to another
one in a test tube) electrons flow from one equilibrium to the other. That
upsets the equilibria, and Le Chatelier's Principle applies. The positions of
equilibrium move - and keep on moving if the electrons continue to be
transferred.
The two equilibria essentially turn into two one-way reactions:

The equilibrium with the more negative (or less positive) E value
will move to the left.
The equilibrium with the more positive (or less negative) E value
will move to the right.

Note: If you aren't confident about this, please go back and


start from the beginning of this sequence of pages. All these
ideas are explored in a gentle and logical way. Links at the
bottom of each page will bring you back here again.

Will magnesium react with dilute sulphuric acid?


Of course it does! I'm choosing this as an introductory example because
everybody will know the right answer before we start. We have also
explored this from a slightly different point of view on the previous page in
this sequence.
The E values are:

You are starting with magnesium metal and hydrogen ions in the acid. The
sulphate ions are spectator ions and play no part in the reaction.

http://www.chemguide.co.uk/physical/redoxeqia/predict.html (2 of 14)30/12/2004 11:30:19

making predictions using redox (electrode) potentials

Think of it like this. There is no need to write anything down unless you
want to. With a small amount of practice, all you need to do is just look at
the numbers.

Is there anything to stop the sort of movements we have suggested? No!


The magnesium can freely turn into magnesium ions and give electrons to
the hydrogen ions producing hydrogen gas. The reaction is feasible.

Now for a reaction which turns out not to be feasible . . .


Will copper react with dilute sulphuric acid?
You know that the answer is that it won't. How do the E values predict
this?

http://www.chemguide.co.uk/physical/redoxeqia/predict.html (3 of 14)30/12/2004 11:30:19

making predictions using redox (electrode) potentials

Doing the same sort of thinking as before:

The diagram shows the way that the E values are telling us that the
equilibria will tend to move. Is this possible? No!
If we start from copper metal, the copper equilibrium is already completely
to the right. If it were to react at all, the equilibrium will have to move to the
left - directly opposite to what the E values are saying.
Similarly, if we start from hydrogen ions (from the dilute acid), the
hydrogen equilibrium is already as far to the left as possible. For it to react,
it would have to move to the right - against what the E values demand.
There is no possibility of a reaction.

In the next couple of examples, decide for yourself whether or not the
reaction is feasible before you read the text.
Will oxygen oxidise iron(II) hydroxide to iron(III) hydroxide under
alkaline conditions?

http://www.chemguide.co.uk/physical/redoxeqia/predict.html (4 of 14)30/12/2004 11:30:19

making predictions using redox (electrode) potentials

The E values are:

Think about this before you read on. Remember that the equilibrium with
the more negative E value will tend to move to the left. The other one
tends to move to the right. Is that possible?

Yes, it is possible. Given what we are starting with, both of these equilibria
can move in the directions required by the E values. The reaction is
feasible.

Will chlorine oxidise manganese(II) ions to manganate(VII) ions?


The E values are:

http://www.chemguide.co.uk/physical/redoxeqia/predict.html (5 of 14)30/12/2004 11:30:19

making predictions using redox (electrode) potentials

Again, think about this before you read on.

Given what you are starting from, these equilibrium shifts are impossible.
The manganese equilibrium has the more positive E value and so will
tend to move to the right. However, because we are starting from
manganese(II) ions, it is already as far to the right as possible. In order to
get any reaction, the equilibrium would have to move to the left. That is
against what the E values are saying.
This reaction isn't feasible.

Will dilute nitric acid react with copper?


This is going to be more complicated because there are two different ways
in which dilute nitric acid might possibly react with copper. The copper
might react with the hydrogen ions or with the nitrate ions. Nitric acid
reactions are always more complex than the simpler acids like sulphuric or
hydrochloric acid because of this problem.
Here are the E values:

http://www.chemguide.co.uk/physical/redoxeqia/predict.html (6 of 14)30/12/2004 11:30:19

making predictions using redox (electrode) potentials

We have already discussed the possibility of copper reacting with


hydrogen ions further up this page. Go back and look at it again if you
need to, but the argument (briefy) goes like this:
The copper equilibrium has a more positive E value than the hydrogen
one. That means that the copper equilibium will tend to move to the right
and the hydrogen one to the left.
However, if we start from copper and hydrogen ions, the equilibria are
already as far that way as possible. Any reaction would need them to
move in the opposite direction to what the E values want. The reaction
isn't feasible.
What about a reaction between the copper and the nitrate ions?
This is feasible. The nitrate ion equilibrium has the more positive E value
and will move to the right. The copper E value is less positive. That
equilibrium will move to the left. The movements that the E values
suggest are possible, and so a reaction is feasible.
Copper(II) ions are produced together with nitrogen monoxide gas.

Warning! There are all sorts of ways of dealing with this


feasibility question - all of them, I believe, more complicated
than this. Some methods actually force you to do some
(simple) calculations. Unfortunately, some examiners ask
questions which make you use their own inefficient methods of
working out whether a reaction is feasible, and you need to
know if this is going to be a problem.
UK A level students should refer to their syllabuses and past
papers. Follow this link if you haven't got the necessary
information.

http://www.chemguide.co.uk/physical/redoxeqia/predict.html (7 of 14)30/12/2004 11:30:19

making predictions using redox (electrode) potentials

You will find the calculation approach covered in my chemistry


calculations book, together with more examples using the
method developed on these pages. However, I find the
calculation approach so pointless that I refuse to include it on
this site! Why do a calculation if you can just look at two
numbers and decide in seconds whether or not a reaction is
feasible?

Two examples where the E values seem to give the wrong answer
It sometimes happens that E values suggest that a reaction ought to
happen, but it doesn't. Occasionally, a reaction happens although the E
values seem to be the wrong way around. These next two examples
explain how that can happen. By coincidence, both involve the dichromate
(VI) ion in potassium dichromate(VI).
Will acidified potassium dichromate(VI) oxidise water?
The E values are:

The relative sizes of the E values show that the reaction is feasible:

http://www.chemguide.co.uk/physical/redoxeqia/predict.html (8 of 14)30/12/2004 11:30:19

making predictions using redox (electrode) potentials

However, in the test tube nothing happens however long you wait. An
acidified solution of potassium dichromate(VI) doesn't react with the water
that it is dissolved in. So what is wrong with the argument?
In fact, there is nothing wrong with the argument. The problem is that all
the E values show is that a reaction is possible. They don't tell you that it
will actually happen. There may be very large activation barriers to the
reaction which prevent it from taking place.
Always treat what E values tell you with some caution. All they tell you is
whether a reaction is feasible - they tell you nothing about how fast the
reaction will happen.
Will acidified potassium dichromate(VI) oxidise chloride ions to
chlorine?
The E values are:

Because the chlorine E value is slightly greater than the dichromate(VI)


one, there shouldn't be any reaction. For a reaction to occur, the equilibria
would have to move in the wrong directions.
http://www.chemguide.co.uk/physical/redoxeqia/predict.html (9 of 14)30/12/2004 11:30:19

making predictions using redox (electrode) potentials

Unfortunately, in the test tube, potassium dichromate(VI) solution does


oxidise concentrated hydrochloric acid to chlorine. The hydrochloric acid
serves as the source of the hydrogen ions in the dichromate(VI)
equilibrium and of the chloride ions.
The problem here is that E values only apply under standard conditions. If
you change the conditions you will change the position of an equilibrium and that will change its E value. (Notice that you can't call it an E value
any more, because the conditions are no longer standard.)
The standard condition for concentration is 1 mol dm-3. But concentrated
hydrochloric acid is approximately 10 mol dm-3. The concentrations of the
hydrogen ions and chloride ions are far in excess of standard.
What effect does that have on the two positions of equilibrium?

http://www.chemguide.co.uk/physical/redoxeqia/predict.html (10 of 14)30/12/2004 11:30:19

making predictions using redox (electrode) potentials

Because the E values are so similar, you don't have to change them very
much to make the dichromate(VI) one the more positive. As soon as that
happens, it will react with the chloride ions to produce chlorine.
In most cases, there is enough difference between E values that you can
ignore the fact that you aren't doing a reaction under strictly standard
conditions. But sometimes it does make a difference. Be careful!

Selecting an oxidising or reducing agent for a reaction


There is nothing remotely new in this. It is just a slight variation on what
we've just been looking at.
Choosing an oxidising agent
Remember:

Oxidation is loss of electrons.


An oxidising agent oxidises something by removing electrons from
it. That means that the oxidising agent gains electrons.

http://www.chemguide.co.uk/physical/redoxeqia/predict.html (11 of 14)30/12/2004 11:30:19

making predictions using redox (electrode) potentials

It is easier to explain this with a specific example. What could you use to
oxidise iron(II) ions to iron(III) ions? The E value for this reaction is:

To change iron(II) ions into iron(III) ions, you need to persuade this
equilibrium to move to the left. That means that when you couple it to a
second equilibrium, this iron E value must be the more negative (less
positive one).
You could use anything which has a more positive E value. For example,
you could use any of the these which we have already looked at over the
last page or two:
Dilute nitric acid:

Acidified potassium dichromate(VI):

Chlorine:

Acidified potassium manganate(VII):

Choosing a reducing agent


Remember:

Reduction is gain of electrons.


A reducing agent reduces something by giving electrons to it. That
means that the reducing agent loses electrons.

http://www.chemguide.co.uk/physical/redoxeqia/predict.html (12 of 14)30/12/2004 11:30:19

making predictions using redox (electrode) potentials

You have to be a little bit more careful this time, because the substance
losing electrons is found on the right-hand side of one of these redox
equilibria. Again, a specific example makes it clearer.
For example, what could you use to reduce chromium(III) ions to chromium
(II) ions? The E value is:

You need this equilibrium to move to the right. That means that when you
couple it with a second equilibrium, this chromium E value must be the
most positive (least negative).
In principle, you could choose anything with a more negative E value - for
example, zinc:

You would have to remember to start from metallic zinc, and not zinc ions.
You need this second equilibrium to be able to move to the left to provide
the electrons. If you started with zinc ions, it would already be on the left and would have no electrons to give away. Nothing could possibly happen
if you mixed chromium(III) ions and zinc ions.
That is fairly obvious in this simple case. If you were dealing with a more
complicated equilibrium, you would have to be careful to think it through
properly.

Where would you like to go now?


To the redox equilibria menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

http://www.chemguide.co.uk/physical/redoxeqia/predict.html (13 of 14)30/12/2004 11:30:19

making predictions using redox (electrode) potentials

Jim Clark 2002

http://www.chemguide.co.uk/physical/redoxeqia/predict.html (14 of 14)30/12/2004 11:30:19

Acid-base equilibria menu

Understanding Chemistry

ACID-BASE EQUILIBRIA MENU

Theories of acids and bases . . .


Describes the Arrhenius, Bronsted-Lowry, and Lewis theories of
acids and bases, and explains the relationship between them.
Includes the meaning of the term conjugate as applied to acidbase pairs.
Strong and weak acids . . .
Explains the terms strong and weak as applied to acids. Defines
pH, Ka and pKa.
The ionic product for water . . .
Explains Kw and pKw, and how these terms define the pH of pure
water at various temperatures.
Strong and weak bases . . .
Explains the terms strong and weak as applied to bases. Defines
Kb and pKb for a weak base.
pH curves . . .
Describes the way that pH changes during various acid-base
titrations.
acid-base indicators . . .

http://www.chemguide.co.uk/physical/acideqiamenu.html (1 of 2)30/12/2004 11:30:20

Acid-base equilibria menu

Explains how simple indicators work, and what determines the


choice of indicator for a particular acid-base titration. This page
assumes that you know about pH curves.
Buffer solutions . . .
Explains what a buffer solution is and how simple buffer solutions
work.

Go to physical chemistry menu . . .


Go to Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/acideqiamenu.html (2 of 2)30/12/2004 11:30:20

the ionic product for water: Kw

THE IONIC PRODUCT FOR WATER, Kw

This page explains what is meant by the ionic product for water. It looks
at how the ionic product varies with temperature, and how that
determines the pH of pure water at different temperatures.

Kw and pKw
The important equilibrium in water
Water molecules can function as both acids and bases. One water
molecule (acting as a base) can accept a hydrogen ion from a second
one (acting as an acid). This will be happening anywhere there is even a
trace of water - it doesn't have to be pure.
A hydroxonium ion and a hydroxide ion are formed.

However, the hydroxonium ion is a very strong acid, and the hydroxide
ion is a very strong base. As fast as they are formed, they react to
poduce water again.
The net effect is that an equilibrium is set up.
http://www.chemguide.co.uk/physical/acidbaseeqia/kw.html (1 of 7)30/12/2004 11:30:25

the ionic product for water: Kw

At any one time, there are incredibly small numbers of hydroxonium ions
and hydroxide ions present. Further down this page, we shall calculate
the concentration of hydroxonium ions present in pure water. It turns out
to be 1.00 x 10-7 mol dm-3 at room temperature.
You may well find this equilibrium written in a simplified form:

This is OK provided you remember that H+(aq) actually refers to a


hydroxonium ion.

Defining the ionic product for water, Kw


Kw is essentially just an equilibrium constant for the reactions shown.
You may meet it in two forms:
Based on the fully written equilibrium . . .

. . . or on the simplified equilibrium:

You may find them written with or without the state symbols. Whatever
version you come across, they all mean exactly the same thing!
You may wonder why the water isn't written on the bottom of these
equilibrium constant expressions. So little of the water is ionised at any
one time, that its concentration remains virtually unchanged - a constant.
Kw is defined to avoid making the expression unnecessarily complicated
by including another constant in it.
http://www.chemguide.co.uk/physical/acidbaseeqia/kw.html (2 of 7)30/12/2004 11:30:25

the ionic product for water: Kw

The value of Kw
Like any other equilibrium constant, the value of Kw varies with
temperature. Its value is usually taken to be 1.00 x 10-14 mol2 dm-6 at
room temperature. In fact, this is its value at a bit less than 25C.

The units of Kw: Kw is found by multiplying two


concentration terms together. Each of these has the units of
mol dm-3.
Multiplying mol dm-3 x mol dm-3 gives you the units above.

pKw
The relationship between Kw and pKw is exactly the same as that
between Ka and pKa, or [H+] and pH.

The Kw value of 1.00 x 10-14 mol2 dm-6 at room temperature gives you a
pKw value of 14. Try it on your calculator! Notice that pKw doesn't have
any units.

The pH of pure water


Why does pure water have a pH of 7?
That question is actually misleading! In fact, pure water only has a pH of
7 at a particular temperature - the temperature at which the Kw value is
1.00 x 10-14 mol2 dm-6.

http://www.chemguide.co.uk/physical/acidbaseeqia/kw.html (3 of 7)30/12/2004 11:30:25

the ionic product for water: Kw

This is how it comes about:


To find the pH you need first to find the hydrogen ion concentration (or
hydroxonium ion concentration - it's the same thing). Then you convert it
to pH.
In pure water at room temperature the Kw value tells you that:
[H+] [OH-] = 1.00 x 10-14
But in pure water, the hydrogen ion (hydroxonium ion) concentration
must be equal to the hydroxide ion concentration. For every hydrogen ion
formed, there is a hydroxide ion formed as well.
That means that you can replace the [OH-] term in the Kw expression by
another [H+].
[H+]2 = 1.00 x 10-14
Taking the square root of each side gives:
[H+] = 1.00 x 10-7 mol dm-3
Converting that into pH:
pH = - log10 [H+]
pH = 7
That's where the familiar value of 7 comes from.

The variation of the pH of pure water with temperature


The formation of hydrogen ions (hydroxonium ions) and hydroxide ions
from water is an endothermic process. Using the simpler version of the
equilibrium:

http://www.chemguide.co.uk/physical/acidbaseeqia/kw.html (4 of 7)30/12/2004 11:30:25

the ionic product for water: Kw

The forward reaction absorbs heat.


According to Le Chatelier's Principle, if you make a change to the
conditions of a reaction in dynamic equilibrium, the position of equilibrium
moves to counter the change you have made.

Note: If you don't understand Le Chatelier's Principle, you


should follow this link before you go on. Make sure that you
understand the effect of temperature on position of
equilibrium.
Use the BACK button on your browser when you are ready to
return to this page.

According to Le Chatelier, if you increase the temperature of the water,


the equilibrium will move to lower the temperature again. It will do that by
absorbing the extra heat.
That means that the forward reaction will be favoured, and more
hydrogen ions and hydroxide ions will be formed. The effect of that is to
increase the value of Kw as temperature increases.
The table below shows the effect of temperature on Kw. For each value
of Kw, a new pH has been calculated using the same method as above. It
might be useful if you were to check these pH values yourself.

T (C)

Kw (mol2 dm-6)

pH

0.114 x 10-14

7.47

10

0.293 x 10-14

7.27

http://www.chemguide.co.uk/physical/acidbaseeqia/kw.html (5 of 7)30/12/2004 11:30:25

the ionic product for water: Kw

20

0.681 x 10-14

7.08

25

1.008 x 10-14

7.00

30

1.471 x 10-14

6.92

40

2.916 x 10-14

6.77

50

5.476 x 10-14

6.63

100

51.3 x 10-14

6.14

You can see that the pH of pure water falls as the temperature increases.
A word of warning!
If the pH falls as temperature increases, does this mean that water
becomes more acidic at higher temperatures? NO!
A solution is acidic if there is an excess of hydrogen ions over hydroxide
ions. In the case of pure water, there are always the same number of
hydrogen ions and hydroxide ions. That means that the water remains
neutral - even if its pH changes.
The problem is that we are all so familiar with 7 being the pH of pure
water, that anything else feels really strange. Remember that you
calculate the neutral value of pH from Kw. If that changes, then the
neutral value for pH changes as well.
At 100C, the pH of pure water is 6.14. That is the neutral point on the
pH scale at this higher temperature. A solution with a pH of 7 at this
temperature is slightly alkaline because its pH is a bit higher than the
neutral value of 6.14.
Similarly, you can argue that a solution with a pH of 7 at 0C is slightly
acidic, because its pH is a bit lower than the neutral value of 7.47 at this
temperature.

http://www.chemguide.co.uk/physical/acidbaseeqia/kw.html (6 of 7)30/12/2004 11:30:25

the ionic product for water: Kw

Where would you like to go now?


To the acid-base equilibria menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/acidbaseeqia/kw.html (7 of 7)30/12/2004 11:30:25

strong and weak bases

STRONG AND WEAK BASES

This page explains the terms strong and weak as applied to bases. As a
part of this it defines and explains Kb and pKb.
We are going to use the Bronsted-Lowry definition of a base as a
substance which accepts hydrogen ions (protons).

Note: If you don't know what the Bronsted-Lowry theory is,


you should read about theories of acids and bases on
another page in this section. You don't need to spend time
reading about Lewis acids and bases for the purposes of this
present page.
Use the BACK button on your browser when you are ready to
return to this page.

The usual way of comparing the strengths of bases is to see how readily
they produce hydroxide ions in solution. This may be because they
already contain hydroxide ions, or because they take hydrogen ions from
water molecules to produce hydroxide ions.

Strong bases
Explaining the term "strong base"
A strong base is something like sodium hydroxide or potassium
hydroxide which is fully ionic. You can think of the compound as being
100% split up into metal ions and hydroxide ions in solution.
Each mole of sodium hydroxide dissolves to give a mole of hydroxide
ions in solution.
http://www.chemguide.co.uk/physical/acidbaseeqia/bases.html (1 of 7)30/12/2004 11:30:32

strong and weak bases

Some strong bases like calcium hydroxide aren't very soluble in water.
That doesn't matter - what does dissolve is still 100% ionised into
calcium ions and hydroxide ions. Calcium hydroxide still counts as a
strong base because of that 100% ionisation.
Working out the pH of a strong base
Remember that:

Since pH is a measure of hydrogen ion concentration, how can a solution


which contains hydroxide ions have a pH? To understand this, you need
to know about the ionic product for water.

Important: If you don't understand about the ionic product


for water you must follow this link before you go on.
Use the BACK button on your browser when you are ready to
return to this page.

http://www.chemguide.co.uk/physical/acidbaseeqia/bases.html (2 of 7)30/12/2004 11:30:32

strong and weak bases

Wherever there is water, an equilibrium is set up. Using the simplified


version of this equilibrium:

In the presence of extra hydroxide ions from, say, sodium hydroxide, the
equilibrium is still there, but the position of equilibrium has been shifted
well to the left according to Le Chatelier's Principle.

Note: If you don't understand Le Chatelier's Principle, you


should follow this link before you go on. Make sure that you
understand the effect of concentration on position of
equilibrium.
Use the BACK button on your browser when you are ready to
return to this page.

There will be far fewer hydrogen ions than there are in pure water, but
there will still be hydrogen ions present. The pH is a measure of the
concentration of these.
An outline of the method of working out the pH of a strong base

Work out the concentration of the hydroxide ions.

Use Kw to work out the hydrogen ion concentration.

Convert the hydrogen ion concentration to a pH.

An example
To find the pH of 0.500 mol dm-3 sodium hydroxide solution:
Because the sodium hydroxide is fully ionic, each mole of it gives that
same number of moles of hydroxide ions in solution.

http://www.chemguide.co.uk/physical/acidbaseeqia/bases.html (3 of 7)30/12/2004 11:30:32

strong and weak bases

[OH-] = 0.500 mol dm-3

Note: You would have to be careful here if you had a base


like calcium hydroxide, Ca(OH)2. Each mole of calcium
hydroxide would produce twice as many hydroxide ions in
solution.

Now you use the value of Kw at the temperature of your solution. You
normally take this as 1.00 x 10-14 mol2 dm-6.
[H+] [OH-] = 1.00 x 10-14
This is true whether the water is pure or not. In this case we have a value
for the hydroxide ion concentration. Substituting that gives:
[H+] x 0.500 = 1.00 x 10-14
If you solve that for [H+], and then convert it into pH, you get a pH of 13.7.

Note: If you want more examples to look at and to try


yourself (with fully worked solutions given), you may be
interested in my chemistry calculations book. This also
includes the reverse problem which is slightly more confusing
- working out the concentration of a strong base from its pH.

http://www.chemguide.co.uk/physical/acidbaseeqia/bases.html (4 of 7)30/12/2004 11:30:32

strong and weak bases

Weak bases
Explaining the term "weak base"
Ammonia is a typical weak base. Ammonia itself obviously doesn't
contain hydroxide ions, but it reacts with water to produce ammonium
ions and hydroxide ions.

However, the reaction is reversible, and at any one time about 99% of
the ammonia is still present as ammonia molecules. Only about 1% has
actually produced hydroxide ions.
A weak base is one which doesn't convert fully into hydroxide ions in
solution.

Important: What follows isn't required by any of the current


UK A' level syllabuses.

Comparing the strengths of weak bases in solution: Kb


When a weak base reacts with water, the position of equilibrium varies
from base to base. The further to the left it is, the weaker the base.

You can get a measure of the position of an equilibrium by writing an


equilibrium constant for the reaction. The lower the value for the
constant, the more the equilibrium lies to the left.
http://www.chemguide.co.uk/physical/acidbaseeqia/bases.html (5 of 7)30/12/2004 11:30:32

strong and weak bases

In this case the equilibrium constant is called Kb. This is defined as:

Note: If you want to know why the water has been omitted
from the bottom of this expression, you will find it explained
on the page about strong and weak acids under the
corresponding constant, Ka.
Use the BACK button on your browser when you are ready to
return to this page.

pKb
The relationship between Kb and pKb is exactly the same as all the other
"p" terms in this topic:

The table shows some values for Kb and pKb for some weak bases.
Kb (mol dm-3)

pKb

C6H5NH2

4.17 x 10-10

9.38

NH3

1.78 x 10-5

4.75

CH3NH2

4.37 x 10-4

3.36

CH3CH2NH2

5.37 x 10-4

3.27

base

http://www.chemguide.co.uk/physical/acidbaseeqia/bases.html (6 of 7)30/12/2004 11:30:32

strong and weak bases

As you go down the table, the value of Kb is increasing. That means that
the bases are getting stronger.
As Kb gets bigger, pKb gets smaller. The lower the value of pKb, the
stronger the base.
This is exactly in line with the corresponding term for acids, pKa - the
smaller the value, the stronger the acid.

Note: If you are interested in exploring organic bases


further, you will find them explained elsewhere on the site.

Where would you like to go now?


To the acid-base equilibria menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/acidbaseeqia/bases.html (7 of 7)30/12/2004 11:30:32

pH curves (titration curves)

pH (TITRATION) CURVES

This page describes how pH changes during various acid-base titrations.

The equivalence point of a titration


Sorting out some confusing terms
When you carry out a simple acid-base titration, you use an indicator to
tell you when you have the acid and alkali mixed in exactly the right
proportions to "neutralise" each other. When the indicator changes
colour, this is often described as the end point of the titration.
In an ideal world, the colour change would happen when you mix the two
solutions together in exactly equation proportions. That particular mixture
is known as the equivalence point.
For example, if you were titrating sodium hydroxide solution with
hydrochloric acid, both with a concentration of 1 mol dm-3, 25 cm3 of
sodium hydroxide solution would need exactly the same volume of the
acid - because they react 1 : 1 according to the equation.

In this particular instance, this would also be the neutral point of the
titration, because sodium chloride solution has a pH of 7.
But that isn't necessarily true of all the salts you might get formed.
For example, if you titrate ammonia solution with hydrochloric acid, you
would get ammonium chloride formed. The ammonium ion is slightly
acidic, and so pure ammonium chloride has a slightly acidic pH.
That means that at the equivalence point (where you had mixed the
solutions in the correct proportions according to the equation), the
solution wouldn't actually be neutral. To use the term "neutral point" in
http://www.chemguide.co.uk/physical/acidbaseeqia/phcurves.html (1 of 13)30/12/2004 11:30:50

pH curves (titration curves)

this context would be misleading.


Similarly, if you titrate sodium hydroxide solution with ethanoic acid, at
the equivalence point the pure sodium ethanoate formed has a slightly
alkaline pH because the ethanoate ion is slightly basic.
To summarise:

The term "neutral point" is best avoided.


The term "equivalence point" means that the solutions have been
mixed in exactly the right proportions according to the equation.
The term "end point" is where the indicator changes colour. As you
will see on the page about indicators, that isn't necessarily exactly
the same as the equivalence point.

Note: You can find out about indicators by following this link
(also available from the acid-base equilibria menu). You
should read the present page first though.

Simple pH curves
All the following titration curves are based on both acid and alkali having
a concentration of 1 mol dm-3. In each case, you start with 25 cm3 of one
of the solutions in the flask, and the other one in a burette.
Although you normally run the acid from a burette into the alkali in a
flask, you may need to know about the titration curve for adding it the
other way around as well. Alternative versions of the curves have been
described in most cases.
Titration curves for strong acid v strong base
We'll take hydrochloric acid and sodium hydroxide as typical of a strong
acid and a strong base.
http://www.chemguide.co.uk/physical/acidbaseeqia/phcurves.html (2 of 13)30/12/2004 11:30:50

pH curves (titration curves)

Running acid into the alkali

You can see that the pH only falls a very small amount until quite near
the equivalence point. Then there is a really steep plunge. If you
calculate the values, the pH falls all the way from 11.3 when you have
added 24.9 cm3 to 2.7 when you have added 25.1 cm3.

Note: If you need to know how to calculate pH changes


during a titration, you may be interested in my chemistry
calculations book.

http://www.chemguide.co.uk/physical/acidbaseeqia/phcurves.html (3 of 13)30/12/2004 11:30:50

pH curves (titration curves)

Running alkali into the acid


This is very similar to the previous curve except, of course, that the pH
starts off low and increases as you add more sodium hydroxide solution.

Again, the pH doesn't change very much until you get close to the
equivalence point. Then it surges upwards very steeply.

Titration curves for strong acid v weak base


This time we are going to use hydrochloric acid as the strong acid and
ammonia solution as the weak base.

Running acid into the alkali

http://www.chemguide.co.uk/physical/acidbaseeqia/phcurves.html (4 of 13)30/12/2004 11:30:50

pH curves (titration curves)

Because you have got a weak base, the beginning of the curve is
obviously going to be different. However, once you have got an excess of
acid, the curve is essentially the same as before.
At the very beginning of the curve, the pH starts by falling quite quickly
as the acid is added, but the curve very soon gets less steep. This is
because a buffer solution is being set up - composed of the excess
ammonia and the ammonium chloride being formed.

Note: You can find out more about buffer solutions by


following this link. However, this is a very minor point in the
present context, and you would probably do better to read the
whole of the current page before you follow this up.

http://www.chemguide.co.uk/physical/acidbaseeqia/phcurves.html (5 of 13)30/12/2004 11:30:50

pH curves (titration curves)

Notice that the equivalence point is now somewhat acidic ( a bit less than
pH 5), because pure ammonium chloride isn't neutral. However, the
equivalence point still falls on the steepest bit of the curve. That will turn
out to be important in choosing a suitable indicator for the titration.
Running alkali into the acid
At the beginning of this titration, you have an excess of hydrochloric acid.
The shape of the curve will be the same as when you had an excess of
acid at the start of a titration running sodium hydroxide solution into the
acid.
It is only after the equivalence point that things become different.
A buffer solution is formed containing excess ammonia and ammonium
chloride. This resists any large increase in pH - not that you would expect
a very large increase anyway, because ammonia is only a weak base.

Titration curves for weak acid v strong base


We'll take ethanoic acid and sodium hydroxide as typical of a weak acid
and a strong base.

http://www.chemguide.co.uk/physical/acidbaseeqia/phcurves.html (6 of 13)30/12/2004 11:30:50

pH curves (titration curves)

Running acid into the alkali


For the first part of the graph, you have an excess of sodium hydroxide.
The curve will be exactly the same as when you add hydrochloric acid to
sodium hydroxide. Once the acid is in excess, there will be a difference.

Past the equivalence point you have a buffer solution containing sodium
ethanoate and ethanoic acid. This resists any large fall in pH.
Running alkali into the acid

http://www.chemguide.co.uk/physical/acidbaseeqia/phcurves.html (7 of 13)30/12/2004 11:30:50

pH curves (titration curves)

The start of the graph shows a relatively rapid rise in pH but this slows
down as a buffer solution containing ethanoic acid and sodium ethanoate
is produced. Beyond the equivalence point (when the sodium hydroxide
is in excess) the curve is just the same as that end of the HCl - NaOH
graph.

Titration curves for weak acid v weak base


The common example of this would be ethanoic acid and ammonia.

It so happens that these two are both about equally weak - in that case,
the equivalence point is approximately pH 7.
Running acid into the alkali
This is really just a combination of graphs you have already seen. Up to
the equivalence point it is similar to the ammonia - HCl case. After the
equivalence point it is like the end of the ethanoic acid - NaOH curve.

http://www.chemguide.co.uk/physical/acidbaseeqia/phcurves.html (8 of 13)30/12/2004 11:30:50

pH curves (titration curves)

Notice that there isn't any steep bit on this graph. Instead, there is just
what is known as a "point of inflexion". That lack of a steep bit means
that it is difficult to do a titration of a weak acid against a weak base.

Note: Because you almost never do titrations with this


combination, there is no real point in giving the graph where
they are added the other way round. It isn't difficult to work
out what it might look like if you are interested - take the
beginning of the sodium hydroxide added to ethanoic acid
curve, and the end of the ammonia added to hydrochloric
acid one.
The reason that it is difficult to do these titrations is discussed
on the page about indicators.

http://www.chemguide.co.uk/physical/acidbaseeqia/phcurves.html (9 of 13)30/12/2004 11:30:50

pH curves (titration curves)

A summary of the important curves


The way you normally carry out a titration involves adding the acid to the
alkali. Here are reduced versions of the graphs described above so that
you can see them all together.

More complicated titration curves


Adding hydrochloric acid to sodium carbonate solution
The overall equation for the reaction between sodium carbonate solution
and dilute hydrochloric acid is:

If you had the two solutions of the same concentration, you would have

http://www.chemguide.co.uk/physical/acidbaseeqia/phcurves.html (10 of 13)30/12/2004 11:30:50

pH curves (titration curves)

to use twice the volume of hydrochloric acid to reach the equivalence


point - because of the 1 : 2 ratio in the equation.
Suppose you start with 25 cm3 of sodium carbonate solution, and that
both solutions have the same concentration of 1 mol dm-3. That means
that you would expect the steep drop in the titration curve to come after
you had added 50 cm3 of acid.
The actual graph looks like this:

The graph is more complicated than you might think - and curious things
happen during the titration.
You expect carbonates to produce carbon dioxide when you add acids to
them, but in the early stages of this titration, no carbon dioxide is given
off at all.
Then - as soon as you get past the half-way point in the titration - lots of
carbon dioxide is suddenly released.
The graph is showing two end points - one at a pH of 8.3 (little more than
a point of inflexion), and a second at about pH 3.7. The reaction is
obviously happening in two distinct parts.
In the first part, complete at A in the diagram, the sodium carbonate is
http://www.chemguide.co.uk/physical/acidbaseeqia/phcurves.html (11 of 13)30/12/2004 11:30:50

pH curves (titration curves)

reacting with the acid to produce sodium hydrogencarbonate:

You can see that the reaction doesn't produce any carbon dioxide.
In the second part, the sodium hydrogencarbonate produced goes on to
react with more acid - giving off lots of CO2.

That reaction is finished at B on the graph.


It is possible to pick up both of these end points by careful choice of
indicator. That is explained on the separate page on indicators.

Adding sodium hydroxide solution to dilute ethanedioic acid


Ethanedioic acid was previously known as oxalic acid. It is a diprotic
acid, which means that it can give away 2 protons (hydrogen ions) to a
base. Something which can only give away one (like HCl) is known as a
monoprotic acid.

The reaction with sodium hydroxide takes place in two stages because
one of the hydrogens is easier to remove than the other. The two
successive reactions are:

http://www.chemguide.co.uk/physical/acidbaseeqia/phcurves.html (12 of 13)30/12/2004 11:30:50

pH curves (titration curves)

If you run sodium hydroxide solution into ethanedioic acid solution, the
pH curve shows the end points for both of these reactions.

The curve is for the reaction between sodium hydroxide and ethanedioic
acid solutions of equal concentrations.

Where would you like to go now?


To the acid-base equilibria menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/acidbaseeqia/phcurves.html (13 of 13)30/12/2004 11:30:50

acid-base indicators

ACID-BASE INDICATORS

This page describes how simple acid-base indicators work, and how to
choose the right one for a particular titration.

Warning: This page assumes that you know about pH


curves for all the common acid-base combinations mentioned
by your syllabus, and weak acids (including pKa). If you aren't
happy about either of these things, you must follow these
links before you go any further.

How simple indicators work


Indicators as weak acids
Litmus
Litmus is a weak acid. It has a seriously complicated molecule which we
will simplify to HLit. The "H" is the proton which can be given away to
something else. The "Lit" is the rest of the weak acid molecule.
There will be an equilibrium established when this acid dissolves in
water. Taking the simplified version of this equilibrium:

Note: If you don't understand what I mean by "the simplified


version of this equilibrium", you need to follow up the weak
acids link before you go any further.

http://www.chemguide.co.uk/physical/acidbaseeqia/indicators.html (1 of 13)30/12/2004 11:31:04

acid-base indicators

The un-ionised litmus is red, whereas the ion is blue.


Now use Le Chatelier's Principle to work out what would happen if you
added hydroxide ions or some more hydrogen ions to this equilibrium.

Note: If you don't understand Le Chatelier's Principle, follow


this link before you go any further, and make sure that you
understand about the effect of changes of concentration on
the position of equilibrium.
Use the BACK button on your browser to return to this page.

Adding hydroxide ions:

Adding hydrogen ions:

http://www.chemguide.co.uk/physical/acidbaseeqia/indicators.html (2 of 13)30/12/2004 11:31:04

acid-base indicators

If the concentrations of HLit and Lit - are equal:


At some point during the movement of the position of equilibrium, the
concentrations of the two colours will become equal. The colour you see
will be a mixture of the two.

The reason for the inverted commas around "neutral" is that there is no
reason why the two concentrations should become equal at pH 7. For
litmus, it so happens that the 50 / 50 colour does occur at close to pH 7 that's why litmus is commonly used to test for acids and alkalis. As you
will see below, that isn't true for other indicators.
Methyl orange
Methyl orange is one of the indicators commonly used in titrations. Its
structure is simple enough to be able to see what is happening as it loses
and gains hydrogen ions.

http://www.chemguide.co.uk/physical/acidbaseeqia/indicators.html (3 of 13)30/12/2004 11:31:04

acid-base indicators

You have the same sort of equilibrium as in the litmus case - but the
colours are different.

You should be able to work out for yourself why the colour changes when
you add an acid or an alkali. The explanation is identical to the litmus
case - all that differs are the colours.

Note: If you have problems with this, it is because you don't


really understand Le Chatelier's Principle. Sort it out!
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/physical/acidbaseeqia/indicators.html (4 of 13)30/12/2004 11:31:04

acid-base indicators

In the methyl orange case, the half-way stage where the mixture of red
and yellow produces an orange colour happens at pH 3.7 - nowhere near
neutral. This will be explored further down this page.
Phenolphthalein
Phenolphthalein is another commonly used indicator for titrations, and is
another weak acid.

In this case, the weak acid is colourless and its ion is bright pink. Adding
extra hydrogen ions shifts the position of equilibrium to the left, and turns
the indicator colourless. Adding hydroxide ions removes the hydrogen
ions from the equilibrium which tips to the right to replace them - turning
the indicator pink.
The half-way stage happens at pH 9.3. Since a mixture of pink and
colourless is simply a paler pink, this is difficult to detect with any
accuracy!

The pH range of indicators


The importance of pKind
Think about a general indicator, HInd - where "Ind" is all the rest of the
indicator apart from the hydrogen ion which is given away:

Because this is just like any other weak acid, you can write an
expression for Ka for it. We will call it Kind to stress that we are talking
about the indicator.

http://www.chemguide.co.uk/physical/acidbaseeqia/indicators.html (5 of 13)30/12/2004 11:31:04

acid-base indicators

Note: If this doesn't mean anything to you, thenyou won't be


able to understand any of what follows without first reading
the page on weak acids.
Use the BACK button on your browser to return to this page.

Think of what happens half-way through the colour change. At this point
the concentrations of the acid and its ion are equal. In that case, they will
cancel out of the Kind expression.

You can use this to work out what the pH is at this half-way point. If you
re-arrange the last equation so that the hydrogen ion concentration is on
the left-hand side, and then convert to pH and pKind, you get:

That means that the end point for the indicator depends entirely on what
its pKind value is. For the indicators we've looked at above, these are:
indicator

pKind

litmus

6.5

methyl orange

3.7

http://www.chemguide.co.uk/physical/acidbaseeqia/indicators.html (6 of 13)30/12/2004 11:31:04

acid-base indicators

phenolphthalein

9.3

The pH range of indicators


Indicators don't change colour sharply at one particular pH (given by their
pKind). Instead, they change over a narrow range of pH.
Assume the equilibrium is firmly to one side, but now you add something
to start to shift it. As the equilibrium shifts, you will start to get more and
more of the second colour formed, and at some point the eye will start to
detect it.
For example, suppose you had methyl orange in an alkaline solution so
that the dominant colour was yellow. Now start to add acid so that the
equilibrium begins to shift.
At some point there will be enough of the red form of the methyl orange
present that the solution will begin to take on an orange tint. As you go
on adding more acid, the red will eventually become so dominant that
you can no longe see any yellow.
There is a gradual smooth change from one colour to the other, taking
place over a range of pH. As a rough "rule of thumb", the visible change
takes place about 1 pH unit either side of the pKind value.
The exact values for the three indicators we've looked at are:
pKind

pH range

litmus

6.5

5-8

methyl orange

3.7

3.1 - 4.4

phenolphthalein

9.3

8.3 - 10.0

indicator

The litmus colour change happens over an unusually wide range, but it is
useful for detecting acids and alkalis in the lab because it changes colour
around pH 7. Methyl orange or phenolphthalein would be less useful.

http://www.chemguide.co.uk/physical/acidbaseeqia/indicators.html (7 of 13)30/12/2004 11:31:04

acid-base indicators

This is more easily seen diagramatically.

For example, methyl orange would be yellow in any solution with a pH


greater than 4.4. It couldn't distinguish between a weak acid with a pH of
5 or a strong alkali with a pH of 14.

Choosing indicators for titrations


Remember that the equivalence point of a titration is where you have
mixed the two substances in exactly equation proportions. You obviously
need to choose an indicator which changes colour as close as possible
to that equivalence point. That varies from titration to titration.
Strong acid v strong base
The next diagram shows the pH curve for adding a strong acid to a
strong base. Superimposed on it are the pH ranges for methyl orange
and phenolphthalein.

http://www.chemguide.co.uk/physical/acidbaseeqia/indicators.html (8 of 13)30/12/2004 11:31:04

acid-base indicators

You can see that neither indicator changes colour at the equivalence
point.
However, the graph is so steep at that point that there will be virtually no
difference in the volume of acid added whichever indicator you choose.
However, it would make sense to titrate to the best possible colour with
each indicator.
If you use phenolphthalein, you would titrate until it just becomes
colourless (at pH 8.3) because that is as close as you can get to the
equivalence point.
On the other hand, using methyl orange, you would titrate until there is
the very first trace of orange in the solution. If the solution becomes red,
you are getting further from the equivalence point.

Strong acid v weak base

http://www.chemguide.co.uk/physical/acidbaseeqia/indicators.html (9 of 13)30/12/2004 11:31:04

acid-base indicators

This time it is obvious that phenolphthalein would be completely useless.


However, methyl orange starts to change from yellow towards orange
very close to the equivalence point.
You have to choose an indicator which changes colour on the steep bit of
the curve.

Weak acid v strong base

http://www.chemguide.co.uk/physical/acidbaseeqia/indicators.html (10 of 13)30/12/2004 11:31:04

acid-base indicators

This time, the methyl orange is hopeless! However, the phenolphthalein


changes colour exactly where you want it to.

Weak acid v weak base


The curve is for a case where the acid and base are both equally weak for example, ethanoic acid and ammonia solution. In other cases, the
equivalence point will be at some other pH.

You can see that neither indicator is any use. Phenolphthalein will have
finished changing well before the equivalence point, and methyl orange
falls off the graph altogether.
It may be possible to find an indicator which starts to change or finishes
changing at the equivalence point, but because the pH of the
equivalence point will be different from case to case, you can't generalise.
On the whole, you would never titrate a weak acid and a weak base in
the presence of an indicator.

Sodium carbonate solution and dilute hydrochloric acid

http://www.chemguide.co.uk/physical/acidbaseeqia/indicators.html (11 of 13)30/12/2004 11:31:04

acid-base indicators

This is an interesting special case. If you use phenolphthalein or methyl


orange, both will give a valid titration result - but the value with
phenolphthalein will be exactly half the methyl orange one.

It so happens that the phenolphthalein has finished its colour change at


exactly the pH of the equivalence point of the first half of the reaction in
which sodium hydrogencarbonate is produced.

The methyl orange changes colour at exactly the pH of the equivalence


point of the second stage of the reaction.

Where would you like to go now?


To the acid-base equilibria menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

http://www.chemguide.co.uk/physical/acidbaseeqia/indicators.html (12 of 13)30/12/2004 11:31:04

acid-base indicators

Jim Clark 2002

http://www.chemguide.co.uk/physical/acidbaseeqia/indicators.html (13 of 13)30/12/2004 11:31:04

buffer solutions

BUFFER SOLUTIONS

This page describes simple acidic and alkaline buffer solutions and
explains how they work.

What is a buffer solution?


Definition
A buffer solution is one which resists changes in pH when small
quantities of an acid or an alkali are added to it.
Acidic buffer solutions
An acidic buffer solution is simply one which has a pH less than 7. Acidic
buffer solutions are commonly made from a weak acid and one of its
salts - often a sodium salt.
A common example would be a mixture of ethanoic acid and sodium
ethanoate in solution. In this case, if the solution contained equal molar
concentrations of both the acid and the salt, it would have a pH of 4.76. It
wouldn't matter what the concentrations were, as long as they were the
same.
You can change the pH of the buffer solution by changing the ratio of
acid to salt, or by choosing a different acid and one of its salts.

Note: If you need to know about calculations involving buffer


solutions, you may be interest in my chemistry calculations
book.

http://www.chemguide.co.uk/physical/acidbaseeqia/buffers.html (1 of 7)30/12/2004 11:31:11

buffer solutions

Alkaline buffer solutions


An alkaline buffer solution has a pH greater than 7. Alkaline buffer
solutions are commonly made from a weak base and one of its salts.
A frequently used example is a mixture of ammonia solution and
ammonium chloride solution. If these were mixed in equal molar
proportions, the solution would have a pH of 9.25. Again, it doesn't
matter what concentrations you choose as long as they are the same.

How do buffer solutions work?


A buffer solution has to contain things which will remove any hydrogen
ions or hydroxide ions that you might add to it - otherwise the pH will
change. Acidic and alkaline buffer solutions achieve this in different ways.
Acidic buffer solutions
We'll take a mixture of ethanoic acid and sodium ethanoate as typical.
Ethanoic acid is a weak acid, and the position of this equilibrium will be
well to the left:

Adding sodium ethanoate to this adds lots of extra ethanoate ions.


According to Le Chatelier's Principle, that will tip the position of the
equilibrium even further to the left.

http://www.chemguide.co.uk/physical/acidbaseeqia/buffers.html (2 of 7)30/12/2004 11:31:11

buffer solutions

Note: If you don't understand Le Chatelier's Principle, follow


this link before you go any further, and make sure that you
understand about the effect of changes of concentration on
the position of equilibrium.
Use the BACK button on your browser to return to this page.

The solution will therefore contain these important things:

lots of un-ionised ethanoic acid;

lots of ethanoate ions from the sodium ethanoate;

enough hydrogen ions to make the solution acidic.

Other things (like water and sodium ions) which are present aren't
important to the argument.
Adding an acid to this buffer solution
The buffer solution must remove most of the new hydrogen ions
otherwise the pH would drop markedly.
Hydrogen ions combine with the ethanoate ions to make ethanoic acid.
Although the reaction is reversible, since the ethanoic acid is a weak
acid, most of the new hydrogen ions are removed in this way.

Since most of the new hydrogen ions are removed, the pH won't change
very much - but because of the equilibria involved, it will fall a little bit.
Adding an alkali to this buffer solution
Alkaline solutions contain hydroxide ions and the buffer solution removes
most of these.

http://www.chemguide.co.uk/physical/acidbaseeqia/buffers.html (3 of 7)30/12/2004 11:31:11

buffer solutions

This time the situation is a bit more complicated because there are two
processes which can remove hydroxide ions.
Removal by reacting with ethanoic acid
The most likely acidic substance which a hydroxide ion is going to collide
with is an ethanoic acid molecule. They will react to form ethanoate ions
and water.

Note: You might be surprised to find this written as a slightly


reversible reaction. Because ethanoic acid is a weak acid, its
conjugate base (the ethanoate ion) is fairly good at picking up
hydrogen ions again to re-form the acid. It can get these from
the water molecules. You may well find this reaction written
as one-way, but to be fussy about it, it is actually reversible!

Because most of the new hydroxide ions are removed, the pH doesn't
increase very much.
Removal of the hydroxide ions by reacting with hydrogen ions
Remember that there are some hydrogen ions present from the
ionisation of the ethanoic acid.

Hydroxide ions can combine with these to make water. As soon as this
happens, the equilibrium tips to replace them. This keeps on happening
until most of the hydroxide ions are removed.

http://www.chemguide.co.uk/physical/acidbaseeqia/buffers.html (4 of 7)30/12/2004 11:31:11

buffer solutions

Again, because you have equilibria involved, not all of the hydroxide ions
are removed - just most of them. The water formed re-ionises to a very
small extent to give a few hydrogen ions and hydroxide ions.

Alkaline buffer solutions


We'll take a mixture of ammonia and ammonium chloride solutions as
typical.
Ammonia is a weak base, and the position of this equilibrium will be well
to the left:

Adding ammonium chloride to this adds lots of extra ammonium ions.


According to Le Chatelier's Principle, that will tip the position of the
equilibrium even further to the left.
The solution will therefore contain these important things:

lots of unreacted ammonia;

lots of ammonium ions from the ammonium chloride;

enough hydroxide ions to make the solution alkaline.

Other things (like water and chloride ions) which are present aren't
important to the argument.
http://www.chemguide.co.uk/physical/acidbaseeqia/buffers.html (5 of 7)30/12/2004 11:31:11

buffer solutions

Adding an acid to this buffer solution


There are two processes which can remove the hydrogen ions that you
are adding.
Removal by reacting with ammonia
The most likely basic substance which a hydrogen ion is going to collide
with is an ammonia molecule. They will react to form ammonium ions.

Most, but not all, of the hydrogen ions will be removed. The ammonium
ion is weakly acidic, and so some of the hydrogen ions will be released
again.
Removal of the hydrogen ions by reacting with hydroxide ions
Remember that there are some hydroxide ions present from the reaction
between the ammonia and the water.

Hydrogen ions can combine with these hydroxide ions to make water. As
soon as this happens, the equilibrium tips to replace the hydroxide ions.
This keeps on happening until most of the hydrogen ions are removed.

Again, because you have equilibria involved, not all of the hydrogen ions
are removed - just most of them.

http://www.chemguide.co.uk/physical/acidbaseeqia/buffers.html (6 of 7)30/12/2004 11:31:11

buffer solutions

Adding an alkali to this buffer solution


The hydroxide ions from the alkali are removed by a simple reaction with
ammonium ions.

Because the ammonia formed is a weak base, it can react with the water
- and so the reaction is slightly reversible. That means that, again, most
(but not all) of the the hydroxide ions are removed from the solution.

Where would you like to go now?


To the acid-base equilibria menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/physical/acidbaseeqia/buffers.html (7 of 7)30/12/2004 11:31:11

Phase equilibria menu

Understanding Chemistry

PHASE EQUILIBRIA MENU

Introducing vapour pressure . . .


Explains what the saturated vapour pressure of a liquid means,
and how it varies with temperature.
Phase diagrams for pure substances . . .
Explains how to interpret a simple phase diagram for a pure
substance, including a look at the special cases of water and
carbon dioxide.
Raoult's Law and non-volatile solutes . . .
Explains Raoult's Law and how it applies to solutions containing
non-volatile solutes like salt. Shows how the lowering of vapour
pressure affects the boiling point and freezing point of the solvent.
Liquid-solid phase diagrams: tin and lead . . .
Explains the relationship between the cooling curves for liquid
mixtures of tin and lead, and the resulting phase diagram. Includes
the concept of a eutectic mixture.
More to come over the next couple of months

Go to physical chemistry menu . . .

http://www.chemguide.co.uk/physical/phaseeqiamenu.html (1 of 2)30/12/2004 11:31:12

Phase equilibria menu

Go to Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/physical/phaseeqiamenu.html (2 of 2)30/12/2004 11:31:12

saturated vapour pressure - an introduction

AN INTRODUCTION TO SATURATED VAPOUR


PRESSURE

This page looks at how the equilibrium between a liquid (or a solid) and
its vapour leads to the idea of a saturated vapour pressure. It also looks
at how saturated vapour pressure varies with temperature, and the
relationship between saturated vapour pressure and boiling point.

The origin of saturated vapour pressure


The evaporation of a liquid
The average energy of the particles in a liquid is governed by the
temperature. The higher the temperature, the higher the average energy.
But within that average, some particles have energies higher than the
average, and others have energies lower than the average.
Some of the more energetic particles on the surface of the liquid can be
moving fast enough to escape from the attractive forces holding the liquid
together. They evaporate.
The diagram shows a small region of a liquid near its surface.

Notice that evaporation only takes place on the surface of the liquid.
That's quite different from boiling which happens when there is enough
energy to disrupt the attractive forces throughout the liquid. That's why, if
http://www.chemguide.co.uk/physical/phaseeqia/vapourpress.html (1 of 9)30/12/2004 11:31:20

saturated vapour pressure - an introduction

you look at boiling water, you see bubbles of gas being formed all the
way through the liquid.
If you look at water which is just evaporating in the sun, you don't see
any bubbles. Water molecules are simply breaking away from the
surface layer.
Eventually, the water will all evaporate in this way. The energy which is
lost as the particles evaporate is replaced from the surroundings. As the
molecules in the water jostle with each other, new molecules will gain
enough energy to escape from the surface.

The evaporation of a liquid in a closed container


Now imagine what happens if the liquid is in a closed container. Common
sense tells you that water in a sealed bottle doesn't seem to evaporate or at least, it doesn't disappear over time.
But there is constant evaporation from the surface. Particles continue to
break away from the surface of the liquid - but this time they are trapped
in the space above the liquid.

As the gaseous particles bounce around, some of them will hit the
surface of the liquid again, and be trapped there. There will rapidly be an
equilibrium set up in which the number of particles leaving the surface is
http://www.chemguide.co.uk/physical/phaseeqia/vapourpress.html (2 of 9)30/12/2004 11:31:20

saturated vapour pressure - an introduction

exactly balanced by the number rejoining it.

In this equilibrium, there will be a fixed number of the gaseous particles


in the space above the liquid.
When these particles hit the walls of the container, they exert a pressure.
This pressure is called the saturated vapour pressure (also known as
saturation vapour pressure) of the liquid.

Measuring the saturated vapour pressure


It isn't difficult to show the existence of this saturated vapour pressure
(and to measure it) using a simple piece of apparatus.

Note: This experiment is much easier to talk about than do,


given the safety problems in handling mercury because of its
poisonous vapour. This is particularly going to be a problem if
you want to find the saturated vapour pressure of a liquid at a
higher temperature. You would have to use a more complex
bit of apparatus. That isn't a problem you need to worry about
for UK A level purposes.

http://www.chemguide.co.uk/physical/phaseeqia/vapourpress.html (3 of 9)30/12/2004 11:31:20

saturated vapour pressure - an introduction

If you have a mercury barometer tube in a trough of mercury, at 1


atmosphere pressure the column will be 760 mm tall. 1 atmosphere is
sometimes quoted as 760 mmHg ("millimetres of mercury").

If you squirt a few drops of liquid into the tube, it will rise to form a thin
layer floating on top of the mercury. Some of the liquid will evaporate and
you will get the equilibrium we've just been talking about - provided there
is still some liquid on top of the mercury. It is only an equilibrium if both
liquid and vapour are present.

http://www.chemguide.co.uk/physical/phaseeqia/vapourpress.html (4 of 9)30/12/2004 11:31:20

saturated vapour pressure - an introduction

The saturated vapour pressure of the liquid will force the mercury level
down a bit. You can measure the drop - and this gives a value for the
saturated vapour pressure of the liquid at this temperature. In this case,
the mercury has been forced down by a distance of 760 - 630 mm. The
saturated vapour pressure of this liquid at the temperature of the
experiment is 130 mmHg.
You could convert this into proper SI units (pascals) if you wanted to. 760
mmHg is equivalent to 101325 Pa.
A value of 130 mmHg is quite a high vapour pressure if we are talking
about room temperature. Water's saturated vapour pressure is about 20
mmHg at this temperature. A high vapour pressure means that the liquid
must be volatile - molecules escape from its surface relatively easily, and
aren't very good at sticking back on again either.
That will result in larger numbers of them in the gas state once
equilibrium is reached.
The liquid in the example must have significantly weaker intermolecular
forces than water.

http://www.chemguide.co.uk/physical/phaseeqia/vapourpress.html (5 of 9)30/12/2004 11:31:20

saturated vapour pressure - an introduction

The variation of saturated vapour pressure with


temperature
The effect of temperature on the equilibrium between liquid and
vapour
You can look at this in two ways.
There is a common sense way. If you increase the temperature, you are
increasing the average energy of the particles present. That means that
more of them are likely to have enough energy to escape from the
surface of the liquid. That will tend to increase the saturated vapour
pressure.
Or you can look at it in terms of Le Chatelier's Principle - which works
just as well in this kind of physical situation as it does in the more familiar
chemical examples.

Note: You could follow this link if you aren't sure about Le
Chatelier's Principle.
Use the BACK button on your browser to return to this page
later.

When the space above the liquid is saturated with vapour particles, you
have this equilibrium occurring on the surface of the liquid:

The forward change (liquid to vapour) is endothermic. It needs heat to


convert the liquid into the vapour.
According to Le Chatelier, increasing the temperature of a system in a
dynamic equilibrium favours the endothermic change. That means that
increasing the temperature increases the amount of vapour present, and
so increases the saturated vapour pressure.

http://www.chemguide.co.uk/physical/phaseeqia/vapourpress.html (6 of 9)30/12/2004 11:31:20

saturated vapour pressure - an introduction

The effect of temperature on the saturated vapour pressure of water


The graph shows how the saturated vapour pressure (svp) of water
varies from 0C to 100 C. The pressure scale (the vertical one) is
measured in kilopascals (kPa). 1 atmosphere pressure is 101.325 kPa.

Saturated vapour pressure and boiling point


A liquid boils when its saturated vapour pressure becomes equal to the
external pressure on the liquid. When that happens, it enables bubbles of
vapour to form throughout the liquid - those are the bubbles you see
when a liquid boils.
If the external pressure is higher than the saturated vapour pressure,
these bubbles are prevented from forming, and you just get evaporation
at the surface of the liquid.
If the liquid is in an open container and exposed to normal atmospheric
pressure, the liquid boils when its saturated vapour pressure becomes
equal to 1 atmosphere (or 101325 Pa or 101.325 kPa or 760 mmHg).
This happens with water when the temperature reaches 100C.
http://www.chemguide.co.uk/physical/phaseeqia/vapourpress.html (7 of 9)30/12/2004 11:31:20

saturated vapour pressure - an introduction

But at different pressures, water will boil at different temperatures. For


example, at the top of Mount Everest the pressure is so low that water
will boil at about 70C. Depressions from the Atlantic can easily lower the
atmospheric pressure in the UK enough so that water will boil at 99C even lower with very deep depressions.
Whenever we just talk about "the boiling point" of a liquid, we always
assume that it is being measured at exactly 1 atmosphere pressure. In
practice, of course, that is rarely exactly true.

Saturated vapour pressure and solids


Sublimation
Solids can also lose particles from their surface to form a vapour, except
that in this case we call the effect sublimation rather than evaporation.
Sublimation is the direct change from solid to vapour (or vice versa)
without going through the liquid stage.
In most cases, at ordinary temperatures, the saturated vapour pressures
of solids range from low to very, very, very low. The forces of attraction in
many solids are too high to allow much loss of particles from the surface.
However, there are some which do easily form vapours. For example,
naphthalene (used in old-fashioned "moth balls" to deter clothes moths)
has quite a strong smell. Molecules must be breaking away from the
surface as a vapour, because otherwise you wouldn't be able to smell it.
Another fairly common example (discussed in detail on another page) is
solid carbon dioxide - "dry ice". This never forms a liquid at atmospheric
pressure and always converts directly from solid to vapour. That's why it
is known as dry ice.

http://www.chemguide.co.uk/physical/phaseeqia/vapourpress.html (8 of 9)30/12/2004 11:31:20

saturated vapour pressure - an introduction

Note: This link will take you straight to the page about phase
diagrams where this is discussed in detail. Follow this link as
well if you are interested in the logical next step in this topic
which is a discussion of the phase diagrams of pure
substances, including water and carbon dioxide.

Where would you like to go now?


To the phase equilibrium menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/physical/phaseeqia/vapourpress.html (9 of 9)30/12/2004 11:31:20

phase diagrams of pure substances

PHASE DIAGRAMS OF PURE SUBSTANCES

This page explains how to interpret the phase diagrams for simple pure
substances - including a look at the special cases of the phase diagrams
of water and carbon dioxide. This is going to be a long page, because I
have tried to do the whole thing as gently as possible.

The basic phase diagram


What is a phase?
At its simplest, a phase can be just another term for solid, liquid or gas. If
you have some ice floating in water, you have a solid phase present and
a liquid phase. If there is air above the mixture, then that is another
phase.
But the term can be used more generally than this. For example, oil
floating on water also consists of two phases - in this case, two liquid
phases. If the oil and water are contained in a bucket, then the solid
bucket is yet another phase. In fact, there might be more than one solid
phase if the handle is attached separately to the bucket rather than
moulded as a part of the bucket.
You can recognise the presence of the different phases because there is
an obvious boundary between them - a boundary between the solid ice
and the liquid water, for example, or the boundary between the two
liquids.

Phase diagrams
A phase diagram lets you work out exactly what phases are present at
any given temperature and pressure. In the cases we'll be looking at on
this page, the phases will simply be the solid, liquid or vapour (gas)
states of a pure substance.

http://www.chemguide.co.uk/physical/phaseeqia/phasediags.html (1 of 17)30/12/2004 11:31:31

phase diagrams of pure substances

This is the phase diagram for a typical pure substance.

These diagrams (including this one) are nearly always drawn highly
distorted in order to see what is going on more easily. There are usually
two major distortions. We'll discuss these when they become relevant.
If you look at the diagram, you will see that there are three lines, three
areas marked "solid", "liquid" and "vapour", and two special points
marked "C" and "T".
The three areas
These are easy! Suppose you have a pure substance at three different
sets of conditions of temperature and pressure corresponding to 1, 2 and
3 in the next diagram.

http://www.chemguide.co.uk/physical/phaseeqia/phasediags.html (2 of 17)30/12/2004 11:31:31

phase diagrams of pure substances

Under the set of conditions at 1 in the diagram, the substance would be a


solid because it falls into that area of the phase diagram. At 2, it would be
a liquid; and at 3, it would be a vapour (a gas).

Note: I'm using the terms vapour and gas as if they were
interchangeable. There are subtle differences between them
that I'm not ready to explain for a while yet. Be patient!

Moving from solid to liquid by changing the temperature:


Suppose you had a solid and increased the temperature while keeping
the pressure constant - as shown in the next diagram. As the
temperature increases to the point where it crosses the line, the solid will
turn to liquid. In other words, it melts.

http://www.chemguide.co.uk/physical/phaseeqia/phasediags.html (3 of 17)30/12/2004 11:31:31

phase diagrams of pure substances

If you repeated this at a higher fixed pressure, the melting temperature


would be higher because the line between the solid and liquid areas
slopes slightly forward.

http://www.chemguide.co.uk/physical/phaseeqia/phasediags.html (4 of 17)30/12/2004 11:31:31

phase diagrams of pure substances

Note: This is one of the cases where we distort these


diagrams to make them easier to discuss. This line is much
more vertical in practice than we normally draw it. There
would be very little change in melting point at a higher
pressure. The diagram would be very difficult to follow if we
didn't exaggerate it a bit.

So what actually is this line separating the solid and liquid areas of the
diagram?
It simply shows the effect of pressure on melting point.
Anywhere on this line, there is an equilibrium between solid and liquid.
You can apply Le Chatelier's Principle to this equilibrium just as if it was
a chemical equilibrium. If you increase the pressure, the equilibrium will
move in such a way as to counter the change you have just made.

If it converted from liquid to solid, the pressure would tend to decrease


again because the solid takes up slightly less space for most substances.
That means that increasing the pressure on the equilibrium mixture of
solid and liquid at its original melting point will convert the mixture back
into the solid again. In other words, it will no longer melt at this
temperature.
To make it melt at this higher pressure, you will have to increase the
temperature a bit. Raising the pressure raises the melting point of most
solids. That's why the melting point line slopes forward for most
substances.
Moving from solid to liquid by changing the pressure:
http://www.chemguide.co.uk/physical/phaseeqia/phasediags.html (5 of 17)30/12/2004 11:31:31

phase diagrams of pure substances

You can also play around with this by looking at what happens if you
decrease the pressure on a solid at constant temperature.

Note: You have got to be a bit careful about this, because


exactly what happens if you decrease the pressure depends
on exactly what your starting conditions are. We'll talk some
more about this when we look at the line separating the solid
region from the vapour region.

Moving from liquid to vapour:


In the same sort of way, you can do this either by changing the
temperature or the pressure.

http://www.chemguide.co.uk/physical/phaseeqia/phasediags.html (6 of 17)30/12/2004 11:31:31

phase diagrams of pure substances

The liquid will change to a vapour - it boils - when it crosses the


boundary line between the two areas. If it is temperature that you are
varying, you can easily read off the boiling temperature from the phase
diagram. In the diagram above, it is the temperature where the red arrow
crosses the boundary line.
So, again, what is the significance of this line separating the two areas?
Anywhere along this line, there will be an equilibrium between the liquid
and the vapour. The line is most easily seen as the effect of pressure on
the boiling point of the liquid.
As the pressure increases, so the boiling point increases.

http://www.chemguide.co.uk/physical/phaseeqia/phasediags.html (7 of 17)30/12/2004 11:31:31

phase diagrams of pure substances

Note: I don't want to make any very big deal over this, but
this line is actually exactly the same as the graph for the
effect of temperature on the saturated vapour pressure of the
liquid. Saturated vapour pressure is dealt with on a separate
page. A liquid will boil when its saturated vapour pressure is
equal to the external pressure.
Suppose you measured the saturated vapour pressure of a
liquid at 50C, and it turned out to be 75 kPa. You could plot
that as one point on a vapour pressure curve, and then go on
to measure other saturated vapour pressures at different
temperatures and plot those as well.
Now, suppose that you had the liquid exposed to a total
external pressure of 75 kPa, and gradually increased the
temperature. The liquid would boil when its saturated vapour
pressure became equal to the external pressure - in this case
at 50C. If you have the complete vapour pressure curve, you
could equally well find the boiling point corresponding to any
other external pressure.
That means that the plot of saturated vapour pressure
against temperature is exactly the same as the curve relating
boiling point and external pressure - they are just two ways of
looking at the same thing.
If all you are interested in doing is interpreting one of these
phase diagrams, you probably don't have to worry too much
about this.

http://www.chemguide.co.uk/physical/phaseeqia/phasediags.html (8 of 17)30/12/2004 11:31:31

phase diagrams of pure substances

The critical point


You will have noticed that this liquid-vapour equilibrium curve has a top
limit that I have labelled as C in the phase diagram.
This is known as the critical point. The temperature and pressure
corresponding to this are known as the critical temperature and critical
pressure.
If you increase the pressure on a gas (vapour) at a temperature lower
than the critical temperature, you will eventually cross the liquid-vapour
equilibrium line and the vapour will condense to give a liquid.

This works fine as long as the gas is below the critical temperature.
What, though, if your temperature was above the critical temperature?
There wouldn't be any line to cross!
That is because, above the critical temperature, it is impossible to
condense a gas into a liquid just by increasing the pressure. All you get
is a highly compressed gas. The particles have too much energy for the
intermolecular attractions to hold them together as a liquid.
The critical temperature obviously varies from substance to substance
and depends on the strength of the attractions between the particles. The
http://www.chemguide.co.uk/physical/phaseeqia/phasediags.html (9 of 17)30/12/2004 11:31:31

phase diagrams of pure substances

stronger the intermolecular attractions, the higher the critical temperature.

Note: This is now a good point for a quick comment about


the use of the words "gas" and "vapour". To a large extent
you just use the term which feels right. You don't usually talk
about "ethanol gas", although you would say "ethanol
vapour". Equally, you wouldn't talk about oxygen as being a
vapour - you always call it a gas.
There are various guide-lines that you can use if you want to.
For example, if the substance is commonly a liquid at or
around room temperature, you tend to call what comes away
from it a vapour. A slightly wider use would be to call it a
vapour if the substance is below its critical point, and a gas if
it is above it. Certainly it would be unusual to call anything a
vapour if it was above its critical point at room temperature oxygen or nitrogen or hydrogen, for example. These would all
be described as gases.
This is absolutely NOT something that is at all worth getting
worked up about!

Moving from solid to vapour:


There's just one more line to look at on the phase diagram. This is the
line in the bottom left-hand corner between the solid and vapour areas.
That line represents solid-vapour equilibrium. If the conditions of
temperature and pressure fell exactly on that line, there would be solid
and vapour in equilibrium with each other - the solid would be subliming.
(Sublimation is the change directly from solid to vapour or vice versa
without going through the liquid phase.)
Once again, you can cross that line by either increasing the temperature
of the solid, or decreasing the pressure.
The diagram shows the effect of increasing the temperature of a solid at
a (probably very low) constant pressure. The pressure obviously has to
be low enough that a liquid can't form - in other words, it has to happen
http://www.chemguide.co.uk/physical/phaseeqia/phasediags.html (10 of 17)30/12/2004 11:31:31

phase diagrams of pure substances

below the point labelled as T.

You could read the sublimation temperature off the diagram. It will be the
temperature at which the line is crossed.

The triple point


Point T on the diagram is called the triple point.
If you think about the three lines which meet at that point, they represent
conditions of:

solid-liquid equilibrium

liquid-vapour equilibrium

solid-vapour equilibrium

Where all three lines meet, you must have a unique combination of
temperature and pressure where all three phases are in equilibrium
together. That's why it is called a triple point.
http://www.chemguide.co.uk/physical/phaseeqia/phasediags.html (11 of 17)30/12/2004 11:31:31

phase diagrams of pure substances

If you controlled the conditions of temperature and pressure in order to


land on this point, you would see an equilibrium which involved the solid
melting and subliming, and the liquid in contact with it boiling to produce
a vapour - and all the reverse changes happening as well.
If you held the temperature and pressure at those values, and kept the
system closed so that nothing escaped, that's how it would stay. A
strange set of affairs!

Normal melting and boiling points


The normal melting and boiling points are those when the pressure is 1
atmosphere. These can be found from the phase diagram by drawing a
line across at 1 atmosphere pressure.

The phase diagram for water

http://www.chemguide.co.uk/physical/phaseeqia/phasediags.html (12 of 17)30/12/2004 11:31:31

phase diagrams of pure substances

There is only one difference between this and the phase diagram that
we've looked at up to now. The solid-liquid equilibrium line (the melting
point line) slopes backwards rather than forwards.
In the case of water, the melting point gets lower at higher pressures.
Why?

If you have this equilibrium and increase the pressure on it, according to
Le Chatelier's Principle the equilibrium will move to reduce the pressure
again. That means that it will move to the side with the smaller volume.
Liquid water is produced.
To make the liquid water freeze again at this higher pressure, you will
have to reduce the temperature. Higher pressures mean lower melting
(freezing) points.
Now lets put some numbers on the diagram to show the exact positions
of the critical point and triple point for water.

http://www.chemguide.co.uk/physical/phaseeqia/phasediags.html (13 of 17)30/12/2004 11:31:31

phase diagrams of pure substances

Notice that the triple point for water occurs at a very low pressure. Notice
also that the critical temperature is 374C. It would be impossible to
convert water from a gas to a liquid by compressing it above this
temperature.
The normal melting and boiling points of water are found in exactly the
same way as we have already discussed - by seeing where the 1
atmosphere pressure line crosses the solid-liquid and then the liquidvapour equilibrium lines.

Note: Further up the page I mentioned two ways in which


these diagrams are distorted to make them easier to follow. I
have already pointed out that the solid-liquid equilibrium line
should really be much more vertical. This last diagram
illustrates the other major distortion - which is to the scales of
both pressure and temperature. Look, for example, at the
gaps between the various quoted pressure figures and then
imagine that you had to plot those on a bit of graph paper!
The temperature scale is equally haphazard.

http://www.chemguide.co.uk/physical/phaseeqia/phasediags.html (14 of 17)30/12/2004 11:31:31

phase diagrams of pure substances

Just one final example of using this diagram (because it appeals to me).
Imagine lowering the pressure on liquid water along the line in the
diagram below.

The phase diagram shows that the water would first freeze to form ice as
it crossed into the solid area. When the pressure fell low enough, the ice
would then sublime to give water vapour. In other words, the change is
from liquid to solid to vapour. I find that satisfyingly bizarre!

The phase diagram for carbon dioxide

http://www.chemguide.co.uk/physical/phaseeqia/phasediags.html (15 of 17)30/12/2004 11:31:31

phase diagrams of pure substances

The only thing special about this phase diagram is the position of the
triple point which is well above atmospheric pressure. It is impossible to
get any liquid carbon dioxide at pressures less than 5.11 atmospheres.
That means that at 1 atmosphere pressure, carbon dioxide will sublime
at a temperature of -78C.
This is the reason that solid carbon dioxide is often known as "dry ice".
You can't get liquid carbon dioxide under normal conditions - only the
solid or the vapour.

Where would you like to go now?


To the phase equilibrium menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

http://www.chemguide.co.uk/physical/phaseeqia/phasediags.html (16 of 17)30/12/2004 11:31:31

phase diagrams of pure substances

Jim Clark 2004

http://www.chemguide.co.uk/physical/phaseeqia/phasediags.html (17 of 17)30/12/2004 11:31:31

Raoult's Law and non-volatile solutes

RAOULT'S LAW AND NON-VOLATILE SOLUTES

This page deals with Raoult's Law and how it applies to solutions in
which the solute is non-volatile - for example, a solution of salt in water.
A non-volatile solute (the salt, for example) hasn't got any tendency to
form a vapour at the temperature of the solution.
It goes on to explain how the resulting lowering of vapour pressure
affects the boiling point and freezing point of the solution.

Important: If you haven't already read the page about


saturated vapour pressure, you should follow this link before
you go on.
Use the BACK button on your browser to return to this page
when you are ready.

Raoult's Law
There are several ways of stating Raoult's Law, and you tend to use
slightly different versions depending on the situation you are talking
about. You can use the simplified definition in the box below in the case
of a single volatile liquid (the solvent) and a non-volatile solute.

The vapour pressure of a solution of a non-volatile solute is equal


to the vapour pressure of the pure solvent at that temperature
multiplied by its mole fraction.

In equation form, this reads:


http://www.chemguide.co.uk/physical/phaseeqia/raoultnonvol.html (1 of 12)30/12/2004 11:31:41

Raoult's Law and non-volatile solutes

In this equation, Po is the vapour pressure of the pure solvent at a


particular temperature.
xsolv is the mole fraction of the solvent. That is exactly what it says it is the fraction of the total number of moles present which is solvent.
You calculate this using:

Suppose you had a solution containing 10 moles of water and 0.1 moles
of sugar. The total number of moles is therefore 10.1
The mole fraction of the water is:

A simple explanation of why Raoult's Law works


There are two ways of explaining why Raoult's Law works - a simple
visual way, and a more sophisticated way based on entropy. Because of
the level I am aiming at, I'm just going to look at the simple way.
Remember that saturated vapour pressure is what you get when a liquid
is in a sealed container. An equilibrium is set up where the number of
particles breaking away from the surface is exactly the same as the
number sticking on to the surface again.

http://www.chemguide.co.uk/physical/phaseeqia/raoultnonvol.html (2 of 12)30/12/2004 11:31:41

Raoult's Law and non-volatile solutes

Now suppose you added enough solute so that the solvent molecules
only occupied 50% of the surface of the solution.

A certain fraction of the solvent molecules will have enough energy to


escape from the surface (say, 1 in 1000 or 1 in a million, or whatever). If
you reduce the number of solvent molecules on the surface, you are
going to reduce the number which can escape in any given time.
But it won't make any difference to the ability of molecules in the vapour
to stick to the surface again. If a solvent molecule in the vapour hits a bit
of surface occupied by the solute particles, it may well stick. There are
obviously attractions between solvent and solute otherwise you wouldn't
have a solution in the first place.
http://www.chemguide.co.uk/physical/phaseeqia/raoultnonvol.html (3 of 12)30/12/2004 11:31:41

Raoult's Law and non-volatile solutes

The net effect of this is that when equilibrium is established, there will be
fewer solvent molecules in the vapour phase - it is less likely that they
are going to break away, but there isn't any problem about them
returning.
If there are fewer particles in the vapour at equilibrium, the saturated
vapour pressure is lower.

Limitations on Raoult's Law


Raoult's Law only works for ideal solutions. An ideal solution is defined
as one which obeys Raoult's Law.
Features of an ideal solution
In practice, there's no such thing! However, very dilute solutions obey
Raoult's Law to a reasonable approximation. The solution in the last
diagram wouldn't actually obey Raoult's Law - it is far too concentrated. I
had to draw it that concentrated to make the point more clearly.
In an ideal solution, it takes exactly the same amount of energy for a
solvent molecule to break away from the surface of the solution as it did
in the pure solvent. The forces of attraction between solvent and solute
are exactly the same as between the original solvent molecules - not a
very likely event!

http://www.chemguide.co.uk/physical/phaseeqia/raoultnonvol.html (4 of 12)30/12/2004 11:31:41

Raoult's Law and non-volatile solutes

Suppose that in the pure solvent, 1 in 1000 molecules had enough


energy to overcome the intermolecular forces and break away from the
surface in any given time. In an ideal solution, that would still be exactly
the same proportion.
Fewer would, of course, break away because there are now fewer
solvent molecules on the surface - but of those that are on the surface,
the same proportion still break away.
If there were strong solvent-solute attractions, this proportion may be
reduced to 1 in 2000, or 1 in 5000 or whatever.
In any real solution of, say, a salt in water, there are strong attractions
between the water molecules and the ions. That would tend to slow down
the loss of water molecules from the surface. However, if the solution is
sufficiently dilute, there will be good-sized regions on the surface where
you still have water molecules on their own. The solution will then
approach ideal behaviour.

The nature of the solute


There is another thing that you have to be careful of if you are going to
do any calculations on Raoult's Law (beyond the scope of this site). You
may have noticed in the little calculation about mole fraction further up
the page, that I used sugar as a solute rather than salt. There was a
good reason for that!
What matters isn't actually the number of moles of substance that you
put into the solution, but the number of moles of particles formed. For
each mole of sodium chloride dissolved, you get 1 mole of sodium ions
and 1 mole of chloride ions - in other words, you get twice the number of
moles of particles as of original salt.

So, if you added 0.1 moles of sodium chloride, there would actually be
0.2 moles of particles in the solution - and that's the figure you would
http://www.chemguide.co.uk/physical/phaseeqia/raoultnonvol.html (5 of 12)30/12/2004 11:31:41

Raoult's Law and non-volatile solutes

have to use in the mole fraction calculation.


Unless you think carefully about it, Raoult's Law only works for solutes
which don't change their nature when they dissolve. For example, they
mustn't ionise or associate (in other words, if you put in substance A, it
mustn't form A2 in solution).
If it does either of these things, you have to treat Raoult's law with great
care.

Note: This isn't a problem you are likely to have to worry


about if you are a UK A level student. Just be aware that the
problem exists.

Raoult's Law and melting and boiling points


The effect of Raoult's Law is that the saturated vapour pressure of a
solution is going to be lower than that of the pure solvent at any particular
temperature. That has important effects on the phase diagram of the
solvent.
The next diagram shows the phase diagram for pure water in the region
around its normal melting and boiling points. The 1 atmosphere line
shows the conditions for measuring the normal melting and boiling points.

http://www.chemguide.co.uk/physical/phaseeqia/raoultnonvol.html (6 of 12)30/12/2004 11:31:41

Raoult's Law and non-volatile solutes

Note: In common with most phase diagrams, this is drawn


highly distorted in order to show more clearly what is going
on.
If you haven't already read my page about phase diagrams
for pure substances, you should follow this link before you go
on to make proper sense of what comes next.
Use the BACK button on your browser to return to this page
when you are ready.

http://www.chemguide.co.uk/physical/phaseeqia/raoultnonvol.html (7 of 12)30/12/2004 11:31:41

Raoult's Law and non-volatile solutes

The line separating the liquid and vapour regions is the set of conditions
where liquid and vapour are in equilibrium.
It can be thought of as the effect of pressure on the boiling point of the
water, but it is also the curve showing the effect of temperature on the
saturated vapour pressure of the water. These two ways of looking at the
same line are discussed briefly in a note about half-way down the page
about phase diagrams (follow the last link above).
If you draw the saturated vapour pressure curve for a solution of a nonvolatile solute in water, it will always be lower than the curve for the pure
water.

http://www.chemguide.co.uk/physical/phaseeqia/raoultnonvol.html (8 of 12)30/12/2004 11:31:41

Raoult's Law and non-volatile solutes

Note: The curves for the pure water and for the solution are
often drawn parallel to each other. That has got to be wrong!
Suppose you have a solution where the mole fraction of the
water is 0.99 and the vapour pressure of the pure water at
that temperature is 100 kPa. The vapour pressure of the
solution will be 99 kPa - a fall of 1 kPa. At a lower
temperature, where the vapour pressure of the pure water is
10 kPa, the fall will only be 0.1 kPa. For the curves to be
parallel the falls would have to be the same over the whole
temperature range. They aren't!

If you look closely at the last diagram, you will see that the point at which
the liquid-vapour equilibrium curve meets the solid-vapour curve has
moved. That point is the triple point of the system - a unique set of
temperature and pressure conditions at which it is possible to get solid,
liquid and vapour all in equilibrium with each other at the same time.
Since the triple point has solid-liquid equilibrium present (amongst other
equilibria), it is also a melting point of the system - although not the
normal melting point because the pressure isn't 1 atmosphere.
That must mean that the phase diagram needs a new melting point line
(a solid-liquid equilibrium line) passing through the new triple point. That
is shown in the next diagram.

http://www.chemguide.co.uk/physical/phaseeqia/raoultnonvol.html (9 of 12)30/12/2004 11:31:41

Raoult's Law and non-volatile solutes

Now we are finally in a position to see what effect a non-volatile solute


has on the melting and freezing points of the solution. Look at what
happens when you draw in the 1 atmosphere pressure line which lets
you measure the melting and boiling points. The diagram also includes
the melting and boiling points of the pure water from the original phase
diagram for pure water (black lines).

http://www.chemguide.co.uk/physical/phaseeqia/raoultnonvol.html (10 of 12)30/12/2004 11:31:41

Raoult's Law and non-volatile solutes

Because of the changes to the phase diagram, you can see that:

the boiling point of the solvent in a solution is higher than that of


the pure solvent;
the freezing point (melting point) of the solvent in a solution is
lower than that of the pure solvent.

We have looked at this with water as the solvent, but using a different
solvent would make no difference to the argument or the conclusions.
The only difference is in the slope of the solid-liquid equilibrium lines. For
most solvents, these slope forwards whereas the water line slopes
backwards. You could prove to yourself that that doesn't affect what we
have been looking at by re-drawing all these diagrams with the slope of
that particular line changed.
You will find it makes no difference whatsoever.

Where would you like to go now?


To the phase equilibrium menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/physical/phaseeqia/raoultnonvol.html (11 of 12)30/12/2004 11:31:41

Raoult's Law and non-volatile solutes

http://www.chemguide.co.uk/physical/phaseeqia/raoultnonvol.html (12 of 12)30/12/2004 11:31:41

solid-liquid phase diagrams: tin and lead

SOLID-LIQUID PHASE DIAGRAMS: TIN AND


LEAD

This page explains the relationship between the cooling curves for liquid
mixtures of tin and lead, and the resulting phase diagram. It also offers a
simple introduction to the idea of a eutectic mixture.

Important: This page is only really designed to be an


introduction to the topic suitable for courses for 16-18 year
olds such as UK A level chemistry. Be aware that the phase
diagram used is a simplified version of the real thing. In
particular, it ignores the formation of solid solutions of tin and
lead.

Cooling curves
Cooling curves for pure substances
Suppose you have some pure molten lead and allow it to cool down until
it has all solidified, plotting the temperature of the lead against time as
you go. You would end up with a typical cooling curve for a pure
substance.

http://www.chemguide.co.uk/physical/phaseeqia/snpb.html (1 of 16)30/12/2004 11:31:53

solid-liquid phase diagrams: tin and lead

Note: Just before the liquid freezes, there is sometimes a


slight dip in the curve below the freezing point. This is called
supercooling. As soon as some solid forms, the temperature
recovers to the normal freezing point. Supercooling isn't
important for the present discussion, so I'm ignoring it on the
diagrams.

Throughout the whole experiment, heat is being lost to the surroundings and yet the temperature doesn't fall at all while the lead is freezing. This
is because the freezing process liberates heat at exactly the same rate
that it is being lost to the surroundings.
Energy is released when new bonds form - in this case, the strong
metallic bonds in the solid lead.
If you repeated this process for pure liquid tin, the shape of the graph
would be exactly the same, except that the freezing point would now be
at 232C. (The graph for this is further down the page.)

Cooling curves for tin-lead mixtures

http://www.chemguide.co.uk/physical/phaseeqia/snpb.html (2 of 16)30/12/2004 11:31:53

solid-liquid phase diagrams: tin and lead

A sample curve
If you add some tin to the lead, the shape of the cooling curve changes.
The next graph shows what happens if you cool a liquid mixture
containing about 67% lead and 33% tin by mass.

There are lots of things to look at:

Notice that nothing happens at all at the normal freezing point of


the lead. Adding the tin to it lowers its freezing point.
Freezing starts for this mixture at about 250C. You would start to
get some solid lead formed - but no tin. At that point the rate of
cooling slows down - the curve gets less steep.
However, the graph doesn't go horizontal yet. Although energy is
being given off as the lead turns to a solid, there isn't anything
similar happening to the tin. That means that there isn't enough
energy released to keep the temperature constant.

The temperature does stop falling at 183C. Now both tin and lead
are freezing. Once everything has solidified, the temperature
continues to fall.

http://www.chemguide.co.uk/physical/phaseeqia/snpb.html (3 of 16)30/12/2004 11:31:53

solid-liquid phase diagrams: tin and lead

Changing the proportions of tin and lead


If you had less tin in the mixture, the overall shape of the curve stays
much the same, but the point at which the lead first starts to freeze
changes.
The less tin there is, the smaller the drop in the freezing point of the lead.
For a mixture containing only 20% of tin, the freezing point of the lead is
about 275C. That's where the graph would suddenly become less steep.
BUT . . . you will still get the graph going horizontal (showing the freezing
of both the tin and lead) at exactly the same temperature: 183C.
As you increase the proportion of tin, the first signs of solid lead appear
at lower and lower temperatures, but the final freezing of the whole
mixture still happens at 183C.
That continues until you have added enough tin that the mixture contains
62% tin and 38% lead. At that point, the graph changes.

This particular mixture of lead and tin has a cooling curve which looks
exactly like that of a pure substance rather than a mixture. There is just
the single horizontal part of the graph where everything is freezing.

http://www.chemguide.co.uk/physical/phaseeqia/snpb.html (4 of 16)30/12/2004 11:31:53

solid-liquid phase diagrams: tin and lead

However, it is still a mixture. If you use a microscope to look at the solid


formed after freezing, you can see the individual crystals of tin and lead.
This particular mixture is known as a eutectic mixture. The word
"eutectic" comes from Greek and means "easily melted".
The eutectic mixture has the lowest melting point (which is, of course, the
same as the freezing point) of any mixture of lead and tin. The
temperature at which the eutectic mixture freezes or melts is known as
the eutectic temperature.

What happens if there is more than 62% of tin in the mixture?


You can trace it through in exactly the same way, by imagining starting
with pure tin and then adding lead to it.
The cooling curve for pure liquid tin looks like this:

It's just like the pure lead cooling curve except that tin's freezing point is
lower.
If you add small amounts of lead to the tin, so that you have perhaps
80% tin and 20% lead, you will get a curve like this:
http://www.chemguide.co.uk/physical/phaseeqia/snpb.html (5 of 16)30/12/2004 11:31:53

solid-liquid phase diagrams: tin and lead

Notice the lowered freezing point of the tin. Notice also the final freezing
of the whole mixture again takes place at 183C.
As you increase the amount of lead (or decrease the amount of tin same thing!) until there is 62% of tin and 38% of lead, you will again get
the eutectic mixture with the curve we've already looked at.

http://www.chemguide.co.uk/physical/phaseeqia/snpb.html (6 of 16)30/12/2004 11:31:53

solid-liquid phase diagrams: tin and lead

The phase diagram


Constructing the phase diagram
You start from data obtained from the cooling curves. You draw a graph
of the temperature at which freezing first starts against the proportion of
tin and lead in the mixture. The only unusual thing is that you draw the
temperature scale at each end of the diagram instead of only at the lefthand side.

Notice that at the left-hand side and right-hand sides of the curves you
have the freezing points (melting points) of the pure lead and tin.

http://www.chemguide.co.uk/physical/phaseeqia/snpb.html (7 of 16)30/12/2004 11:31:53

solid-liquid phase diagrams: tin and lead

Note: The two lines meeting at the eutectic point have been
simplified slightly so that they are drawn as straight lines
rather than slight curves. It doesn't affect the argument in any
way. I haven't been able to find the actual data to plot them
accurately, so the simplification is to avoid giving the
impression that I actually know exactly what the curves look
like!

To finish off the phase diagram, all you have to do is draw a single
horizontal line across at the eutectic temperature. Then you label each
area of the diagram with what you would find under the various different
conditions.

Using the phase diagram

http://www.chemguide.co.uk/physical/phaseeqia/snpb.html (8 of 16)30/12/2004 11:31:53

solid-liquid phase diagrams: tin and lead

Important: One of the problems with the various sorts of


phase diagrams is that they are all interpreted slightly
differently. You can't assume that because you know how to
use one sort of phase diagram that you can treat others
exactly the same. If you already know about other phase
diagrams, look at this one completely afresh.

Suppose you have a mixture of 67% lead and 33% tin. That's the mixture
from the first cooling curve plotted above. Suppose it is at a temperature
of 300C.
That corresponds to a set of conditions in the area of the phase diagram
labelled as molten tin and lead.

Now consider what happens if you cool that mixture. Eventually the
temperature will drop to a point where it crosses the line into the next

http://www.chemguide.co.uk/physical/phaseeqia/snpb.html (9 of 16)30/12/2004 11:31:53

solid-liquid phase diagrams: tin and lead

region of the diagram.


At that point, the mixture will start to produce some solid lead - in other
words, the lead (but not the tin) starts to freeze. That happens at a
temperature of about 250C.

Now, it is the next bit that needs careful thinking about, because there
are two different ways you can look at it. If you have been taught to do it
one way, then stick with that - otherwise you risk getting very confused!

Thinking about changes in the composition of the liquid


When the first of the lead freezes, the composition of the remaining liquid
changes. It obviously becomes proportionally richer in tin. That lowers
the freezing point of the lead a bit more, and so the next bit of lead
freezes at a slightly lower temperature - leaving a liquid still richer in tin.
This process goes on. The liquid gets richer and richer in tin, and the
temperature needed to freeze the next lot of lead continues to fall. The
http://www.chemguide.co.uk/physical/phaseeqia/snpb.html (10 of 16)30/12/2004 11:31:53

solid-liquid phase diagrams: tin and lead

set of conditions of temperature and composition of the liquid essentially


moves down the curve - until it reaches the eutectic point.

Once it has reached the eutectic point, if the temperature continues to


fall, you obviously just move into the region of a mixture of solid lead and
solid tin - in other words, all the remaining liquid freezes.

Thinking about the composition of the system as a whole


We've seen that as the liquid gradually freezes, its composition changes.
But if you look at the system as a whole, obviously the proportions of
lead and tin remain constant - you aren't taking anything away or adding
anything. All that is happening is that things are changing from liquids to
solids.
So suppose we continue the cooling beyond the temperature that the first
solid lead appears and the temperature drops to the point shown in the
next diagram - a point clearly in the "solid lead and molten mixture" area.

http://www.chemguide.co.uk/physical/phaseeqia/snpb.html (11 of 16)30/12/2004 11:31:53

solid-liquid phase diagrams: tin and lead

What would you see in the mixture? To find out, you draw a horizontal tie
line through that point, and then look at the ends of it.

http://www.chemguide.co.uk/physical/phaseeqia/snpb.html (12 of 16)30/12/2004 11:31:53

solid-liquid phase diagrams: tin and lead

At the left-hand end, you have 100% lead. That represents the solid lead
that has frozen from the mixture. At the right-hand end, you have the
composition of the liquid mixture. This is now much richer in tin than the
whole system is - because obviously a fair bit of solid lead has separated
out.
As the temperature continues to fall, the composition of the liquid mixture
(as shown by the right-hand end of the tie line) will get closer and closer
to the eutectic mixture.

http://www.chemguide.co.uk/physical/phaseeqia/snpb.html (13 of 16)30/12/2004 11:31:53

solid-liquid phase diagrams: tin and lead

It will finally reach the eutectic composition when the temperature drops
to the eutectic temperature - and the whole lot then freezes.
At a temperature lower than the eutectic temperature, you are obviously
in the solid lead plus solid tin region. That's fairly obvious!

If you cooled a liquid mixture on the right-hand side of the phase diagram
(to the right of the eutectic mixture), everything would work exactly the
same except that solid tin would be formed instead of solid lead. If you
have understood what has gone before, it isn't at all difficult to work out
what happens.
Finally . . . what happens if you cool a liquid mixture which has exactly
the eutectic composition?
It simply stays as a liquid mixture until the temperature falls enough that
it all solidifies. You never get into the awkward areas of the phase
diagram.
http://www.chemguide.co.uk/physical/phaseeqia/snpb.html (14 of 16)30/12/2004 11:31:53

solid-liquid phase diagrams: tin and lead

Tin-lead mixtures as solder


Traditionally, tin-lead mixtures have been used as solder, but these are
being phased out because of health concerns over the lead. This is
especially the case where the solder is used to join water pipes where
the water is used for drinking. New non-lead solders have been
developed as safer replacements.
Typical old-fashioned solders include:

60% tin and 40% lead. This is close to the eutectic composition
(62% tin and 38% lead), giving a low melting point. It will also melt
and freeze cleanly over a very limited temperature range. This is
useful for electrical work.
50% tin and 50% lead. This will melt and freeze over a wider range
of temperatures. When it is molten it will start to freeze at about
220C and finally solidify at the eutectic temperature of 183C.
That means that it stays workable for a useful amount of time.
That's helpful if it is being used for plumbing joints.

Where would you like to go now?


To the phase equilibrium menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/physical/phaseeqia/snpb.html (15 of 16)30/12/2004 11:31:53

solid-liquid phase diagrams: tin and lead

http://www.chemguide.co.uk/physical/phaseeqia/snpb.html (16 of 16)30/12/2004 11:31:53

orders of reaction and mechanisms

ORDERS OF REACTION AND MECHANISMS

This page looks at the relationship between orders of reaction and


mechanisms in some simple cases. It explores what a mechanism is,
and the idea of a rate determining step. It also explains the difference
between the sometimes confusing terms "order of reaction" and
"molecularity of reaction".

Note: If you aren't sure about orders of reaction you ought to


read the introductory page before you go on. You will find a
link back to here at the bottom of the introductory page.

Reaction mechanisms
What is a reaction mechanism?
In any chemical change, some bonds are broken and new ones are
made. Quite often, these changes are too complicated to happen in one
simple stage. Instead, the reaction may involve a series of small changes
one after the other.
A reaction mechanism describes the one or more steps involved in the
reaction in a way which makes it clear exactly how the various bonds are
broken and made. The following example comes from organic chemistry.
It doesn't matter in the least if it is unfamiliar to you!
This is a reaction between 2-bromo-2-methylpropane and the hydroxide
ions from sodium hydroxide solution:

The overall reaction replaces the bromine atom in the organic compound
by an OH group.
The first thing that happens is that the carbon-bromine bond in a small
http://www.chemguide.co.uk/physical/basicrates/ordermech.html (1 of 8)30/12/2004 11:31:59

orders of reaction and mechanisms

proportion of the organic compound breaks to give ions:

Carbon-bromine bonds are reasonably strong, so this is a slow change. If


the ions hit each other again, the covalent bond will reform. The curly
arrow in the equation shows the movement of a pair of electrons.
If there is a high concentration of hydroxide ions present, the positive ion
stands a high chance of hitting one of those. This step of the overall
reaction will be very fast. A new covalent bond is made between the
carbon and the oxygen, using one of the lone pairs on the oxygen atom.

Because carbon-oxygen bonds are strong, once the OH group has


attached to the carbon atom, it tends to stay attached.
The mechanism shows that the reaction takes place in two steps and
describes exactly how those steps happen in terms of bonds being
broken or made. It also shows that the steps have different rates of
reaction - one slow and one fast.

Note: If you are interested in exploring more organic


reaction mechanisms you will find probably more than you
will want to know about by following this link!

http://www.chemguide.co.uk/physical/basicrates/ordermech.html (2 of 8)30/12/2004 11:31:59

orders of reaction and mechanisms

The rate determining step


The overall rate of a reaction (the one which you would measure if you
did some experiments) is controlled by the rate of the slowest step. In the
example above, the hydroxide ion can't combine with the positive ion
until that positive ion has been formed. The second step is in a sense
waiting around for the first slow step to happen.
The slow step of a reaction is known as the rate determining step.
As long as there is a lot of difference between the rates of the various
steps, when you measure the rate of a reaction, you are actually
measuring the rate of the rate determining step.

Reaction mechanisms and orders of reaction


The examples we use at this level are the very simple ones where the
orders of reaction with respect to the various substances taking part are
0, 1 or 2. These tend to have the slow step of the reaction happening
before any fast step(s).
To try to explain how fractional orders of reaction can arise is beyond the
scope of UK A' level courses.
Example 1
Here is the mechanism we have already looked at. How do we know that
it works like this?

http://www.chemguide.co.uk/physical/basicrates/ordermech.html (3 of 8)30/12/2004 11:31:59

orders of reaction and mechanisms

By doing rate of reaction experiments, you find this rate equation:

The reaction is first order with respect to the organic compound, and zero
order with respect to the hydroxide ions. The concentration of the
hydroxide ions isn't affecting the overall rate of the reaction.
If the hydroxide ions were taking part in the slow step of the reaction,
increasing their concentration would speed the reaction up. Since their
concentration doesn't seem to matter, they must be taking part in a later
fast step.
Increasing the concentration of the hydroxide ions will speed up the fast
step, but that won't have a noticeable effect on the overall rate of the
reaction. That is governed by the speed of the slow step.

Note: If you decreased the concentration of hydroxide ions


enough, you will eventually slow down the second step of this
reaction to the point where both steps have similar rates. At
that point, the concentration of the hydroxide ions will matter.
At normal concentrations, the rates of the two steps differ
widely, and so this problem doesn't arise.

In a simple case like this, where the slow step of the reaction is the first
step, the rate equation tells you what is taking part in that slow step. In
this case, the reaction is first order with respect to the organic molecule and that's all.
This gives you a starting point for working out a possible mechanism.
Having come up with a mechanism, you would need to find more
evidence to confirm it. For example, in this case you might try to detect
the presence of the positive ion that is formed in the first step.
Example 2

http://www.chemguide.co.uk/physical/basicrates/ordermech.html (4 of 8)30/12/2004 11:31:59

orders of reaction and mechanisms

At first sight this reaction seems identical with the last one. A bromine
atom is being replaced by an OH group in an organic compound.

However, the rate equation for this apparently similar reaction turns out
to be quite different. That means that the mechanism must be different.

The reaction this time is first order with respect to both the organic
compound and the hydroxide ions. Both of these must be taking part in
the slow step of the reaction. The reaction must happen by a
straightforward collision between them.

The carbon atom which is hit by the hydroxide ion has a slight positive
charge on it and the bromine a slight negative one because of the
difference in their electronegativities.
As the hydroxide ion approaches, the bromine is pushed off in one
smooth action.

Note: If you are interested in understanding these


mechanisms in more detail, you could follow this link. For the
purposes of this page, all that matters is that the rate
equations show that the two apparently similar reactions
happen by different mechanisms.

http://www.chemguide.co.uk/physical/basicrates/ordermech.html (5 of 8)30/12/2004 11:31:59

orders of reaction and mechanisms

Molecularity of a reaction
You may come across an older term known as the molecularity of a
reaction. This has largely dropped out of UK A' level syllabuses, but if
you meet it, it is important that you understand the difference between
this and the order of a reaction. The terms were sometimes used
carelessly as if they mean the same thing - they don't!
Order of reaction
The important thing to realise is that this is something which can only be
found by doing experiments. It gives you information about which
concentrations affect the rate of the reaction. You cannot look at an
equation for a reaction and deduce what the order of the reaction is
going to be - you have to do some practical work!
Having found the order of the reaction experimentally, you may be able
to make suggestions about the mechanism for the reaction - at least in
simple cases.
Molecularity of a reaction
This starts at the other end! If you know the mechanism for a reaction,
you can write down equations for a series of steps which make it up.
Each of those steps has a molecularity.
The molecularity of a step simply counts the number of species
(molecules, ions, atoms or free radicals) taking part in that step. For
example, going back to the mechanisms we've been looking at:

This step involves a single molecule breaking into ions. Because only
one species is involved in the reaction, it has a molecularity of 1. It could
be described as unimolecular.
The second step of this mechanism, involves two ions reacting together.

http://www.chemguide.co.uk/physical/basicrates/ordermech.html (6 of 8)30/12/2004 11:31:59

orders of reaction and mechanisms

This step has a molecularity of 2 - a bimolecular reaction.


The other reaction we looked at happened in a single step:

Because of the two species involved (one molecule and one ion), this
reaction is also bimolecular.
Unless an overall reaction happens in one step (like this last one), you
can't assign it a molecularity. You have to know the mechanism, and
then each individual step has its own molecularity.
There's nothing the least bit complicated about the term molecularity.
The only confusion is that you may sometimes find it used as if it meant
the same as order. It doesn't!

Where would you like to go now?


To the rates of reaction menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

http://www.chemguide.co.uk/physical/basicrates/ordermech.html (7 of 8)30/12/2004 11:31:59

orders of reaction and mechanisms

Jim Clark 2002

http://www.chemguide.co.uk/physical/basicrates/ordermech.html (8 of 8)30/12/2004 11:31:59

complex ions - names

NAMING COMPLEX METAL IONS

This page explains how to name some common complex metal ions.

How to work out the names of complex ions


Although the names of complex ions can look long and worrying, the
formulae are simply being coded in much the same way that organic
names are coded. Once you have sorted out that code, the names are
entirely descriptive.

Naming the ligands


Coding for the ligand
The table shows some common ligands and the code for them in the
name of a complex ion. The old names sometimes differ by a letter or so,
but never enough for it to be confusing.

ligand

coded by

(old name)

H2O

aqua

aquo

NH3

ammine

ammino

OH-

hydroxo

hydroxy

Cl-

chloro

F-

fluoro

http://www.chemguide.co.uk/inorganic/complexions/names.html (1 of 5)30/12/2004 11:32:04

complex ions - names

CN-

cyano

Take care with the code for ammonia as a ligand - it has 2 "m"s in its
name. If you miss one of these out so that you are left with "amine" or
"amino", you are refering to the NH2 group in an organic compound.
This is probably the only point of confusion with these names.
Coding for the number of ligands
The normal prefixes apply if there is more than one ligand.

no of ligands

coded by

di

tri

tetra

penta

hexa

Putting this together


For a complex ion containing only one type of ligand, there is no
problem. For example:
[Cu(H2O)6]2+ is called the hexaaquacopper(II) ion.
(Don't worry about the copper(II) bit for the moment.) The fact that there
are two "a"s next to each other in the name is OK.
With more than one type of ligand in an ion, the ligands are named in
alphabetical order - ignoring the prefixes. For example:
http://www.chemguide.co.uk/inorganic/complexions/names.html (2 of 5)30/12/2004 11:32:04

complex ions - names

[Cu(NH3)4(H2O)2]2+ is called the tetraamminediaquacopper(II) ion.


The "ammine" is named before the "aqua" because "am" comes before
"aq" in the alphabet. The "tetra" and "di" are ignored.

Naming the metal


You might have thought that this was fairly obvious, but it isn't
necessarily. It depends on whether the complex ion ends up as positively
or negatively charged.
For positively charged complex ions
A positively charged complex ion is called a cationic complex. A cation
is a positively charged ion.
The metal in this is named exactly as you would expect, with the addition
of its oxidation state.

Note: If you aren't sure about oxidation states, you could


follow this link. On the other hand, for the purposes of this
page, the oxidation state is simply the charge on the original
metal ion before it was complexed.
If you follow this link, use the BACK button on your browser
to return quickly to this page.

http://www.chemguide.co.uk/inorganic/complexions/names.html (3 of 5)30/12/2004 11:32:04

complex ions - names

Going back to a previous example, [Cu(H2O)6]2+ is called the


hexaaquacopper(II) ion because the copper's oxidation state is +2.
Copper's oxidation is +2 because the original uncomplexed ion was Cu2+
- NOT because the complex carries 2+ charges.
The oxidation state is frequently left out if a metal only ever has one
oxidation state. For example, in its compounds aluminium always has an
oxidation state of +3. [Al(H2O)6]3+ is usually just called the
hexaaquaaluminium ion rather than the hexaaquaaluminium(III) ion.
For negatively charged complex ions
A negatively charged complex ion is called an anionic complex. An
anion is a negatively charged ion.
In this case the name of the metal is modified to show that it has ended
up in a negative ion. This is shown by the ending -ate.
With many metals, the basic name of the metal is changed as well sometimes drastically!
Common examples include:

metal

changed to

cobalt

cobaltate

aluminium

aluminate

chromium

chromate

vanadium

vanadate

copper

cuprate

iron

ferrate

http://www.chemguide.co.uk/inorganic/complexions/names.html (4 of 5)30/12/2004 11:32:04

complex ions - names

So, for example, suppose you bond 4 chloride ions around a Cu2+ ion to
give [CuCl4]2-.
The name shows the 4 (tetra) chlorines (chloro) around a copper in an
overall negative ion (cuprate). The copper has on oxidation state of +2.
This is the tetrachlorocuprate(II) ion.
Similarly, [Al(H2O)2(OH)4]- is called the diaquatetrahydroxoaluminate ion.
Take the name to pieces so that you can see exactly what refers to what.
Don't forget that the two different ligands are named in alphabetical order
- aqua before hydroxo - ignoring the prefixes, di and tetra.
The oxidation state of the aluminium could be shown, but isn't absolutely
necessary because aluminium only has the one oxidation state in its
compounds. The full name is the diaquatetrahydroxoaluminate(III) ion.

Where would you like to go now?


To the complex ion menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/inorganic/complexions/names.html (5 of 5)30/12/2004 11:32:04

complex ions - shapes

THE SHAPES OF COMPLEX METAL IONS

This page describes the shapes of some common complex metal ions. It
goes on to look at some simple examples of stereoisomerism (geometric
and optical) in complex ions. If you aren't sure what stereoisomerism is,
you will find a helpful link further down the page.

Some simple shapes for complex ions


These shapes are for complex ions formed using monodentate ligands ligands which only form one bond to the central metal ion.
You will probably be familiar with working out the shapes of simple
compounds using the electron pair repulsion theory. Unfortunately that
doesn't work for most complex metal ions involving transition metals. The
answer is just to learn the shapes you need to know about. As you will
see, it isn't difficult.

Note: If you are interested (although it is completely


irrelevant to this page), you will find an explanation of how
you work out the shapes of simple molecules and ions by
following this link.
You don't need to know about this to understand the rest of
this page!

http://www.chemguide.co.uk/inorganic/complexions/shapes.html (1 of 8)30/12/2004 11:32:10

complex ions - shapes

6-co-ordinated complex ions


These are complex ions in which the central metal ion is forming six
bonds. In the simple cases we are talking about, that means that it will be
attached to six ligands.
These ions have an octahedral shape. Four of the ligands are in one
plane, with the fifth one above the plane, and the sixth one below the
plane.
The diagram shows four fairly random examples of octahedral ions.

http://www.chemguide.co.uk/inorganic/complexions/shapes.html (2 of 8)30/12/2004 11:32:10

complex ions - shapes

Note: Remember that the ligands attached to a wedge


shaped arrow are coming out of the screen or paper towards
you. Those attached to a dotted arrow are behind the plane
of the screen or paper. The two ligands attached to the
ordinary arrows are above and below the plane of the rest.

It doesn't matter what the ligands are. If you have six of them, this is the
shape they will take up. Easy!

4-co-ordinated complex ions


These are far less common, and they can take up one of two different
shapes.
Tetrahedral ions
These are the ones you are most likely to need for A' level purposes in
the UK. There are two very similar ions which crop up commonly at this
level: [CuCl4]2- and [CoCl4]2-.
The copper(II) and cobalt(II) ions have four chloride ions bonded to them
rather than six, because the chloride ions are too big to fit any more
around the central metal ion.

http://www.chemguide.co.uk/inorganic/complexions/shapes.html (3 of 8)30/12/2004 11:32:10

complex ions - shapes

That's not very difficult to remember either!

A square planar complex


Occasionally a 4-co-ordinated complex turns out to be square planar.
There's no easy way of predicting that this is going to happen. The only
one you might possibly come across at this level is cisplatin which is
used as an anti-cancer drug.
Cisplatin is a neutral complex, Pt(NH3)2Cl2. It is neutral because the 2+
charge of the original platinum(II) ion is exactly cancelled by the two
negative charges supplied by the chloride ions.

The platinum, the two chlorines, and the two nitrogens are all in the same
plane. We will have more to say about cisplatin immediately below.

Stereoisomerism in complex ions


Some complex ions can show either optical or geometric isomerism.

http://www.chemguide.co.uk/inorganic/complexions/shapes.html (4 of 8)30/12/2004 11:32:10

complex ions - shapes

Warning! If you don't know about optical and geometric


isomers, it is essential that you explore the isomerism menu
in the organic chemistry section of this site before you go any
further. The rest of this page is unlikely to mean anything to
you if you don't have this background!
Use the BACK button, or the History file, or the Go menu on
your browser to return to this page - depending on how
waylaid you get!

Geometric isomerism
This occurs in planar complexes like the Pt(NH3)2Cl2 we've just looked
at. There are two completely different ways in which the ammonias and
chloride ions could arrange themselves around the central platinum ion:

The two structures drawn are isomers because there is no way that you
can just twist one to turn it into the other. The complexes are both locked
into their current forms.
The termscis and trans are used in the same way as they are in organic
chemistry. Trans implies "opposite" - notice that the ammonias are
arranged opposite each other in that version, and so are the chlorines.
Cis implies "on the same side" - in this instance, that just means that the
ammonias and the chlorines are next door to each other.

http://www.chemguide.co.uk/inorganic/complexions/shapes.html (5 of 8)30/12/2004 11:32:10

complex ions - shapes

Optical isomerism
You recognise optical isomers because they have no plane of symmetry.
In the organic case, it is fairly easy to recognise the possibiliy of this by
looking for a carbon atom with four different things attached to it. It isn't
qute so easy with the complex ions - either to draw or to visualise!
The examples you are most likely to need occur in octahedral complexes
which contain bidentate ligands - ions like [Ni(NH2CH2CH2NH2)3]2+ or [Cr
(C2O4)3]3-.

Help! If these ions look scarily unfamiliar, you must read the
introductory page on complex ions before you go on. The
shapes of the ions given below will be simplified in a way that
is explained on that page.
Use the BACK button on your browser to return to this page.

The diagram below shows a simplified view of one of these ions.


Essentially, they all have the same shape - all that differs is the nature of
the "headphones".
I have deliberately left the charges off the ion, because obviously they
will vary from case to case. The shape shown applies to any ion of this
kind.

http://www.chemguide.co.uk/inorganic/complexions/shapes.html (6 of 8)30/12/2004 11:32:10

complex ions - shapes

If your visual imagination will cope, you may be able to see that this ion
has no plane of symmetry. If you find this difficult to visualise, the only
solution is to make the ion out of a lump of plasticene (or a bit of clay or
dough) and three bits of cardboard cut to shape.
A substance with no plane of symmetry is going to have optical isomers one of which is the mirror image of the other. One of the isomers will
rotate the plane of polarisation of plane polarised light clockwise; the
other rotates it anti-clockwise.

Help! If you don't understand what I am talking about, it is


because you didn't take my advice and follow the link to the
isomerism part of the site earlier on this page! You need to
read about optical isomerism in the organic chemistry section
of this site.
Use the BACK button on your browser to return to this page .

In this case, the two isomers are:

If you have a really impressive visual imagination, you may be able to


see that there is no way of rotating the second isomer in space so that it
looks exactly the same as the first one. I can't do this! The only way I
have ever been able to convince myself that they are different is to make
models.
http://www.chemguide.co.uk/inorganic/complexions/shapes.html (7 of 8)30/12/2004 11:32:10

complex ions - shapes

For exam purposes, this doesn't matter in the slightest. As long as you
draw the isomers carefully, with the second one a true reflection of the
first, the two structures will be different.

Where would you like to go now?


To the complex ion menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/inorganic/complexions/shapes.html (8 of 8)30/12/2004 11:32:10

shapes of molecules and ions containing single bonds

SHAPES OF MOLECULES AND IONS

This page explains how to work out the shapes of molecules and ions
containing only single bonds. If you are interested in the shapes of
molecules and ions containing double bonds, you will find a link at the
bottom of the page.

The electron pair repulsion theory


The shape of a molecule or ion is governed by the arrangement of the
electron pairs around the central atom. All you need to do is to work out
how many electron pairs there are at the bonding level, and then arrange
them to produce the minimum amount of repulsion between them. You
have to include both bonding pairs and lone pairs.
How to work out the number of electron pairs
You can do this by drawing dots-and-crosses pictures, or by working out
the structures of the atoms using electrons-in-boxes and worrying about
promotion, hybridisation and so on. But this is all very tedious! You can
get exactly the same information in a much quicker and easier way for
the examples you will meet at A'level.

Warning: This method won't work without some modification


for many ions containing metals, and no simple method gives
reliable results where the central atom is a transition metal.
The method will, however, cope with all the substances that
you are likely to meet in this section of the syllabus. When
you deal with transition metal chemistry, you will be expected
to know the shapes of some ions formed by transition metals,
but not to work them out. At that point, learn the ones your
syllabus wants you to know.
It is important to know exactly which molecules and ions your
syllabus expects you to be able to work out the shapes for in
this part of the syllabus. You should also check past exam
http://www.chemguide.co.uk/atoms/bonding/shapes.html (1 of 12)30/12/2004 11:32:26

shapes of molecules and ions containing single bonds

papers. If you haven't got copies of your syllabus and past


papers follow this link to find out how to get them.

First you need to work out how many electrons there are around the
central atom:

Write down the number of electrons in the outer level of the central
atom. That will be the same as the Periodic Table group number,
except in the case of the noble gases which form compounds,
when it will be 8.
Add one electron for each bond being formed. (This allows for the
electrons coming from the other atoms.)
Allow for any ion charge. For example, if the ion has a 1- charge,
add one more electron. For a 1+ charge, deduct an electron.

Now work out how many bonding pairs and lone pairs of electrons there
are:

Divide by 2 to find the total number of electron pairs around the


central atom.
Work out how many of these are bonding pairs, and how many are
lone pairs. You know how many bonding pairs there are because
you know how many other atoms are joined to the central atom
(assuming that only single bonds are formed).
For example, if you have 4 pairs of electrons but only 3 bonds,
there must be 1 lone pair as well as the 3 bonding pairs.

Finally, you have to use this information to work out the shape:

Arrange these electron pairs in space to minimise repulsions. How


this is done will become clear in the examples which follow.

http://www.chemguide.co.uk/atoms/bonding/shapes.html (2 of 12)30/12/2004 11:32:26

shapes of molecules and ions containing single bonds

Don't panic! This is all much easier to do in practice than it


is to describe in a long list like this one!

Two electron pairs around the central atom


The only simple case of this is beryllium chloride, BeCl2. The
electronegativity difference between beryllium and chlorine isn't enough
to allow the formation of ions.
Beryllium has 2 outer electrons because it is in group 2. It forms bonds to
two chlorines, each of which adds another electron to the outer level of
the beryllium. There is no ionic charge to worry about, so there are 4
electrons altogether - 2 pairs.
It is forming 2 bonds so there are no lone pairs. The two bonding pairs
arrange themselves at 180 to each other, because that's as far apart as
they can get. The molecule is described as being linear.

Three electron pairs around the central atom


The simple cases of this would be BF3 or BCl3.
Boron is in group 3, so starts off with 3 electrons. It is forming 3 bonds,
adding another 3 electrons. There is no charge, so the total is 6 electrons
- in 3 pairs.
Because it is forming 3 bonds there can be no lone pairs. The 3 pairs
arrange themselves as far apart as possible. They all lie in one plane at
120 to each other. The arrangement is called trigonal planar.

http://www.chemguide.co.uk/atoms/bonding/shapes.html (3 of 12)30/12/2004 11:32:26

shapes of molecules and ions containing single bonds

In the diagram, the other electrons on the fluorines have been left out
because they are irrelevant.

Four electron pairs around the central atom


There are lots of examples of this. The simplest is methane, CH4.

Note: Elsewhere on the site, you will find the shape of


methane worked out in detail using modern bonding theory.
Here we are doing it the quick and easy way!
If you are interested in the bonding in methane you can find it
in the organic section by following this link, or in a page on
covalent bonding by following this one.

Carbon is in group 4, and so has 4 outer electrons. It is forming 4 bonds


to hydrogens, adding another 4 electrons - 8 altogether, in 4 pairs.
Because it is forming 4 bonds, these must all be bonding pairs.
Four electron pairs arrange themselves in space in what is called a
tetrahedral arrangement. A tetrahedron is a regular triangularly-based
pyramid. The carbon atom would be at the centre and the hydrogens at
the four corners. All the bond angles are 109.5.

http://www.chemguide.co.uk/atoms/bonding/shapes.html (4 of 12)30/12/2004 11:32:26

shapes of molecules and ions containing single bonds

Note: It is important that you understand the use of various


sorts of line to show the 3-dimensional arrangement of the
bonds. In diagrams of this sort, an ordinary line represents a
bond in the plane of the screen or paper. A dotted line shows
a bond going away from you into the screen or paper. A
wedge shows a bond coming out towards you.

Other examples with four electron pairs around the central atom
Ammonia, NH3
Nitrogen is in group 5 and so has 5 outer electrons. Each of the 3
hydrogens is adding another electron to the nitrogen's outer level,
making a total of 8 electrons in 4 pairs. Because the nitrogen is only
forming 3 bonds, one of the pairs must be a lone pair. The electron pairs
arrange themselves in a tetrahedral fashion as in methane.

In this case, an additional factor comes into play. Lone pairs are in
orbitals that are shorter and rounder than the orbitals that the bonding
pairs occupy. Because of this, there is more repulsion between a lone
pair and a bonding pair than there is between two bonding pairs.
That forces the bonding pairs together slightly - reducing the bond angle
from 109.5 to 107. It's not much, but the examiners will expect you to
http://www.chemguide.co.uk/atoms/bonding/shapes.html (5 of 12)30/12/2004 11:32:26

shapes of molecules and ions containing single bonds

know it.
Remember this:
Greatest repulsion

lone pair - lone pair


lone pair - bond pair

Least repulsion

bond pair - bond pair

Be very careful when you describe the shape of ammonia. Although the
electron pair arrangement is tetrahedral, when you describe the shape,
you only take notice of the atoms. Ammonia is pyramidal - like a
pyramid with the three hydrogens at the base and the nitrogen at the top.

Water, H2O

Following the same logic as before, you will find that the oxygen has four
pairs of electrons, two of which are lone pairs. These will again take up a
tetrahedral arrangement. This time the bond angle closes slightly more to
104, because of the repulsion of the two lone pairs.
The shape isn't described as tetrahedral, because we only "see" the
oxygen and the hydrogens - not the lone pairs. Water is described as
bent or V-shaped.

The ammonium ion, NH4+


The nitrogen has 5 outer electrons, plus another 4 from the four
hydrogens - making a total of 9.

http://www.chemguide.co.uk/atoms/bonding/shapes.html (6 of 12)30/12/2004 11:32:26

shapes of molecules and ions containing single bonds

But take care! This is a positive ion. It has a 1+ charge because it has
lost 1 electron. That leaves a total of 8 electrons in the outer level of the
nitrogen. There are therefore 4 pairs, all of which are bonding because of
the four hydrogens.
The ammonium ion has exactly the same shape as methane, because it
has exactly the same electronic arrangement. NH4+ is tetrahedral.

Note: To simplify diagrams, bonding electrons won't be


shown from now on. Each line, of course, represents a
bonding pair. It is essential, however, to draw lone pairs.

Methane and the ammonium ion are said to be isoelectronic. Two


species (atoms, molecules or ions) are isoelectronic if they have exactly
the same number and arrangement of electrons (including the distinction
between bonding pairs and lone pairs).

The hydroxonium ion, H3O+


Oxygen is in group 6 - so has 6 outer electrons. Add 1 for each
hydrogen, giving 9. Take one off for the +1 ion, leaving 8. This gives 4
pairs, 3 of which are bond pairs. The hydroxonium ion is isoelectronic
with ammonia, and has an identical shape - pyramidal.

http://www.chemguide.co.uk/atoms/bonding/shapes.html (7 of 12)30/12/2004 11:32:26

shapes of molecules and ions containing single bonds

Five electron pairs around the central atom


A simple example: phosphorus(V) fluoride, PF5
(The argument for phosphorus(V) chloride, PCl5, would be identical.)
Phosphorus (in group 5) contributes 5 electrons, and the five fluorines 5
more, giving 10 electrons in 5 pairs around the central atom. Since the
phosphorus is forming five bonds, there can't be any lone pairs.
The 5 electron pairs take up a shape described as a trigonal bipyramid
- three of the fluorines are in a plane at 120 to each other; the other two
are at right angles to this plane. The trigonal bipyramid therefore has two
different bond angles - 120 and 90.

A tricky example, ClF3


Chlorine is in group 7 and so has 7 outer electrons. The three fluorines
contribute one electron each, making a total of 10 - in 5 pairs. The
chlorine is forming three bonds - leaving you with 3 bonding pairs and 2
lone pairs, which will arrange themselves into a trigonal bipyramid.

http://www.chemguide.co.uk/atoms/bonding/shapes.html (8 of 12)30/12/2004 11:32:26

shapes of molecules and ions containing single bonds

But don't jump to conclusions. There are actually three different ways in
which you could arrange 3 bonding pairs and 2 lone pairs into a trigonal
bipyramid. The right arrangement will be the one with the minimum
amount of repulsion - and you can't decide that without first drawing all
the possibilities.

These are the only possible arrangements. Anything else you might think
of is simply one of these rotated in space.
We need to work out which of these arrangements has the minimum
amount of repulsion between the various electron pairs.
A new rule applies in cases like this:

If you have more than four electron pairs arranged


around the central atom, you can ignore repulsions at
angles of greater than 90.

One of these structures has a fairly obvious large amount of repulsion.

In this diagram, two lone pairs are at 90 to each other, whereas in the
other two cases they are at more than 90, and so their repulsions can
be ignored. ClF3 certainly won't take up this shape because of the strong
lone pair-lone pair repulsion.
To choose between the other two, you need to count up each sort of
http://www.chemguide.co.uk/atoms/bonding/shapes.html (9 of 12)30/12/2004 11:32:26

shapes of molecules and ions containing single bonds

repulsion.
In the next structure, each lone pair is at 90 to 3 bond pairs, and so
each lone pair is responsible for 3 lone pair-bond pair repulsions.

Because of the two lone pairs there are therefore 6 lone pair-bond pair
repulsions. And that's all. The bond pairs are at an angle of 120 to each
other, and their repulsions can be ignored.
Now consider the final structure.

Each lone pair is at 90 to 2 bond pairs - the ones above and below the
plane. That makes a total of 4 lone pair-bond pair repulsions - compared
with 6 of these relatively strong repulsions in the last structure. The other
fluorine (the one in the plane) is 120 away, and feels negligible repulsion
from the lone pairs.
The bond to the fluorine in the plane is at 90 to the bonds above and
below the plane, so there are a total of 2 bond pair-bond pair repulsions.
The structure with the minimum amount of repulsion is therefore this last
one, because bond pair-bond pair repulsion is less than lone pair-bond
pair repulsion. ClF3 is described as T-shaped.

http://www.chemguide.co.uk/atoms/bonding/shapes.html (10 of 12)30/12/2004 11:32:26

shapes of molecules and ions containing single bonds

Warning! If your syllabus expects you to discuss examples


with more than 4 pairs of electrons around the central atom,
check past exam papers to see if nasty questions like this
one involving ClF3 ever come up. If so, don't leave this
example until you are sure that you understand it. It is by far
the most complicated one on this page.

Six electron pairs around the central atom


A simple example: SF6
6 electrons in the outer level of the sulphur, plus 1 each from the six
fluorines, makes a total of 12 - in 6 pairs. Because the sulphur is forming
6 bonds, these are all bond pairs. They arrange themselves entirely at
90, in a shape described as octahedral.

Two slightly more difficult examples


XeF4
Xenon forms a range of compounds, mainly with fluorine or oxygen, and
this is a typical one. Xenon has 8 outer electrons, plus 1 from each
fluorine - making 12 altogether, in 6 pairs. There will be 4 bonding pairs
(because of the four fluorines) and 2 lone pairs.

http://www.chemguide.co.uk/atoms/bonding/shapes.html (11 of 12)30/12/2004 11:32:26

shapes of molecules and ions containing single bonds

There are two possible structures, but in one of them the lone pairs
would be at 90. Instead, they go opposite each other. XeF4 is described
as square planar.

ClF4Chlorine is in group 7 and so has 7 outer electrons. Plus the 4 from the
four fluorines. Plus one because it has a 1- charge. That gives a total of
12 electrons in 6 pairs - 4 bond pairs and 2 lone pairs. The shape will be
identical with that of XeF4.

Where would you like to go now?


To look at shapes involving double bonds . . .
To the bonding menu . . .
To the atomic structure and bonding menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/atoms/bonding/shapes.html (12 of 12)30/12/2004 11:32:26

shapes of molecules and ions containing double bonds

SHAPES OF MOLECULES AND IONS


CONTAINING DOUBLE BONDS

This page explains how to work out the shapes of molecules and ions
containing double bonds. It assumes that you have just read the page on
shapes of molecules and ions containing only single bonds.

Important! If you have come straight to this page via a


search engine, you should read about the shapes of more
simple molecules and ions before you go on.

The shapes of simple molecules containing double bonds


Carbon dioxide, CO2
Carbon dioxide could be drawn (without making any assumptions about
the shape) as

The carbon originally had 4 electrons in its outer level (group 4). Each
oxygen contributes 2 electrons - 1 for each bond. That means there are a
total of 8 electrons around the carbon, in 4 pairs. Because there are 4
bonds, these are all bond pairs.
Each double bond uses 2 bond pairs - which are then thought of as a
single unit. Those two double bond units will try to get as far apart as
possible, and so the molecule is linear. The structure we've drawn above
does in fact represent the shape of the molecule.

http://www.chemguide.co.uk/atoms/bonding/shapesdouble.html (1 of 7)30/12/2004 11:32:31

shapes of molecules and ions containing double bonds

Sulphur dioxide, SO2


Sulphur dioxide could be drawn exactly the same as carbon dioxide
(again without making any assumptions about the shape):

The argument develops differently though. Sulphur has 6 electrons in its


outer level, and the oxygens between them contribute another 4 (1 for
each bond). That gives 10 electrons in total - 5 pairs. 4 pairs are needed
for the bonds, leaving 1 lone pair. Each double bond uses 2 bond pairs
and can be thought of as a single unit.
There are 2 double bond units and 1 lone pair, which will try to get as far
apart as possible - taking up a trigonal planar arrangement.

Because the lone pair isn't counted when you describe the shape, SO2 is
described as bent or V-shaped. In this case, our original structure
misrepresented the shape.

The shapes of ions containing double bonds


This is much more tricky. Before you can do anything else, you need to
know exactly how the ion is bonded - which bonds are double, which
bonds are single, and where the charges are. Check your syllabus to find
out whether you really need to bother about all this.
Fortunately, there is a simplifying factor. In the three ions you might need
to know about, not one of them has a lone pair on the central atom. That
means that the shapes are all pretty obvious as long as you know the
bonding.

http://www.chemguide.co.uk/atoms/bonding/shapesdouble.html (2 of 7)30/12/2004 11:32:31

shapes of molecules and ions containing double bonds

The sulphate ion, SO42The sulphate ion is bonded like this:

The negative charges are found on two of the oxygen atoms.

Beyond A'level: In fact, all the four sulphur-oxygen bonds


are identical, and the two negative charges are spread out
over all four of the oxygens. The ion has some complicated
delocalisation.

The sulphur is in Group 6 and so has 6 outer electrons. It is forming six


bonds to the various oxygens, so that adds a further 6 electrons, making
12 in all. You don't have to add any electrons for the negative charges,
because those charges aren't found on the sulphur atom.
Because there are 6 pairs of electrons, and a total of 6 bonds, all the
pairs of electrons are bonding pairs - there aren't any lone pairs.
Each double bond accounts for 2 of those pairs, and can be thought of
as a single unit. The two double bond units, and the two single bonds
arrange themselves as far apart as possible to give a
tetrahedralarrangement.

http://www.chemguide.co.uk/atoms/bonding/shapesdouble.html (3 of 7)30/12/2004 11:32:31

shapes of molecules and ions containing double bonds

The carbonate ion, CO32The carbonate ion is bonded like this:

Like the sulphate ion, the negative charges are again found on two of the
oxygen atoms.

Beyond A'level: Again, this is a simplification. All the carbonoxygen bonds are identical, and the two negative charges are
spread over the three oxygens. The ion is again delocalised.

The carbon has 4 outer electrons, and the four bonds to the oxygens add
another 4 - making 8 in total. Once again, you don't need to add any
electrons for the negative charges because those charges aren't on the
carbon.
There are 4 pairs of electrons, and the carbon is forming 4 bonds - so
there aren't any lone pairs to worry about. Two of the pairs of electrons
make up a single unit - the double bond. The double bond unit and the
two single bonds arrange themselves as far apart as possible in a
trigonal planar arrangement.

http://www.chemguide.co.uk/atoms/bonding/shapesdouble.html (4 of 7)30/12/2004 11:32:31

shapes of molecules and ions containing double bonds

The nitrate ion, NO3The nitrate ion is distinctly awkward and there isn't any simple way of
working out its shape. The problem is that it contains a co-ordinate
(dative covalent) bond.

Note: If you are a bit hazy about co-ordinate (dative


covalent) bonding you could follow this link before you go on.
It probably isn't necessary - all you need to know is that a coordinate bond is one in which both electrons come from the
same atom.

The nitrate ion comes from nitric acid, so we'll start from the structure of
that.

Look carefully at the bonding around the nitrogen. You will see that one
of the bonds is formed entirely from the lone pair on the nitrogen. That's
the co-ordinate bond.

http://www.chemguide.co.uk/atoms/bonding/shapesdouble.html (5 of 7)30/12/2004 11:32:31

shapes of molecules and ions containing double bonds

The nitrate ion is formed by the loss of the hydrogen ion, and so its
structure is:

Around the central nitrogen there are 4 pairs of shared electrons, and no
remaining lone pair. The original lone pair has now become a bonding
pair. Two of those pairs make up a double bond. The double bond unit
and the two single bonds arrange themselves as far apart as possible in
a trigonal planar arrangement - exactly the same as the carbonate ion.

Note: Co-ordinate bonds can be shown using an arrow


going from the atom donating the lone pair towards the atom
receiving it. Once again, this structure is a simplification for
A'level purposes. In fact, all the nitrogen-oxygen bonds are
identical, and the charge is delocalised evenly over all three
oxygen atoms.

http://www.chemguide.co.uk/atoms/bonding/shapesdouble.html (6 of 7)30/12/2004 11:32:31

shapes of molecules and ions containing double bonds

Where would you like to go now?


To the bonding menu . . .
To the atomic structure and bonding menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/atoms/bonding/shapesdouble.html (7 of 7)30/12/2004 11:32:31

complex ions - colour

THE COLOURS OF COMPLEX METAL IONS

This page is going to take a simple look at the origin of colour in complex
ions - in particular, why so many transition metal ions are coloured. Be
aware that this is only an introduction to what can grow into an extremely
complicated topic.

Why do we see some compounds as being coloured?


White light
You will know, of course, that if you pass white light through a prism it
splits into all the colours of the rainbow. Visible light is simply a small part
of an electromagnetic spectrum most of which we can't see - gamma rays,
X-rays, infra-red, radio waves and so on.
Each of these has a particular wavelength, ranging from 10-16 metres for
gamma rays to several hundred metres for radio waves. Visible light has
wavelengths from about 400 to 750 nm. (1 nanometre = 10-9 metres.)
The diagram shows an approximation to the spectrum of visible light.

http://www.chemguide.co.uk/inorganic/complexions/colour.html (1 of 15)30/12/2004 11:32:41

complex ions - colour

Important: This isn't a real spectrum - it's a made-up drawing.


The colours are only an approximation, and so are the
wavelengths assigned to them. Anyone choosing to use this
spectrum as anything more than an illustration should be aware
that it lacks any pretence of accuracy!

Why is copper(II) sulphate solution blue?


If white light (ordinary sunlight, for example) passes through copper(II)
sulphate solution, some wavelengths in the light are absorbed by the
solution. Copper(II) ions in solution absorb light in the red region of the
spectrum.
The light which passes through the solution and out the other side will
have all the colours in it except for the red. We see this mixture of
wavelengths as pale blue (cyan).
The diagram gives an impression of what happens if you pass white light
through copper(II) sulphate solution.

Working out what colour you will see isn't easy if you try to do it by
imagining "mixing up" the remaining colours. You wouldn't have thought
that all the other colours apart from some red would look cyan, for example.
Sometimes what you actually see is quite unexpected. Mixing different
wavelengths of light doesn't give you the same result as mixing paints or
other pigments.
http://www.chemguide.co.uk/inorganic/complexions/colour.html (2 of 15)30/12/2004 11:32:41

complex ions - colour

You can, however, sometimes get some estimate of the colour you would
see using the idea of complementary colours.

Complementary colours
If you arrange some colours in a circle, you get a "colour wheel". The
diagram shows one possible version of this. An internet search (see below)
will throw up many different versions!

Colours directly opposite each other on the colour wheel are said to be
complementary colours. Blue and yellow are complementary colours; red
and cyan are complementary; and so are green and magenta.
Mixing together two complementary colours of light will give you white light.

Beware: That is NOT the same as mixing together paint


colours. If you mix yellow and blue paint you don't get white
paint. Is this confusing? YES!

http://www.chemguide.co.uk/inorganic/complexions/colour.html (3 of 15)30/12/2004 11:32:41

complex ions - colour

What this all means is that if a particular colour is absorbed from white
light, what your eye detects by mixing up all the other wavelengths of light
is its complementary colour. Copper(II) sulphate solution is pale blue
(cyan) because it absorbs light in the red region of the spectrum. Cyan is
the complementary colour of red.

Note: If you are interested in understanding the relationship


between colour absorbed and colour seen (beyond the very
basic description above), you ought to do a full Google (or
similar) search on colour theory or complementary colours,
looking at the difference between additive and subtractive
colours. There is a link to Google on the Main Menu of this site.
Don't spend time on this, though, unless you are prepared for a
lot of confusion and frustration!

The origin of colour in complex ions


Transition metal v other metal complex ions
What is a transition metal?
We often casually talk about the transition metals as being those in the
middle of the Periodic Table where d orbitals are being filled, but these
should really be called d block elements rather than transition elements
(or metals).

http://www.chemguide.co.uk/inorganic/complexions/colour.html (4 of 15)30/12/2004 11:32:41

complex ions - colour

Note: If you don't understand about the filling of orbitals in the


Periodic Table, then you must follow this link before you go on.
Use the BACK button on your browser to return to this page.

This shortened version of the Periodic Table shows the first row of the d
block, where the 3d orbitals are being filled.

The usual definition of a transition metal is one which forms one or more
stable ions which have incompletely filled d orbitals.

Note: One of the UK A' level syllabuses includes the possibility


of the metal itself having incompletely filled d levels. This is
unlikely to be a big problem, but it would pay you to learn the
version your syllabus wants.
If you haven't got a copy of your syllabus, follow this link to find
out how to get one. Use the BACK button on your browser to
return quickly to this page.

http://www.chemguide.co.uk/inorganic/complexions/colour.html (5 of 15)30/12/2004 11:32:41

complex ions - colour

Zinc with the electronic structure [Ar] 3d104s2 doesn't count as a transition
metal whichever definition you use. In the metal, it has a full 3d level.
When it forms an ion, the 4s electrons are lost - again leaving a completely
full 3d level.
At the other end of the row, scandium ( [Ar] 3d14s2 ) doesn't really counts
as a transition metal either. Although there is a partially filled d level in the
metal, when it forms its ion, it loses all three outer electrons.
The Sc3+ ion doesn't count as a transition metal ion because its 3d level is
empty.

Some sample colours


The diagrams show the approximate colours of some typical hexaaqua
metal ions, with the formula [ M(H2O)6 ] n+. The charge on these ions is
typically 2+ or 3+.

Note: If you aren't happy about naming complex ions, you


might find it useful to follow this link.
Use the BACK button on your browser to return to this page.

Non-transition metal ions

http://www.chemguide.co.uk/inorganic/complexions/colour.html (6 of 15)30/12/2004 11:32:41

complex ions - colour

These ions are all colourless. (Sorry, I can't do genuinely colourless!)


Transition metal ions

The corresponding transition metal ions are coloured. Some, like the
hexaaquamanganese(II) ion (not shown) and the hexaaquairon(II) ion, are
quite faintly coloured - but they are coloured.
So . . . what causes transition metal ions to absorb wavelengths from
visible light (causing colour) whereas non-transition metal ions don't? And
why does the colour vary so much from ion to ion?

The origin of colour in complex ions containing transition metals


Complex ions containing transition metals are usually coloured, whereas
the similar ions from non-transition metals aren't. That suggests that the
partly filled d orbitals must be involved in generating the colour in some
way. Remember that transition metals are defined as having partly filled d
orbitals.
Octahedral complexes
For simplicity we are going to look at the octahedral complexes which have
six simple ligands arranged around the central metal ion. The argument
isn't really any different if you have multidentate ligands - it's just slightly
more difficult to imagine!

http://www.chemguide.co.uk/inorganic/complexions/colour.html (7 of 15)30/12/2004 11:32:41

complex ions - colour

Note: If you aren't sure about the shapes of complex ions, you
might find it useful to follow this link before you go on. You only
need to read the beginning of that page.
If you don't know what a ligand is, you should read the
introduction to complex ions as a matter of urgency!
Use the BACK button on your browser to return to this page.

When the ligands bond with the transition metal ion, there is repulsion
between the electrons in the ligands and the electrons in the d orbitals of
the metal ion. That raises the energy of the d orbitals.
However, because of the way the d orbitals are arranged in space, it
doesn't raise all their energies by the same amount. Instead, it splits them
into two groups.
The diagram shows the arrangement of the d electrons in a Cu2+ ion
before and after six water molecules bond with it.

Whenever 6 ligands are arranged around a transition metal ion, the d


http://www.chemguide.co.uk/inorganic/complexions/colour.html (8 of 15)30/12/2004 11:32:41

complex ions - colour

orbitals are always split into 2 groups in this way - 2 with a higher energy
than the other 3.
The size of the energy gap between them (shown by the blue arrows on
the diagram) varies with the nature of the transition metal ion, its oxidation
state (whether it is 3+ or 2+, for example), and the nature of the ligands.
When white light is passed through a solution of this ion, some of the
energy in the light is used to promote an electron from the lower set of
orbitals into a space in the upper set.

Each wavelength of light has a particular energy associated with it. Red
light has the lowest energy in the visible region. Violet light has the
greatest energy.
Suppose that the energy gap in the d orbitals of the complex ion
corresponded to the energy of yellow light.

The yellow light would be absorbed because its energy would be used in
promoting the electron. That leaves the other colours.

http://www.chemguide.co.uk/inorganic/complexions/colour.html (9 of 15)30/12/2004 11:32:41

complex ions - colour

Your eye would see the light passing through as a dark blue, because blue
is the complementary colour of yellow.

Warning: This is a major simplification, but is adequate for this


level (UK A' level). It doesn't, for example, account for
absorption happening over a broad range of wavelengths
rather than a single one, or for cases where there is more than
one colour absorbed from different parts of the spectrum.

What about non-transition metal complex ions?


Non-transition metals don't have partly filled d orbitals. Visible light is only
absorbed if some energy from the light is used to promote an electron over
exactly the right energy gap. Non-transition metals don't have any electron
transitions which can absorb wavelengths from visible light.
For example, although scandium is a member of the d block, its ion (Sc3+)
hasn't got any d electrons left to move around. This is no different from an
ion based on Mg2+ or Al3+. Scandium(III) complexes are colourless
because no visible light is absorbed.
In the zinc case, the 3d level is completely full - there aren't any gaps to
promote an electron in to. Zinc complexes are also colourless.

Tetrahedral complexes
Simple tetrahedral complexes have four ligands arranged around the
central metal ion. Again the ligands have an effect on the energy of the d
electrons in the metal ion. This time, of course, the ligands are arranged
differently in space relative to the shapes of the d orbitals.
The net effect is that when the d orbitals split into two groups, three of
them have a greater energy, and the other two a lesser energy (the
opposite of the arrangement in an octahedral complex).
Apart from this difference of detail, the explanation for the origin of colour

http://www.chemguide.co.uk/inorganic/complexions/colour.html (10 of 15)30/12/2004 11:32:41

complex ions - colour

in terms of the absorption of particular wavelengths of light is exactly the


same as for octahedral complexes.

The factors affecting the colour of a transition metal complex ion


In each case we are going to choose a particular metal ion for the centre of
the complex, and change other factors. Colour changes in a fairly
haphazard way from metal to metal across a transition series.
The nature of the ligand
Different ligands have different effects on the energies of the d orbitals of
the central ion. Some ligands have strong electrical fields which cause a
large energy gap when the d orbitals split into two groups. Others have
much weaker fields producing much smaller gaps.
Remember that the size of the gap determines what wavelength of light is
going to get absorbed.
The list shows some common ligands. Those at the top produce the
smallest splitting; those at the bottom the largest splitting.

The greater the splitting, the more energy is needed to promote an


electron from the lower group of orbitals to the higher ones. In terms of the
colour of the light absorbed, greater energy corresponds to shorter
wavelengths.
That means that as the splitting increases, the light absorbed will tend to
http://www.chemguide.co.uk/inorganic/complexions/colour.html (11 of 15)30/12/2004 11:32:41

complex ions - colour

shift away from the red end of the spectrum towards orange, yellow and so
on.
There is a fairly clear-cut case in copper(II) chemistry.
If you add an excess of ammonia solution to hexaaquacopper(II) ions in
solution, the pale blue (cyan) colour is replaced by a dark inky blue as
some of the water molecules in the complex ion are replaced by ammonia.

The first complex must be absorbing red light in order to give the
complementary colour cyan. The second one must be absorbing in the
yellow region in order to give the complementary colour dark blue.
Yellow light has a higher energy than red light. You need that higher
energy because ammonia causes more splitting of the d orbitals than
water does.
It isn't often as simple to see as this, though! Trying to sort out what is
being absorbed when you have murky colours not on the simple colour
wheel further up the page is much more of a problem.
The diagrams show some approximate colours of some ions based on
chromium(III).

http://www.chemguide.co.uk/inorganic/complexions/colour.html (12 of 15)30/12/2004 11:32:41

complex ions - colour

It is obvious that changing the ligand is changing the colour, but trying to
explain the colours in terms of our simple theory isn't easy.

Note: To be honest, I spent a couple of weeks trying to find a


way of doing this simply, based on a simple colour wheel, and
eventually gave up. Life is too short!

The oxidation state of the metal


As the oxidation state of the metal increases, so also does the amount of
splitting of the d orbitals.
Changes of oxidation state therefore change the colour of the light
absorbed, and so the colour of the light you see.
Taking another example from chromium chemistry involving only a change
of oxidation state (from +2 to +3):

http://www.chemguide.co.uk/inorganic/complexions/colour.html (13 of 15)30/12/2004 11:32:41

complex ions - colour

The 2+ ion is almost the same colour as the hexaaquacopper(II) ion, and
the 3+ ion is the hard-to-describe violet-blue-gey colour.

The co-ordination of the ion


Splitting is greater if the ion is octahedral than if it is tetrahedral, and
therefore the colour will change with a change of co-ordination.
Unfortunately, I can't think of a single example to illustrate this with!
The problem is that an ion will only change co-ordination if you change the
ligand - and changing the ligand will change the colour as well. You can't
isolate out the effect of the co-ordination change.
For example, a commonly quoted case comes from cobalt(II) chemistry,
with the ions [Co(H2O)6]2+ and [CoCl4]2-.

The difference in the colours is going to be a combination of the effect of


the change of ligand, and the change of the number of ligands.

Where would you like to go now?


To the complex ion menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

http://www.chemguide.co.uk/inorganic/complexions/colour.html (14 of 15)30/12/2004 11:32:41

complex ions - colour

Jim Clark 2003

http://www.chemguide.co.uk/inorganic/complexions/colour.html (15 of 15)30/12/2004 11:32:41

substitution in complex ions - ligand exchange

COMPLEX METAL IONS - LIGAND EXCHANGE


REACTIONS

This page describes some common ligand exchange (substitution)


reactions involving complex metal ions. It assumes that you are familiar
with basic ideas about complex ions.

Note: If you aren't happy about complex ions, then explore


the complex ion menu before you go on, or refer back to it if
you come across things that you don't understand.

Examples of ligand exchange reactions


A ligand exchange reaction is exactly what it says - a reaction in which
one ligand in a complex ion is replaced by a different one. The following
examples are taken from UK A' level syllabuses.

Replacing water with chloride ions


Replacing the water in the hexaaquacobalt(II) ion
If you add concentrated hydrochloric acid to a solution containing
hexaaquacobalt(II) ions (for example, cobalt(II) chloride solution), the
solution turns from its original pink colour to a dark rich blue. The six
water molecules are replaced by four chloride ions.

http://www.chemguide.co.uk/inorganic/complexions/ligandexch.html (1 of 11)30/12/2004 11:32:51

substitution in complex ions - ligand exchange

The reaction taking place is reversible.

Concentrated hydrochloric acid is used as the source of chloride ions


because it provides a very high concentration compared to what is
possible with, say, sodium chloride solution. Concentrated hydrochloric
acid has a chloride ion concentration of approximately 10 mol dm-3.
The high chloride ion concentration pushes the position of the equilibrium
to the right according to Le Chatelier's Principle.

Note: You really need to know about Le Chatelier's


Principle, particularly with regard to the effect of changes in
concentration on the position of equilibrium. Follow this link if
you aren't sure.
Use the BACK button on your browser to return quickly to this
page.

http://www.chemguide.co.uk/inorganic/complexions/ligandexch.html (2 of 11)30/12/2004 11:32:51

substitution in complex ions - ligand exchange

Notice the change in the co-ordination of the cobalt. Chloride ions are
bigger than water molecules, and there isn't room to fit six of them
around the central cobalt ion.
This reaction can be easily reversed by adding water to the solution.
Adding water to the right-hand side of the equilibrium has the effect of
moving the position of equilibrium to the left. The pink colour of the
hexaaquacobalt(II) ion is produced again (only paler, of course, because
it is more dilute).

Replacing the water in the hexaaquacopper(II) ion


In terms of the chemistry, this is exactly the same as the last example all that differs are the colours. Unfortunately, these aren't quite so
straightforward. The colour of the tetrachlorocuprate(II) ion is almost
always seen mixed with that of the original hexaaqua ion.
What you normally see is:

http://www.chemguide.co.uk/inorganic/complexions/ligandexch.html (3 of 11)30/12/2004 11:32:51

substitution in complex ions - ligand exchange

The reaction taking place is reversible, and you get a mixture of colours
due to both of the complex ions.

You may find the colour of the tetrachlorocuprate(II) ion variously


described as olive-green or yellow.

Help! If you dissolve copper(II) oxide in concentrated


hydrochloric acid (to make copper(II) chloride which will then
go on to form the tetrachlorocuprate(II) ion), you get a dark
yellowish brown solution. This probably better reflects the
colour of the ion.
What are you going to call the colour in an exam? Don't worry
about it particularly. Your examiners will almost certainly
allow any valid colour. If in doubt look at mark schemes and
support material produced by your examiners. If you are a
UK A' level student and haven't got these, find out how to get
them by visiting the syllabuses page.
Use the BACK button (or the History file or the Go menu depending on how waylaid you get!) to return to this page
later.

http://www.chemguide.co.uk/inorganic/complexions/ligandexch.html (4 of 11)30/12/2004 11:32:51

substitution in complex ions - ligand exchange

Adding water to the green solution, replaces the chloride ions as ligands
by water molecules again, and the solution returns to blue.

Replacing water molecules by ammonia


Water molecules and ammonia molecules are very similar in size, and so
there is no change in co-ordination this time. Unfortunately, the reactions
aren't quite so straightforward to describe.
Ammonia solution can react with hexaaqua metal ions in two quite
distinct ways, because it can act as a base as well as a ligand.
If you add a small amount of ammonia solution you get precipitates of the
metal hydroxide - the ammonia is acting as a base. In some cases, these
precipitates redissolve when you add more ammonia to give solutions in
which a ligand exchange reaction has occurred.
In the diagrams below, both steps are shown, but we are only going to
consider the chemistry of the overall ligand exchange reaction. The
precipitates dissolve because of a complicated series of equilibrium
shifts, and we shan't worry about that for the moment.

http://www.chemguide.co.uk/inorganic/complexions/ligandexch.html (5 of 11)30/12/2004 11:32:51

substitution in complex ions - ligand exchange

Note: You will find full details of the reactions involved in the
formation of the precipitates described on a separate page
about the reaction of ammonia with metal aqua ions.
That page also describes the quite complex reasons why the
precipitate dissolves again.

Replacing the water in the hexaaquacopper(II) ion


This is a slightly untypical case, because only four of the six water
molecules get replaced to give the tetraamminediaquacopper(II) ion, [Cu
(NH3)4(H2O)2]2+.

Notice that the four ammonias all lie in one plane, with the water
molecules above and below.
What you see in a test tube is:

http://www.chemguide.co.uk/inorganic/complexions/ligandexch.html (6 of 11)30/12/2004 11:32:51

substitution in complex ions - ligand exchange

The main equilibrium involved in the ligand exchange reaction is:

The colour of the deep blue complex is so strong that this reaction is
used as a sensitve test for copper(II) ions in solution. Even if you try to
reverse the change by adding large amounts of water to the equilibrium,
the strength of the deep blue (even highly diluted) always masks the pale
blue of the aqua ion.

Replacing the water in the hexaaquacobalt(II) ion


This time, all the water molecules get replaced.

http://www.chemguide.co.uk/inorganic/complexions/ligandexch.html (7 of 11)30/12/2004 11:32:51

substitution in complex ions - ligand exchange

The straw coloured solution formed changes colour very rapidly on


standing to a deep reddish brown. The hexaamminecobalt(II) ions are
oxidised by the air to hexaamminecobalt(III) ions. However, that is a
quite separate reaction, and isn't a part of the ligand exchange reaction.

Replacing the water in the hexaaquachromium(III) ion


Again, all the water molecules get replaced by ammonias. The difference
this time is that the reaction isn't so complete. The precipitate has to be
left to stand in the presence of excess concentrated ammonia solution for
some time in order to get the ammine complex.
Even so, you still get left with some unreacted precipitate.

Two more replacements of the water in the hexaaquachromium(III)


ion
The colour of the hexaaquachromium(III) ion has been shown as a
"difficult to describe" violet-blue-grey in all the diagrams above. In
practice, when it is produced during a reaction in a test tube, it is often
green.
A typical example of this is the use of acidified potassium dichromate(VI)
http://www.chemguide.co.uk/inorganic/complexions/ligandexch.html (8 of 11)30/12/2004 11:32:51

substitution in complex ions - ligand exchange

as an oxidising agent. Whenever this is used, the orange solution turns


green and we nearly always describe the green ion as being Cr3+(aq) implying the hexaaquachromium(III) ion. That's actually an oversimplification.
What happens is that one or more of the ligand water molecules get
replaced by a negative ion in the solution - typically sulphate or chloride.
Replacement of the water by sulphate ions
You can do this simply by warming some chromium(III) sulphate solution.

One of the water molecules is replaced by a sulphate ion. Notice the


change in the charge on the ion. Two of the positive charges are
cancelled by the presence of the two negative charges on the sulphate
ion.
Replacement of the water by chloride ions
In the presence of chloride ions (for example with chromium(III) chloride),
the most commonly observed colour is green. This happens when two of
the water molecules are replaced by chloride ions to give the
tetraaquadichlorochromium(III) ion - [Cr(H2O)4Cl2]+.
Once again, notice that replacing water molecules by chloride ions
changes the charge on the ion.

http://www.chemguide.co.uk/inorganic/complexions/ligandexch.html (9 of 11)30/12/2004 11:32:51

substitution in complex ions - ligand exchange

Note: Technically, the chloride ion ligand should be written


in the formula before the neutral water molecules, but you will
find both versions. The one I have chosen to use is probably
the more commonly found, and has the advantage that it is
consistent with the order the ligands appear in the name.

A ligand exchange reaction in the test for iron(III) ions


This provides an extremely sensitive test for iron(III) ions in solution.
If you add thiocyanate ions, SCN-, (from, say, sodium or potassium or
ammonium thiocyanate solution) to a solution containing iron(III) ions,
you get an intense blood red solution containing the ion [Fe(SCN)(H2O)5]
2+.

http://www.chemguide.co.uk/inorganic/complexions/ligandexch.html (10 of 11)30/12/2004 11:32:51

substitution in complex ions - ligand exchange

A note on the colours: I haven't been able to suggest


properly the intense rich red of the thiocyano complex.
Imagine blood!
The colour of the original solution isn't the colour of the [Fe
(H2O)6]3+ ion. This is actually very pale violet, but is rarely
seen in solution. That complex reacts with water to some
extent to produce other more strongly coloured yellow and
orange complexes. This is explained in more detail on the
page about the acidity of the aqua ions.

Where would you like to go now?


To the complex ion menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/inorganic/complexions/ligandexch.html (11 of 11)30/12/2004 11:32:51

reactions of aqua ions with ammonia solution

REACTIONS OF HEXAAQUA METAL IONS WITH


AMMONIA SOLUTION

This page describes and explains the reactions between complex ions of
the type [M(H2O)6]n+ and ammonia solution.

Important: Before you read this page, you need to know


about the acidity of the hexaaqua ions.
It would also be useful to read two more pages: the one
describing the reactions of these ions with hydroxide ions,
and the one explaining ligand exchange reactions.
The present page draws information from all of these pages,
and you will find it easier to understand if you spend some
time reading these other pages first.
It will probably be easiest to return to this page via the
complex ions menu.

Reactions of the hexaaqua ions with ammonia solution are complicated


by the fact that the ammonia can have two quite different functions. It can
act as a base (in the Bronsted-Lowry sense), but it is also a possible
ligand which can replace water molecules around the central metal ion.
When it acts as a ligand, it is acting as a Lewis base.
We need to look at these two functions separately.

http://www.chemguide.co.uk/inorganic/complexions/aquanh3.html (1 of 13)30/12/2004 11:33:07

reactions of aqua ions with ammonia solution

Note: If you aren't familiar with either of these terms, you


should follow this link to a page about theories of acids and
bases.
Use the BACK button on your browser to return to this page
when you are confident about these terms.

Ammonia acting as a (Bronsted-Lowry) base


The general case
This is what happens when you only add small amounts of dilute
ammonia solution to any of the hexaaqua ions. The ligand effect only
happens with an excess of ammonia or with concentrated ammonia - and
with some metals you don't even see it then.
We'll talk through what happens if you add a small amount of dilute
ammonia solution to a solution containing a 2+ hexaaqua ion.
These have the formula [M(H2O)6]2+, and they are acidic. Their acidity is
shown in the reaction of the hexaaqua ions with water molecules from
the solution:

They are acting as acids by donating hydrogen ions to water molecules


in the solution.
Because of the confusing presence of water from two different sources
(the ligands and the solution), it is easier to simplify this:

http://www.chemguide.co.uk/inorganic/complexions/aquanh3.html (2 of 13)30/12/2004 11:33:07

reactions of aqua ions with ammonia solution

Adding ammonia solution to this equilibrium - stage 1


There are two possible reactions.
Reaction of ammonia with the hydroxonium ions (hydrogen ions)
Ammonia will react with these to produce ammonium ions.
According to Le Chatelier's Principle, the position of equilibrium will move
to the right, producing more of the new complex ion.

Note: You really need to know about Le Chatelier's


Principle, particularly with regard to the effect of changes in
concentration on the position of equilibrium. Follow this link if
you aren't sure.
Use the BACK button on your browser to return quickly to this
page.

http://www.chemguide.co.uk/inorganic/complexions/aquanh3.html (3 of 13)30/12/2004 11:33:07

reactions of aqua ions with ammonia solution

Reaction of ammonia with the hexaaqua ion


Statistically, there is far more chance of an ammonia molecule hitting a
hexaaqua metal ion than of hitting a hydrogen ion. There are far more
hexaaqua ions present.
If that happens, you get exactly the same new complex ion formed as
above.

Notice that this is still a reversible change (unlike the corresponding


change when you add hydroxide ions). Ammonia is only a weak base.
The second stage of the reaction
Whichever of the above reactions happens, you end up with [M(H2O)5
(OH)]+ ions in solution. These are also acidic, and can lose hydrogen
ions from another of the water ligands.
Taking the easier version of the equilibrium:

Adding ammonia again tips the equilibrium to the right - either by reacting
with the hydrogen ions, or by reacting directly with the complex on the
left-hand side.
When this happens, the new complex formed no longer has a charge this is a "neutral complex". It is insoluble in water - and so a precipitate is
formed.
This precipitate is often written without including the remaining water
ligands. In other words we write it as M(OH)2. A precipitate of the metal
hydroxide has been formed.
Summarising what has happened

http://www.chemguide.co.uk/inorganic/complexions/aquanh3.html (4 of 13)30/12/2004 11:33:07

reactions of aqua ions with ammonia solution

You can also usefully write the complete change as an overall


equilibrium reaction. This will be important for later on.

If you did the same reaction with a 3+ ion, the only difference is that you
would have to remove a total of 3 hydrogen ions in order to get to the
neutral complex. That would give the overall equation:

Looking at the ions of specific metals


Remember that we are concentrating for the moment on the ammonia
acting as a base - in other words, on the formation of hydroxide
precipitates when you add small amounts of ammonia solution to
solutions containing hexaaqua metal ions.
The diagrams, however, will show the complete change so I don't have to
repeat them later on. Ignore the cases where the precipitate dissolves in
excess ammonia for the moment.
2+ ions
hexaaquacobalt(II)

http://www.chemguide.co.uk/inorganic/complexions/aquanh3.html (5 of 13)30/12/2004 11:33:07

reactions of aqua ions with ammonia solution

Note: The final solution rapidly darkens in air to a deep redbrown. This is due to oxidation from hexaamminecobalt(II) to
hexaamminecobalt(III) ions.
This oxidation is described in more detail on the page about
cobalt chemistry.
Use the BACK button on your browser to return quickly to this
page.

hexaaquacopper(II)

http://www.chemguide.co.uk/inorganic/complexions/aquanh3.html (6 of 13)30/12/2004 11:33:07

reactions of aqua ions with ammonia solution

hexaaquairon(II)

Iron is very easily oxidised under alkaline conditions. Oxygen in the air
oxidises the iron(II) hydroxide precipitate to iron(III) hydroxide especially
around the top of the tube. The darkening of the precipitate comes from
the same effect. This is NOT a ligand exchange reaction.
hexaaquamanganese(II)

http://www.chemguide.co.uk/inorganic/complexions/aquanh3.html (7 of 13)30/12/2004 11:33:07

reactions of aqua ions with ammonia solution

I have shown the original solution as very pale pink (the palest I can
produce!), but in fact it is virtually colourless. The pale brown precipitate
is oxidised to darker brown manganese(III) oxide in contact with oxygen
from the air. Again, this isn't a ligand exchange reaction.
hexaaquanickel(II)

hexaaquazinc

http://www.chemguide.co.uk/inorganic/complexions/aquanh3.html (8 of 13)30/12/2004 11:33:07

reactions of aqua ions with ammonia solution

You start and finish with colourless solutions, producing a white


precipitate on the way.

Note: I have shown the final zinc ion as [Zn(NH3)4]2+, but I'm
not certain about this. Some sources (although a minority)
show it as [Zn(NH3)4(H2O)2]2+. I don't know which is right!

3+ ions
hexaaquaaluminium

http://www.chemguide.co.uk/inorganic/complexions/aquanh3.html (9 of 13)30/12/2004 11:33:07

reactions of aqua ions with ammonia solution

Starting from a colourless solution, you get a white precipitate.


hexaaquachromium(III)

hexaaquairon(III)

Summary of the effect of adding small amounts of ammonia solution


In each case you get a precipitate of the neutral complex - the metal
hydroxide. Apart from minor differences in the exact shade of colour you
get, these are almost all exactly the same as the precipitates you get
when you add a little sodium hydroxide solution to the solutions of the
http://www.chemguide.co.uk/inorganic/complexions/aquanh3.html (10 of 13)30/12/2004 11:33:07

reactions of aqua ions with ammonia solution

hexaaqua ions. The only real difference lies in the colour of the cobalt
precipitate.

Ammonia acting as a ligand


The ligand exchange reaction
In some cases, ammonia replaces water around the central metal ion to
give another soluble complex. This is known as a ligand exchange
reaction, and involves an equilibrium such as this one:

The formation of this new soluble complex causes the precipitate to


dissolve.
The ammonia attaches to the central metal ion using the lone pair of
electrons on the nitrogen atom. Because it is a lone pair donor, it is
acting as a Lewis base.

Explaining why the precipitate dissolves


Almost all text books leave the argument at this point, assuming that it is
obvious why the formation of the complex causes some precipitates to
dissolve. It isn't! If you want to know the quite complicated reasons, read
on . . .
We'll take the copper case as typical of any of them.
There are two equilibria involved in this. The first is the one in which
ammonia is acting as a base and producing the precipitate:

The other one is the ligand exchange reaction:

http://www.chemguide.co.uk/inorganic/complexions/aquanh3.html (11 of 13)30/12/2004 11:33:07

reactions of aqua ions with ammonia solution

Notice that the hexaaqua ion appears in both of these. There is now an
interaction between the two equilibria:

Looking at it like this is helpful in explaining why some precipitates


dissolve in excess ammonia while others don't. It depends on the
positions of the equilibria.
To get the precipitate to dissolve, you obviously need the ligand
exchange equilibrium to lie well to the right, but you need the acid-base
equilibrium to be easy to pull back to the left.

Where would you like to go now?


http://www.chemguide.co.uk/inorganic/complexions/aquanh3.html (12 of 13)30/12/2004 11:33:07

reactions of aqua ions with ammonia solution

To the complex ion menu . . .


To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/inorganic/complexions/aquanh3.html (13 of 13)30/12/2004 11:33:07

complex ions - acidity of the aqua ions

COMPLEX METAL IONS - THE ACIDITY OF THE


HEXAAQUA IONS

This page explains why complex ions of the type [M(H2O)6]n+ are acidic.

The general explanation


A closer look at the distribution of charge in the ion
The pH's of solutions containing hexaaqua ions vary a lot from one metal
to another (assuming you are comparing solutions of equal
concentrations). However, the underlying explanation is the same for all
of them.
We'll take the hexaaquairon(III) ion, [Fe(H2O)6]3+ as typical. The
structure of the ion is:

Each of the six water molecules are attached to the central iron(III) ion
via a co-ordinate bond using one of the lone pairs on the oxygen.

http://www.chemguide.co.uk/inorganic/complexions/acidity.html (1 of 10)30/12/2004 11:33:14

complex ions - acidity of the aqua ions

Note: This bonding is explained in some detail on the


introductory page on complex ions. Follow this link if you
aren't sure.
Use the BACK button on your browser to return quickly to this
page.

We'll choose one of these water molecules at random (it doesn't make
any difference which one!), and look at the bonding in a bit more detail showing all the bonds around the oxygen.
Imagine for the moment that the 3+ charge is located entirely on the iron.

When the lone pairs on the oxygens form co-ordinate bonds with the
iron, there is obviously a movement of electrons towards the iron.
That has an effect on the electrons in the O-H bonds. These electrons, in
turn, get pulled towards the oxygen even more than usual. That leaves
the hydrogen nuclei more exposed than normal.

http://www.chemguide.co.uk/inorganic/complexions/acidity.html (2 of 10)30/12/2004 11:33:14

complex ions - acidity of the aqua ions

The overall effect is that each of the hydrogen atoms is more positive
than it is in ordinary water molecules. The 3+ charge is no longer located
entirely on the iron but spread out over the whole ion - much of it on the
hydrogen atoms.
The effect of dissolving this ion in water
The theory
The hydrogen atoms attached to the water ligands are sufficiently
positive that they can be pulled off in a reaction involving water
molecules in the solution.
The first stage of this process is:

The complex ion is acting as an acid by donating a hydrogen ion to water


molecules in the solution. The water is, of course, acting as a base by
accepting the hydrogen ion.
Because of the confusing presence of water from two different sources
(the ligands and the solution), it is easier to simplify this:

However, if you write it like this, remember that the hydrogen ion isn't just
falling off the complex ion. It is being pulled off by a water molecule in the
solution. Whenever you write "H+(aq)" what you really mean is a
hydroxonium ion, H3O+.
The hexaaquairon(III) ion is quite strongly acidic giving solutions with
pH's around 1.5, depending on concentration. You can get further loss of
hydrogen ions as well, from a second and a third water molecule.
Losing a second hydrogen ion:

. . . and a third one:


http://www.chemguide.co.uk/inorganic/complexions/acidity.html (3 of 10)30/12/2004 11:33:14

complex ions - acidity of the aqua ions

This time you end up with a complex with no charge on it. This is
described as a neutral complex. Because it has no charge, it doesn't
dissolve in water to any extent, and a precipitate is formed.
In practice
What do you actually get in solution if you dissolve an iron(III) salt in
water? In fact you get a mixture of all the complexes that you have seen
in the equations above. These reactions are all equilibria, so everything
will be present. The proportions depend on how concentrated the
solution is.
The colour of the solution is very variable and depends in part on the
concentration of the solution. Dilute solutions containing iron(III) ions can
be pale yellow. More concentrated ones are much more orange, and
may even produce some orange precipitate.
None of these colours represents the true colour of the [Fe(H2O)6]3+ ion which is a very pale lilac colour! That colour is only really easy to see in
solids containing the ion.

http://www.chemguide.co.uk/inorganic/complexions/acidity.html (4 of 10)30/12/2004 11:33:14

complex ions - acidity of the aqua ions

Looking at the equilibrium showing the loss of the first hydrogen ion:

The colour of the new complex ion on the right-hand side is so strong
that it completely masks the colour of the hexaaqua ion.
In concentrated solutions, the equilibrium position will be even further to
the right-hand side (Le Chatelier's Principle), and so the colour darkens.
You will also get significant loss of other hydrogen ions leading to some
formation of the neutral complex - and so you get some precipitate.

Note: This is a bit of a simplification, but adequate for this


level. In fact you get other changes happening to the new
complex ions as well. For example, the first one dimerises introducing further equilibria and colours! Don't worry about it!

You can move the position of this equilibrium by adding extra hydrogen
ions from a concentrated acid - for example, by adding concentrated
nitric acid to a solution of iron(III) nitrate.
The new hydrogen ions push the position of the equilibrium to the left so
that you can see the colour of the hexaaqua ion. This is slightly easier to
follow if you write the simplified version of the equilibrium.

http://www.chemguide.co.uk/inorganic/complexions/acidity.html (5 of 10)30/12/2004 11:33:14

complex ions - acidity of the aqua ions

Note: I haven't tried to draw this colour change because


when I last tried this I couldn't make it work properly
(although I have done in the past)! The solution concentration
is quite critical. If it is too dilute, the yellow colour disappears
and you are left with a solution which is so faintly lilac that it
is (in all honesty!) colourless. If the solution is too
concentrated, you never get rid of the strong orange colour of
the other complex ions, and so the lilac colour remains
masked.

Why are 3+ ions more acidic than 2+ ions


The effect of charge on the acidity of the hexaaqua ions
Solutions containing 3+ hexaaqua ions tend to have pH's in the range
from 1 to 3. Solutions containing 2+ ions have higher pH's - typically
around 5 - 6, although they can go down to about 3.
Remember that the reason that these ions are acidic is because of the
pull of the electrons towards the positive central ion. An ion with 3+
charges on it is going to pull the electrons more strongly than one with
only 2+ charges.
In 3+ ions, the electrons in the O-H bonds will be pulled further away
from the hydrogens than in 2+ ions.
That means that the hydrogen atoms in the ligand water molecules will
have a greater positive charge in a 3+ ion, and so will be more attracted
to water molecules in the solution.
If they are more attracted, they will be more readily lost - and so the 3+
ions are more acidic.

The effect of ionic radius on acidity

http://www.chemguide.co.uk/inorganic/complexions/acidity.html (6 of 10)30/12/2004 11:33:14

complex ions - acidity of the aqua ions

Charge density and acidity


If you have ions of the same charge, it seems reasonable that the
smaller the volume this charge is packed into, the greater the distorting
effect on the electrons in the O-H bonds.
Ions with the same charge but in a smaller volume (a higher charge
density) would be expected to be more acidic.
You would therefore expect to find that the smaller the radius of the
metal ion, the stronger the acid. Unfortunately, it's not that simple!

Warning! If you like your chemistry all tidy and simple, don't
read any further. I have got to the point that most text books
get to at this level. In fact, as I will show below, it is quite
difficult to show this effect of ionic radius in any simple ions.
There are too many exceptions.

Plotting pKa values against ionic radius


pKa
pKa is a scale for comparing the strengths of weak acids. All you really
need to know for the present topic is that the lower the value of pKa, the
stronger the acid. As a reasonable approximation, if you divide the pKa
value of a weak acid by 2, that will give you the pH of a 1 mol dm-3
solution of the acid.

Note: It isn't necessary for this topic, but if you need to know
more about pKa, you can find it by following this link.
Use the BACK button on your browser to return quickly to this
page.

http://www.chemguide.co.uk/inorganic/complexions/acidity.html (7 of 10)30/12/2004 11:33:14

complex ions - acidity of the aqua ions

Ionic radius
A real problem arises here! Almost every data source that you refer to
quotes different values for ionic radii. This is because the radius of a
metal ion varies depending on what negative ion it is associated with. We
are interested in what happens when the metal ion is bonded to water
molecules, so haven't got simple ions at all! Whatever values we take are
unlikely to represent the real situation.
In the graphs which follow I have taken values for ionic radii from two
common sources so that you can see what a lot of difference it makes to
the argument.
The graphs
If it is true that the smaller the ionic radius, the stronger the acidity of the
hexaaqua ion, you would expect some sort of regular increase in pKa
(showing weaker acids) as ionic radius increases.
The following graphs plot pKa against ionic radii for the 2+ ions of the
elements in the first transition series from vanadium to copper.
The first graph plots pKa against ionic radii taken from Chemistry Data
Book by Stark and Wallace.

You can see that there is a trend for several of the ions, but it is
completely broken by vanadium and chromium.
The second graph uses ionic radii taken from the Nuffield Advanced
http://www.chemguide.co.uk/inorganic/complexions/acidity.html (8 of 10)30/12/2004 11:33:14

complex ions - acidity of the aqua ions

Science Book of Data.

In this case, the pattern is much more confused. Which is right? I don't
know!!
A final comment
During the couple of weeks I spent trying to understand this, I failed
completely. The only conclusion I came to is that there probably is a
relationship between ionic radius and acid strength, but that it is nothing
like as simple and straightforward as most books at this level pretend.
The problem is that there are other more important effects operating as
well (quite apart from differences in charge) and these can completely
swamp the effect of the changes in ionic radius. You have to look in far
more detail at the bonding in the hexaaqua ions and the product ions.
This is all just too difficult to understand at this level. In fact, to be honest,
I did wonder while I was researching this whether anyone really
understood all of this! There is, of course, no reason why they should - it
is nonsense to think that we can explain everything in chemistry at the
present moment.

Where would you like to go now?


To the complex ion menu . . .
To the Inorganic Chemistry menu . . .

http://www.chemguide.co.uk/inorganic/complexions/acidity.html (9 of 10)30/12/2004 11:33:14

complex ions - acidity of the aqua ions

To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/inorganic/complexions/acidity.html (10 of 10)30/12/2004 11:33:14

reactions of aqua ions with hydroxide ions

REACTIONS OF HEXAAQUA METAL IONS WITH


HYDROXIDE IONS

This page describes and explains the reactions between complex ions of
the type [M(H2O)6]n+ and hydroxide ions from, for example, sodium
hydroxide solution.
It assumes that you know why these ions are acidic, and are happy
about the equilibria involved.

Important: If you aren't happy about the acidity of the


hexaaqua ions, it is essential that you follow this link before
you go any further.

The general case


Although there are only minor differences, for simplicity we will look at 2+
ions and 3+ ions separately.
Adding hydroxide ions to 2+ hexaaqua ions
These have the form [M(H2O)6]2+. Their acidity is shown in the reaction
of the hexaaqua ions with water molecules from the solution:

They are acting as acids by donating hydrogen ions to water molecules


in the solution.
Because of the confusing presence of water from two different sources
(the ligands and the solution), it is easier to simplify this:

http://www.chemguide.co.uk/inorganic/complexions/aquaoh.html (1 of 16)30/12/2004 11:33:25

reactions of aqua ions with hydroxide ions

Disturbing this equilibrium by adding hydroxide ions - stage 1


What happens if you add hydroxide ions to this equilibrium? There are
two possible reactions.
Reaction of hydroxide ions with the hydroxonium ions (hydrogen ions)
According to Le Chatelier's Principle, the position of equilibrium will move
to the right, producing more of the new complex ion.

Note: You really need to know about Le Chatelier's


Principle, particularly with regard to the effect of changes in
concentration on the position of equilibrium. Follow this link if
you aren't sure.
Use the BACK button on your browser to return quickly to this
page.

http://www.chemguide.co.uk/inorganic/complexions/aquaoh.html (2 of 16)30/12/2004 11:33:25

reactions of aqua ions with hydroxide ions

Reaction of hydroxide ions with the hexaaqua ion


Statistically, there is far more chance of a hydroxide ion hitting a
hexaaqua metal ion than of hitting a hydrogen ion. There are far more
hexaaqua ions present.
If that happens, you get exactly the same new complex ion formed as
above.

Notice that this isn't a ligand exchange reaction. The hydroxide ion has
removed a hydrogen ion from one of the ligand water molecules. The
reaction has also become virtually one-way.
The second stage of the reaction
Whichever of the above reactions happens, you end up with [M(H2O)5
(OH)]+ ions in solution. These are also acidic, and can lose hydrogen
ions from another of the water ligands.
Taking the easier version of the equilibrium:

Adding hydroxide ions again tips the equilibrium to the right - either by
reacting with the hydrogen ions, or by reacting directly with the complex
on the left-hand side.
When this happens, the new complex formed no longer has a charge we describe it as a "neutral complex". In all the cases we are looking at,
this neutral complex is insoluble in water - and so a precipitate is formed.
This precipitate is often written without including the remaining water
ligands. In other words we write it as M(OH)2. A precipitate of the metal
hydroxide has been formed.

http://www.chemguide.co.uk/inorganic/complexions/aquaoh.html (3 of 16)30/12/2004 11:33:25

reactions of aqua ions with hydroxide ions

Note: Which version of the formula for the metal hydroxide


should you write? The full version, [M(H2O)4(OH)2], shows
more clearly what is going on. Be guided by what your
examiners prefer. It may actually vary depending on what
context the compound crops up in. My preference is for the
full version, and that's what I shall use throughout.

Summarising what has happened so far

Going further
There is no logical reason why the removal of hydrogen ions from the
complex should stop at this point. Further hydrogen ions can be removed
by hydroxide ions to produce anionic complexes - complexes carrying
negative charges.
Whether this actually happens in the test tube to any extent varies from
metal to metal.

http://www.chemguide.co.uk/inorganic/complexions/aquaoh.html (4 of 16)30/12/2004 11:33:25

reactions of aqua ions with hydroxide ions

In fact, if you do this using sodium hydroxide solution of the usual


concentrations, most of the 2+ ions that you will meet at this level don't
go beyond the precipitate. The only one you are likely to come across is
the zinc case - and that has a complication. The final ion is [Zn(OH)4]2- a tetrahedral ion which has lost the remaining 2 water ligands.

Adding hydroxide ions to 3+ hexaaqua ions


The argument here is exactly as before - the only difference is the
number of hydrogen ions which have to be removed from the original
hexaaqua complex to produce the neutral complex.
Going beyond the neutal complex is also rather more common with 3+
than with 2+ ions, and may go as far as having a hydrogen ion removed
from each of the six water molecules.
This is summarised in the same sort of flow scheme as before:

http://www.chemguide.co.uk/inorganic/complexions/aquaoh.html (5 of 16)30/12/2004 11:33:25

reactions of aqua ions with hydroxide ions

http://www.chemguide.co.uk/inorganic/complexions/aquaoh.html (6 of 16)30/12/2004 11:33:25

reactions of aqua ions with hydroxide ions

Note: It would be absurd to sit and learn all this! All you
need to do is to understand that one hydrogen ion gets
removed at a time. When you have a neutral complex, it will
form a precipitate. That is equally true of the 2+ or 3+ cases.
Just be careful with the charges on the complexes.
Remember that for every hydrogen ion you remove, you will
lose a positive charge (or will gain one more negative charge).
When you come to looking at individual metals, you will, of
course, have to remember where the sequence usually stops.

Looking at the ions of specific metals


In each case the formula of the precipitate will be given as if it were the
simple neutral complex. In fact, these "hydroxide" precipitates sometimes
rearrange by losing water from combinations of the attached OH groups.
This produces oxides closely associated with the lost water. These
changes are beyond the scope of this site.
2+ ions
hexaaquacobalt(II)

http://www.chemguide.co.uk/inorganic/complexions/aquaoh.html (7 of 16)30/12/2004 11:33:25

reactions of aqua ions with hydroxide ions

Note: I don't know the reason for the change in colour. One
reliable source quotes it vaguely as a change in coordination
of the metal ion. Another suggests that the blue precipitate
involves a chloride ion (from the cobalt(II) chloride solution if
that's what you used) as a ligand. You are unlikely to need to
worry about this at this level. Remember the colour changes!

hexaaquacopper(II)

hexaaquairon(II)

Iron is very easily oxidised under alkaline conditions. Oxygen in the air
oxidises the iron(II) hydroxide precipitate to iron(III) hydroxide especially
around the top of the tube. The darkening of the precipitate comes from
the same effect.
http://www.chemguide.co.uk/inorganic/complexions/aquaoh.html (8 of 16)30/12/2004 11:33:25

reactions of aqua ions with hydroxide ions

hexaaquamanganese(II)

I have shown the original solution as very pale pink (the palest I can
produce!), but in fact it is virtually colourless. The pale brown precipitate
is oxidised to darker brown manganese(III) oxide in contact with oxygen
from the air.
hexaaquanickel(II)

hexaaquazinc

http://www.chemguide.co.uk/inorganic/complexions/aquaoh.html (9 of 16)30/12/2004 11:33:25

reactions of aqua ions with hydroxide ions

You start and finish with colourless solutions, producing a white


precipitate on the way.

3+ ions
hexaaquaaluminium

This looks exactly the same in the test tube as the corresponding zinc
reaction above - but beware the different formulae of the precipitate and
the final solution.
hexaaquachromium(III)

http://www.chemguide.co.uk/inorganic/complexions/aquaoh.html (10 of 16)30/12/2004 11:33:25

reactions of aqua ions with hydroxide ions

hexaaquairon(III)

Note: The original colour of the iron(III) ion is very variable from a pale yellow to a darkish orange-brown. None of these
colours is actually the colour of the true hexaaquairon(III) ion.
This is explored in more detail on the page about the acidity
of the hexaaqua ions.

http://www.chemguide.co.uk/inorganic/complexions/aquaoh.html (11 of 16)30/12/2004 11:33:25

reactions of aqua ions with hydroxide ions

Reversing the process


If adding hydroxide ions removes hydrogen ions from the hexaaqua
complex one at a time, it doesn't seem unreasonable that you could put
them back again by adding an acid. That's just what happens!
We'll look in detail at what happens in the chromium(III) case, but exactly
the same principle applies to all the other examples we've looked at whether for 2+ or 3+ ions. As long as you understand what is happening,
you can work out the details if you need to.
Adding acid to hexahydroxochromate(III) ions
These are the ions formed at the end of the sequence in which you add
hydroxide ions to a solution containing hexaaquachromium(III) ions.
Their formula is [Cr(OH)6]3-.

Note: If you are bothered about the name of these ions:


"hexa" = 6; "hydroxo" = OH groups; "chromate(III)" =
chromium in the +3 oxidation state, but in a negative ion. If
you need to explore these names in more detail, you could
look at the page on naming, but it isn't very important for the
present topic.

A reminder of the colour changes when you add sodium hydroxide


solution to a solution containing hexaaquachromium(III) ions:

http://www.chemguide.co.uk/inorganic/complexions/aquaoh.html (12 of 16)30/12/2004 11:33:25

reactions of aqua ions with hydroxide ions

If you add an acid (dilute sulphuric acid, for example), the hydrogen ions
get put back on one at a time.

http://www.chemguide.co.uk/inorganic/complexions/aquaoh.html (13 of 16)30/12/2004 11:33:25

reactions of aqua ions with hydroxide ions

You already know the colours of the significant stages (the beginning, the
end, and the neutral complex). It isn't a separate bit of learning!
You can apply this to any case. If you know the colours as you remove
hydrogen ions, you automatically know them as you put the hydrogen
ions back on again.
It also doesn't matter where you start from either - whether, for example,
you add acid to an ionic complex like [Cr(OH)6]3-, or a neutral one like [Fe
(H2O)4(OH)2].
You will know that the [Fe(H2O)4(OH)2] is a dirty green precipitate. When
you add the hydrogen ions back to it, it will revert to the very pale green
solution of the [Fe(H2O)6]2+ ion. None of this is a new bit of learning - you
just have to re-arrange what you already know!

Note: In fact, in the chromium case, it is actually slightly


more difficult than this, because the final solution tends to
end up green rather than the usual strange blue-grey-violet of
the hexaaquachromium(III) ion.
This happens whenever you produce hexaaquachromium(III)
ions in a test tube. The problem is that other ions from the
solution (chloride or sulphate) replace one or more water
molecules in the complex and give green colours. This is
explored some more in the page on ligand exchange
reactions

http://www.chemguide.co.uk/inorganic/complexions/aquaoh.html (14 of 16)30/12/2004 11:33:25

reactions of aqua ions with hydroxide ions

Amphoteric hydroxides
An amphoteric substance has both acidic and basic properties. In other
words, it will react with both bases and acids. Some of the metal
hydroxides we've been looking at are doing exactly that.
Chromium(III) hydroxide as an amphoteric hydroxide
"Chromium(III) hydroxide" is a simple way of naming the neutral complex
[Cr(H2O)3(OH)3]. You have seen that it reacts with bases (hydroxide
ions) to give [Cr(OH)6]3-. It also reacts with acids (hydrogen ions) to give
[Cr(H2O)6]3+.
This is a good example of amphoteric behaviour.
Other examples of amphoteric hydroxides are zinc hydroxide and
aluminium hydroxide.
Copper(II) hydroxide as a basic oxide
Quite a lot of metal hydroxides won't react any further with hydroxide ions
if you use sodium hydroxide solution at the sort of concentrations
normally used in the lab. That means that they don't have any significant
acidic nature.
"Copper(II) hydroxide" is what we would normally call the neutral
complex [Cu(H2O)4(OH)2]. This doesn't dissolve in sodium hydroxide
solution at any concentration normally used in the lab. It doesn't show
any acidic character.
On the other hand, it will react with acids - replacing the lost hydrogen
ions on the water ligands. Because it is accepting hydrogen ions, it is
acting as a base.
Hydroxides like this (which react with acids, but not bases) are not
amphoteric - they are just simple bases.

http://www.chemguide.co.uk/inorganic/complexions/aquaoh.html (15 of 16)30/12/2004 11:33:25

reactions of aqua ions with hydroxide ions

Where would you like to go now?


To the complex ion menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/inorganic/complexions/aquaoh.html (16 of 16)30/12/2004 11:33:25

cobalt

COBALT

This page looks at some aspects of cobalt chemistry required for UK A'
level exams. These are mainly summarised from elsewhere on the site,
with links available to more detailed explanations.

Reactions of cobalt(II) ions in solution


The simplest ion that cobalt forms in solution is the pink hexaaquacobalt
(II) ion - [Co(H2O)6]2+.

Note: If you aren't happy about complex ions (including the


way they are bonded and named), it would pay you to follow
this link and explore the first couple of pages in the complex
ions menu before you go on.
Use the BACK button on your browser to return to this page.

The reaction of hexaaquacobalt(II) ions with hydroxide ions


Hydroxide ions (from, say, sodium hydroxide solution) remove hydrogen
ions from the water ligands attached to the cobalt ion.
Once a hydrogen ion has been removed from two of the water molecules,
you are left with a complex with no charge - a neutral complex. This is
insoluble in water and a precipitate is formed.

http://www.chemguide.co.uk/inorganic/transition/cobalt.html (1 of 8)30/12/2004 11:33:35

cobalt

Note: The colour coding is to show that this isn't a ligand


exchange reaction. The oxygens which were originally
attached to the cobalt are still attached in the neutral complex.

In the test-tube, the colour changes are:

Note: I don't know the reason for the change in colour. One
reliable source quotes it vaguely as a change in coordination
of the metal ion. Another suggests that the blue precipitate
involves a chloride ion (from the cobalt(II) chloride solution if
that's what you used) as a ligand. You are unlikely to need to
worry about this at this level.
You will find the reactions between hexaaqua ions and
hydroxide ions discussed in detail if you follow this link.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/inorganic/transition/cobalt.html (2 of 8)30/12/2004 11:33:35

cobalt

Reactions of hexaaquacobalt(II) ions with ammonia solution


The ammonia acts as both a base and a ligand. With a small amount of
ammonia, hydrogen ions are pulled off the hexaaqua ion exactly as in the
hydroxide ion case to give the same neutral complex.

That precipitate dissolves if you add an excess of ammonia. The


ammonia replaces water as a ligand to give hexaamminecobalt(II) ions.

Note: You might wonder why this second equation is given


starting from the original hexaaqua ion rather than the neutral
complex. Explaining why the precipitate redissolves is quite
complicated. You will find the explanation in full (although by
reference to the corresponding copper case) on the page
about the reactions between hexaaqua ions and ammonia
solution.
Use the BACK button on your browser to return to this page.

The colour changes are:

http://www.chemguide.co.uk/inorganic/transition/cobalt.html (3 of 8)30/12/2004 11:33:35

cobalt

Note: With ammonia, the hydroxide precipitate is usually


produced as a different colour from the one with sodium
hydroxide solution (itself happening in two colour variations!).
Obviously there is some re-complexing, but I don't know what
it is.

The hexaamminecobalt(II) complex is very easily oxidised to the


corresponding cobalt(III) complex. In the test-tube this is seen as a rapid
darkening to a deep red-brown solution.
In fact the hexaamminecobalt(III) ion is yellow! What you see is a mixture
of this ion and various other cobalt(III) ions involving ligand exchange
reactions with both water molecules and negative ions present in the
solution.

Another oxidation of hexaaquacobalt(II) ions


Rather than relying on oxidation by the air, you can add an oxidising
agent such as hydrogen peroxide. You can do this after the addition of
ammonia as in the last case, or you can do it following addition of sodium
hydroxide solution.
With ammonia solution and hydrogen peroxide
http://www.chemguide.co.uk/inorganic/transition/cobalt.html (4 of 8)30/12/2004 11:33:35

cobalt

The reaction with ammonia solution followed by hydrogen peroxide


produces the same dark reddish-brown solution as before - only faster.
The equation for the oxidation of the ammine complex is:

Note: This equation is worked out in detail on one of the


pages about writing ionic equations for redox reactions.
Use the BACK button on your browser to return to this page.

With sodium hydroxide solution and hydrogen peroxide


You get the variably coloured precipitate of the cobalt(II) hydroxide
complex when you add the sodium hydroxide solution.
Addition of hydrogen peroxide produces lots of bubbles of oxygen and a
dark chocolate brown precipitate.

The final precipitate contains cobalt in the +3 oxidation state. I haven't


given a formula for this, because I don't know exactly what it is! I can't
give an equation for the same reason.

http://www.chemguide.co.uk/inorganic/transition/cobalt.html (5 of 8)30/12/2004 11:33:35

cobalt

Note: One reliable source describes the precipitate as a


"hydrous Co2O3" - in other words as cobalt(III) oxide with
closely associated water. Another suggests a formula (in a
similar, but not identical context) of CoO(OH). An internet
search (to within the limits of my patience!) came up with
nothing useful.
The oxygen seen in the reaction is produced from the
decomposition of the hydrogen peroxide in a side reaction.
Many things catalyse this decomposition - presumably, in this
case, one or more of the various cobalt compounds present.

The reaction of hexaaquacobalt(II) ions with carbonate ions


You simply get a precipitate of what you can think of as cobalt(II)
carbonate.

http://www.chemguide.co.uk/inorganic/transition/cobalt.html (6 of 8)30/12/2004 11:33:35

cobalt

Note: Only one of the UK A' level Exam Boards wants this,
and this is the simplification that they make. In fact, the
precipitate is better described as a basic carbonate with a
formula of the type xCoCO3,yCo(OH)2,zH2O.
You will find the reactions between hexaaqua ions and
carbonate ions discussed in detail if you follow this link.
Use the BACK button on your browser to return to this page.

A ligand exchange reaction involving chloride ions


If you add concentrated hydrochloric acid to a solution containing
hexaaquacobalt(II) ions, the solution turns from its original pink colour to
a rich blue. The six water molecules are replaced by four chloride ions.

The reaction taking place is reversible.

If you add water to the blue solution, it returns to the pink colour.

http://www.chemguide.co.uk/inorganic/transition/cobalt.html (7 of 8)30/12/2004 11:33:35

cobalt

Note: You will find this reaction explored in more detail on a


page about ligand exchange reactions.
Use the BACK button on your browser to return to this page.

Where would you like to go now?


To the transition metal menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/inorganic/transition/cobalt.html (8 of 8)30/12/2004 11:33:35

Writing ionic equations for redox reactions done under alkaline conditions

WRITING IONIC EQUATIONS FOR REDOX


REACTIONS DONE UNDER ALKALINE
CONDITIONS

This page explains how to work out electron-half-reactions for oxidation


and reduction processes which are carried out under alkaline conditions,
and then how to combine them to give the overall ionic equation for the
redox reaction. Combining them is easy; working them out may be more
difficult than under acidic conditions.

Warning! If you have come to this page directly via a search


engine you should be aware that it follows-on from an
introductory page on this technique . You would find it much
easier to understand the current page if you read that page
first.

Why is it more difficult to write electron-half-equations for these


reactions?
What you already know
When you are trying to balance electron-half-equations, you are only
allowed to add:

electrons

water

hydrogen ions (unless the reaction is being done under alkaline


conditions - in which case, you can add hydroxide ions instead)

If you are working under acidic or neutral conditions, the sequence is


usually:
http://www.chemguide.co.uk/inorganic/redox/equations2.html (1 of 10)30/12/2004 11:33:54

Writing ionic equations for redox reactions done under alkaline conditions

Balance the atoms apart from oxygen and hydrogen.

Balance the oxygens by adding water molecules.

Balance the hydrogens by adding hydrogen ions.

Balance the charges by adding electrons.

The whole process is fairly automatic, and provided you take care, there
isn't much to go wrong.
How is this different under alkaline conditions?
The problem is that the water and the hydroxide ions that you add to
balance the equations under alkaline conditions contain both hydrogen
and oxygen.
To balance the oxygens, you could in principle add either H2O or OH- to
the equation. The same thing is true for balancing the hydrogens. How do
you know what to start with?

How to tackle the problem


In some cases, it is obvious how to build up the half-equation using
hydroxide ions. Always check this before you get involved in anything
more difficult. You will see what I mean shortly.
If it isn't immediately obvious, work out the electron-half equation as if it
were being done under acidic conditions just as you have learnt to do
on the previous page - in other words by writing in water molecules,
hydrogen ions and electrons.
Once you have got a balanced half-equation, you then convert it to
alkaline conditions. You will see how to do that in the following examples.

Four examples
http://www.chemguide.co.uk/inorganic/redox/equations2.html (2 of 10)30/12/2004 11:33:54

Writing ionic equations for redox reactions done under alkaline conditions

Don't worry if the chemistry in these examples is unfamiliar to you. It


doesn't matter in the slightest. All that matters is how you work out the
equations.
The oxidation of cobalt(II) to cobalt(III) by hydrogen peroxide
If you add an excess of ammonia solution to a solution containing cobalt
(II) ions, you get a complex ion formed called the hexaamminecobalt(II)
ion, Co(NH3)62+. This is oxidised rapidly by hydrogen peroxide solution to
the hexaamminecobalt(III) ion, Co(NH3)63+.
Ammonia solution is, of course, alkaline.
The half-equation for the cobalt reaction is easy. Start by writing down
what you know (or are told):

Everything balances apart from the charges. Add an electron to the righthand side to give both sides an overall charge of 2+.

The hydrogen peroxide half-equation isn't very difficult either, except that
you aren't told what is formed and so have to make a guess. It would
balance very nicely if you ended up with 2 hydroxide ions on the righthand side.
This is a good example of a case where it is fairly obvious where to put
hydroxide ions.

You would then just have to add 2 electrons to the left-hand side to
balance the charges.

http://www.chemguide.co.uk/inorganic/redox/equations2.html (3 of 10)30/12/2004 11:33:54

Writing ionic equations for redox reactions done under alkaline conditions

Combining the half-reactions to make the ionic equation for the reaction
What we have so far is:

The multiplication and addition looks like this:

And that's it - an easy example!

The oxidation of iron(II) hydroxide by the air


If you add sodium hydroxide solution to a solution of an iron(II) compound
you get a green precipitate of iron(II) hydroxide. This is quite quickly
oxidised by oxygen in the air to an orange-brown precipitate of iron(III)
hydroxide.
The half-equation for the iron(II) hydroxide is straightforward. Start with
what you know:

You obviously need another hydroxide ion on the left-hand side. This is
even more straightforward than the previous example.

To balance the charges, add an electron to the right-hand side.

http://www.chemguide.co.uk/inorganic/redox/equations2.html (4 of 10)30/12/2004 11:33:54

Writing ionic equations for redox reactions done under alkaline conditions

The half-equation for the oxygen isn't so easy. You don't know what is
being formed.

It isn't at all obvious whether you need to balance the oxygens by adding
water molecules or hydroxide ions to the right-hand side. OK - treat it as if
it were being done under acidic conditions, and the problem disappears!
In this case, you can only balance the oxygens by adding water
molecules to the right-hand side.

Balance the hydrogens by adding hydrogen ions to the left-hand side.

And then balance the charges by adding 4 electrons:

Now you have got a perfectly balanced half-equation. The problem is, of
course, that it only applies under acidic conditions. We should have
alkaline conditions - with hydroxide ions present not hydrogen ions.
So . . . get rid of the hydrogen ions! Add enough hydroxide ions to both
sides of the equation so that you can neutralise all the hydrogen ions.
Because it is now a balanced equation, you must add the same number
to both sides to maintain the balance.

The hydrogen ions and hydroxide ions on the left-hand side would turn
into 4 water molecules:

Finally, there are water molecules on both sides of the equation. Cancel
out any which aren't changed.
http://www.chemguide.co.uk/inorganic/redox/equations2.html (5 of 10)30/12/2004 11:33:54

Writing ionic equations for redox reactions done under alkaline conditions

This has all been a bit tedious - although you haven't actually had to think
very much! Don't forget to re-check that everything balances now that you
have finished.

Combining the half-reactions to make the ionic equation for the reaction
From here on it's all back to the usual routine. We've worked out the two
half-equations:

The iron equation will have to happen 4 times to supply enough electrons
to the oxygen.

Notice that the hydroxide ions on each side cancel out. (Perhaps to your
surprise - certainly to mine when I worked this out!)

The reduction of manganate(VII) ions to manganate(VI) ions by


hydroxide ions
This is a fairly obscure reaction but it is not too difficult to work out and
balance. It is unusual in that hydroxide ions are acting as reducing agents.
Dark purple potassium manganate(VII) solution is slowly reduced to dark
http://www.chemguide.co.uk/inorganic/redox/equations2.html (6 of 10)30/12/2004 11:33:54

Writing ionic equations for redox reactions done under alkaline conditions

green potassium manganate(VI) solution by concentrated potassium


hydroxide solution. Bubbles of oxygen gas are also given off.
Note: Unless the potassium manganate(VII) solution is very
dilute, its strong purple colour tends to mask the green of the
potassium manganate(VI) in the short term.

The half-equation for the conversion of manganate(VII) ions to manganate


(VI) ions is easy (provided, of course, that you know their formulae!).

So what is happening to the hydroxide ions to turn them into oxygen gas?
It isn't actually very difficult to work out the half-equation directly, but
suppose you want to avoid thinking and go through a standard routine:
Write down what you know, balancing the oxygens in the process:

Balance the hydrogens by adding hydrogen ions:

. . . and now balance the charges:

Now get rid of the hydrogen ions by adding enough hydroxide ions to
both sides of the equation:

. . . and tidy it all up!

http://www.chemguide.co.uk/inorganic/redox/equations2.html (7 of 10)30/12/2004 11:33:54

Writing ionic equations for redox reactions done under alkaline conditions

Combining the half-reactions to make the ionic equation for the reaction
What we have so far is:

The manganese reaction will have to happen four times in order to use up
the four electrons produced by the hydroxide equation. Putting this
together, you get:

The chemistry may be unfamiliar, but working out the equation isn't too
hard!

The oxidation of chromium(III) to chromium(VI)


If you add an excess of sodium hydroxide solution to a solution containing
chromium(III) ions, you get a dark green solution containing the complex
ion hexahydroxochromate(III), Cr(OH)63-.
This can be oxidised to bright yellow chromate(VI) ions, CrO42- by
warming it with hydrogen peroxide solution.
We've already worked out the hydrogen peroxide half-equation where it is
acting as an oxidising agent under alkaline conditions:

So this time we just need to put together the half-equation for the
chromium ions. What we know is:
http://www.chemguide.co.uk/inorganic/redox/equations2.html (8 of 10)30/12/2004 11:33:54

Writing ionic equations for redox reactions done under alkaline conditions

It isn't the least bit obvious where to put hydroxide ions or water
molecules, so treat it as if it were being done under acidic conditions.
That way, you start by balancing the oxygens by adding water molecules.
To get 6 oxygens on each side, you need two waters on the right-hand
side:

Now balance the hydrogens by adding hydrogen ions:

. . . and then balance the charges by adding electrons:

Finally, convert from acidic to alkaline conditions by adding enough


hydroxide ions to both sides to turn the hydrogen ions into water:

. . . and tidy it all up:

Combining the half-reactions to make the ionic equation for the reaction
The two half-equations are:

If you multiply one equation by 3 and the other by 2, that transfers a total
of 6 electrons.

http://www.chemguide.co.uk/inorganic/redox/equations2.html (9 of 10)30/12/2004 11:33:54

Writing ionic equations for redox reactions done under alkaline conditions

Finally, tidy up the hydroxide ions that occur on both sides to leave the
overall ionic equation:

Where would you like to go now?


Return to the first page on electron-half-equations . . .
To the Redox menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/inorganic/redox/equations2.html (10 of 10)30/12/2004 11:33:54

reactions of aqua ions with carbonate ions

REACTIONS OF HEXAAQUA METAL IONS WITH


CARBONATE IONS

This page describes and explains the reactions between complex ions of
the type [M(H2O)6]n+ and carbonate ions from, for example, sodium
carbonate solution.

Important: Before you read this page, you need to know


about the acidity of the hexaaqua ions.
It would also be useful to read the page describing the
reactions of these ions with hydroxide ions.
You will find the present page easier to understand if you
spend some time reading these other two pages first.
It will probably be easiest to return to this page via the
complex ions menu.

The general cases


There is a difference in the reactions depending on whether the metal at
the centre of the hexaaqua ion carries a 2+ or a 3+ charge. We need to
look at the two cases separately.
3+ ions reacting with carbonate ions
The 3+ hexaaqua ions are sufficiently acidic to react with carbonate ions
to release carbon dioxide gas. You also get a precipitate of the metal
hydroxide.
This is how it happens . . .
http://www.chemguide.co.uk/inorganic/complexions/aquaco3.html (1 of 8)30/12/2004 11:34:09

reactions of aqua ions with carbonate ions

The acidity of the 3+ hexaaqua ions


These have the formula [M(H2O)6]3+, and they are fairly acidic. They
react with water molecules from the solution:

They are acting as acids by donating hydrogen ions to water molecules


in the solution.
Because of the confusing presence of water from two different sources
(the ligands and the solution), it is easier to simplify this:

Carbonate ions acting as a base


Carbonate ions combine with hydrogen ions in two stages - first to make
hydrogencarbonate ions, and then to give carbon dioxide and water.

Reacting carbonate ions with the 3+ hexaaqua ions


Provided the proportions are right, the 3+ hexaaqua ions are sufficiently
acidic for the reactions to go all the way to carbon dioxide.
There are two possible reactions.
Reaction of carbonate ions with the hydroxonium ions (hydrogen ions)
According to Le Chatelier's Principle, as the hydrogen ions are removed,
the position of equilibrium will move to the right, producing more of the
http://www.chemguide.co.uk/inorganic/complexions/aquaco3.html (2 of 8)30/12/2004 11:34:09

reactions of aqua ions with carbonate ions

new complex ion.

Reaction of carbonate ions with the hexaaqua ion


Statistically, there is far more chance of a carbonate ion hitting a
hexaaqua metal ion than of hitting a hydrogen ion. It removes a hydrogen
ion directly from the hexaaqua complex.
If that happens, you get exactly the same new complex ion formed as
above.
But it doesn't stop there. Carbonate ions keep on removing hydrogen
ions from the complex until you end up with a neutral complex. You see
that as a precipitate of the metal hydroxide.

http://www.chemguide.co.uk/inorganic/complexions/aquaco3.html (3 of 8)30/12/2004 11:34:09

reactions of aqua ions with carbonate ions

You can also usefully write the complete change as an overall reaction.

2+ ions reacting with carbonate ions


The 2+ hexaaqua ions aren't strongly acidic enough to release carbon
dioxide from carbonates. In these cases, you still get a precipitate - but it
is a precipitate of what is loosely described as the "metal carbonate".

Note: Only one of the UK A' level Exam Boards wants this,
and this is the simplification that they make. In fact, the
precipitates are better described as basic carbonates with a
formula of the type xMCO3,yM(OH)2,zH2O.

http://www.chemguide.co.uk/inorganic/complexions/aquaco3.html (4 of 8)30/12/2004 11:34:09

reactions of aqua ions with carbonate ions

Looking at the ions of specific metals


We'll look at the reactions of three 3+ ions and three 2+ ions. The
important ones are the 3+ ions.
3+ ions
hexaaquaaluminium

Starting from a colourless solution, you get a white precipitate - but with
bubbles of gas as well. The precipitate is identical to the one you get if
you add small amounts of either sodium hydroxide or ammonia solutions
to a solution of the hexaaquaaluminium ions.
hexaaquachromium(III)

Again, the precipitate is just the same as if you had added small amounts
of either sodium hydroxide or ammonia solution.

http://www.chemguide.co.uk/inorganic/complexions/aquaco3.html (5 of 8)30/12/2004 11:34:09

reactions of aqua ions with carbonate ions

hexaaquairon(III)

. . . and again, exactly the same precipitate as if you had added any
other base.

Summary
In each case you get a precipitate of the neutral complex - the metal
hydroxide. This is exactly the same precipitate that you get if you add
small amounts of either sodium hydroxide solution or ammonia solution
to solutions of these ions. Bubbles of carbon dioxide are also given off.

Note: You may miss this carbon dioxide if your proportions


are wrong. Unless there is an excess of the acidic hexaaqua
ion, you may get hydrogencarbonate ions formed in solution
instead of carbon dioxide.

http://www.chemguide.co.uk/inorganic/complexions/aquaco3.html (6 of 8)30/12/2004 11:34:09

reactions of aqua ions with carbonate ions

2+ ions
hexaaquacobalt(II)

No gas this time - just a precipitate of "cobalt(II) carbonate".


hexaaquacopper(II)

Again, there isn't any carbon dioxide - just a precipitate of the "copper(II)
carbonate".
hexaaquairon(II)

http://www.chemguide.co.uk/inorganic/complexions/aquaco3.html (7 of 8)30/12/2004 11:34:09

reactions of aqua ions with carbonate ions

You get a precipitate of the "iron(II) carbonate", but no carbon dioxide.

Summary
Hexaaqua ions with a 2+ charge aren't sufficiently acidic to liberate
carbon dioxide from carbonate ions. Instead you get a precipitate which
you can think of as being the metal carbonate.

Where would you like to go now?


To the complex ion menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/inorganic/complexions/aquaco3.html (8 of 8)30/12/2004 11:34:09

transition metals menu

Understanding Chemistry

TRANSITION METALS MENU

General features of transition metals . . .


This page describes the general features of transition metal
chemistry, and provides links to other pages on the site where
some of the material is covered in more detail.

The chemistry of some specific transition metals


These pages bring together some detailed chemistry of each of
the metals required by UK A' level syllabuses.
Some of the material is taken directly from other pages on the site,
but with added new bits - mainly on redox reactions involving the
ions.
vanadium . . .
chromium . . .
manganese . . .
iron . . .
cobalt . . .
copper . . .

http://www.chemguide.co.uk/inorganic/transitionmenu.html (1 of 2)30/12/2004 11:34:11

transition metals menu

The chemistry of some complex ions


If you are looking particulary for the chemistry of complex ions
(including transition metal ions), you may be better off exploring
the complex ions menu. Follow the next link below.
Go to complex ions menu . . .

Go to inorganic chemistry menu . . .


Go to Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/inorganic/transitionmenu.html (2 of 2)30/12/2004 11:34:11

introducing transition metals

THE GENERAL FEATURES OF TRANSITION


METAL CHEMISTRY

This page explains what a transition metal is in terms of its electronic


structure, and then goes on to look at the general features of transition
metal chemistry. These include variable oxidation state (oxidation
number), complex ion formation, coloured ions, and catalytic activity.
You will find some of this covered quite briefly on this page with links to
other parts of the site where the topics are covered in more detail.

The electronic structures of transition metals


What is a transition metal?
The terms transition metal (or element) and d block element are
sometimes used as if they mean the same thing. They don't - there's a
subtle difference between the two terms.
We'll explore d block elements first:
d block elements
You will remember that when you are building the Periodic Table and
working out where to put the electrons, something odd happens after
argon.
At argon, the 3s and 3p levels are full, but rather than fill up the 3d levels
next, the 4s level fills instead to give potassium and then calcium.
Only after that do the 3d levels fill.

http://www.chemguide.co.uk/inorganic/transition/features.html (1 of 15)30/12/2004 11:34:17

introducing transition metals

Note: If you aren't sure about atomic orbitals and electronic


structures, you really need to follow this link before you go on.
It takes you to a page explaining atomic orbitals and then on to
other pages about electronic structures.
If you do follow the link, use the BACK button on your browser
(or the History file or Go menu) to return quickly to this page.

The elements in the Periodic Table which correspond to the d levels filling
are called d block elements. The first row of these is shown in the
shortened form of the Periodic Table below.

The electronic structures of the d block elements shown are:

Sc

[Ar] 3d14s2

Ti

[Ar] 3d24s2

[Ar] 3d34s2

Cr

[Ar] 3d54s1

Mn

[Ar] 3d54s2

http://www.chemguide.co.uk/inorganic/transition/features.html (2 of 15)30/12/2004 11:34:17

introducing transition metals

Fe

[Ar] 3d64s2

Co

[Ar] 3d74s2

Ni

[Ar] 3d84s2

Cu

[Ar] 3d104s1

Zn

[Ar] 3d104s2

You will notice that the pattern of filling isn't entirely tidy! It is broken at
both chromium and copper.

Note: This is something that you are just going to have to


accept. I know of no simple explanation for it. (In fact, to be
honest, I don't even know a complicated explanation for it!)
People sometimes say that a half-filled d level (with one
electron in each orbital) is especially stable - but they never
say why that might be. Neither can you use the statement that
a full d level (for example, in the copper case) is stable, unless
you can also explain why that is. Again, nobody ever does.
If you can't explain something, it is much better just to accept it
than to make up faulty explanations which sound OK on the
surface but don't stand up to scrutiny!

http://www.chemguide.co.uk/inorganic/transition/features.html (3 of 15)30/12/2004 11:34:17

introducing transition metals

Transition metals
Not all d block elements count as transition metals! There are
discrepancies between the various UK A' level syllabuses, but the majority
use the definition:

A transition metal is one which forms one or more stable ions which
have incompletely filled d orbitals.

Note: One of the UK A' level syllabuses includes the


possibility of the metal itself having incompletely filled d levels.
This is unlikely to be a big problem (it only really arises with
scandium), but it would pay you to learn the version your
syllabus wants.
If you haven't got a copy of your syllabus, follow this link to find
out how to get one. Use the BACK button on your browser to
return quickly to this page.

On the basis of this definition, scandium and zinc don't count as transition
metals - even though they are members of the d block.
Scandium has the electronic structure [Ar] 3d14s2. When it forms ions, it
always loses the 3 outer electrons and ends up with an argon structure.
The Sc3+ ion has no d electrons and so doesn't meet the definition.
Zinc has the electronic structure [Ar] 3d104s2. When it forms ions, it always
loses the two 4s electrons to give a 2+ ion with the electronic structure [Ar]
3d10. The zinc ion has full d levels and doesn't meet the definition either.
By contrast, copper, [Ar] 3d104s1, forms two ions. In the Cu+ ion the
electronic structure is [Ar] 3d10. However, the more common Cu2+ ion has
the structure [Ar] 3d9.
Copper is definitely a transition metal because the Cu2+ ion has an
incomplete d level.
http://www.chemguide.co.uk/inorganic/transition/features.html (4 of 15)30/12/2004 11:34:17

introducing transition metals

Transition metal ions


You have already come across the fact that when the Periodic Table is
being built, the 4s orbital is filled before the 3d orbitals. This is because in
the empty atom, 4s orbitals have a lower energy than 3d orbitals.
However, once the electrons are actually in their orbitals, the energy order
changes - and in all the chemistry of the transition elements, the 4s orbital
behaves as the outermost, highest energy orbital.
The reversed order of the 3d and 4s orbitals only applies to building the
atom up in the first place. In all other respects, you treat the 4s electrons
as being the outer electrons.

Note: I have never seen a simple explanation for this


annoying discrepancy. It is another of those things that you just
have to accept. Just remember that once you have the full
electronic structure for one of these atoms, the 4s electrons are
the outermost electrons.

Remember this:

When d-block elements form ions, the 4s electrons are lost first.

To write the electronic structure for Co2+:

Co

[Ar] 3d74s2

Co2+

[Ar] 3d7

The 2+ ion is formed by the loss of the two 4s electrons.

http://www.chemguide.co.uk/inorganic/transition/features.html (5 of 15)30/12/2004 11:34:17

introducing transition metals

To write the electronic structure for V3+:

[Ar] 3d34s2

V3+

[Ar] 3d2

The 4s electrons are lost first followed by one of the 3d electrons.

Note: You will find more examples of writing the electronic


structures for d block ions, by following this link.
Use the BACK button on your browser to return quickly to this
page.

Variable oxidation state (number)


One of the key features of transition metal chemistry is the wide range of
oxidation states (oxidation numbers) that the metals can show.

Note: If you aren't sure about oxidation states, you really need
to follow this link before you go on.
Use the BACK button on your browser to return quickly to this
page.

http://www.chemguide.co.uk/inorganic/transition/features.html (6 of 15)30/12/2004 11:34:17

introducing transition metals

It would be wrong, though, to give the impression that only transition


metals can have variable oxidation states. For example, elements like
sulphur or nitrogen or chlorine have a very wide range of oxidation states
in their compounds - and these obviously aren't transition metals.
However, this variability is less common in metals apart from the transition
elements. Of the familiar metals from the main groups of the Periodic
Table, only lead and tin show variable oxidation state to any extent.
Examples of variable oxidation states in the transition metals
Iron
Iron has two common oxidation states (+2 and +3) in, for example, Fe2+
and Fe3+. It also has a less common +6 oxidation state in the ferrate(VI)
ion, FeO42-.
Manganese
Manganese has a very wide range of oxidation states in its compounds.
For example:

+2

in Mn2+

+3

in Mn2O3

+4

in MnO2

+6

in MnO42-

+7

in MnO4-

Other examples
You will find the above examples and others looked at in detail if you
explore the chemistry of individual metals from the transition metal menu.
There is a link to this menu at the bottom of the page.
http://www.chemguide.co.uk/inorganic/transition/features.html (7 of 15)30/12/2004 11:34:17

introducing transition metals

Explaining the variable oxidation states in the transition metals


We'll look at the formation of simple ions like Fe2+ and Fe3+.
When a metal forms an ionic compound, the formula of the compound
produced depends on the energetics of the process. On the whole, the
compound formed is the one in which most energy is released. The more
energy released, the more stable the compound.
There are several energy terms to think about, but the key ones are:

The amount of energy needed to ionise the metal (the sum of the
various ionisation energies)
The amount of energy released when the compound forms. This will
either be lattice enthalpy if you are thinking about solids, or the
hydration enthalpies of the ions if you are thinking about solutions.

The more highly charged the ion, the more electrons you have to remove
and the more ionisation energy you will have to provide.
But off-setting this, the more highly charged the ion, the more energy is
released either as lattice enthalpy or the hydration enthalpy of the metal
ion.

Note: What I am talking about here in a general way are BornHaber cycles. These aren't covered anywhere on this site, but
if you want to explore them in more detail, you may be
interested in my chemistry calculations book.

http://www.chemguide.co.uk/inorganic/transition/features.html (8 of 15)30/12/2004 11:34:17

introducing transition metals

Thinking about a typical non-transition metal (calcium)


Calcium chloride is CaCl2. Why is that?
If you tried to make CaCl, (containing a Ca+ ion), the overall process is
slightly exothermic.
By making a Ca2+ ion instead, you have to supply more ionisation energy,
but you get out lots more lattice energy. There is much more attraction
between chloride ions and Ca2+ ions than there is if you only have a 1+
ion. The overall process is very exothermic.
Because the formation of CaCl2 releases much more energy than making
CaCl, then CaCl2 is more stable - and so forms instead.
What about CaCl3? This time you have to remove yet another electron
from calcium.
The first two come from the 4s level. The third one comes from the 3p.
That is much closer to the nucleus and therefore much more difficult to
remove. There is a large jump in ionisation energy between the second
and third electron removed.
Although there will be a gain in lattice enthalpy, it isn't anything like enough
to compensate for the extra ionisation energy, and the overall process is
very endothermic.
It definitely isn't energetically sensible to make CaCl3!
Thinking about a typical transition metal (iron)
Here are the changes in the electronic structure of iron to make the 2+ or
the 3+ ion.

Fe

[Ar] 3d64s2

Fe2+

[Ar] 3d6

http://www.chemguide.co.uk/inorganic/transition/features.html (9 of 15)30/12/2004 11:34:17

introducing transition metals

Fe3+

[Ar] 3d5

The 4s orbital and the 3d orbitals have very similar energies. There isn't a
huge jump in the amount of energy you need to remove the third electron
compared with the first and second.
The figures for the first three ionisation energies (in kJ mol-1) for iron
compared with those of calcium are:

metal

1st IE

2nd IE

3rd IE

Ca

590

1150

4940

Fe

762

1560

2960

There is an increase in ionisation energy as you take more electrons off an


atom because you have the same number of protons attracting fewer
electrons. However, there is much less increase when you take the third
electron from iron than from calcium.
In the iron case, the extra ionisation energy is compensated more or less
by the extra lattice enthalpy or hydration enthalpy evolved when the 3+
compound is made.
The net effect of all this is that the overall enthalpy change isn't vastly
different whether you make, say, FeCl2 or FeCl3. That means that it isn't
too difficult to convert between the two compounds.

The formation of complex ions


What is a complex ion?
A complex ion has a metal ion at its centre with a number of other
molecules or ions surrounding it. These can be considered to be attached
to the central ion by co-ordinate (dative covalent) bonds. (In some cases,

http://www.chemguide.co.uk/inorganic/transition/features.html (10 of 15)30/12/2004 11:34:17

introducing transition metals

the bonding is actually more complicated than that.)


The molecules or ions surrounding the central metal ion are called ligands.
Simple ligands include water, ammonia and chloride ions.

What all these have got in common is active lone pairs of electrons in the
outer energy level. These are used to form co-ordinate bonds with the
metal ion.
Some examples of complex ions formed by transition metals
[Fe(H2O)6]2+
[Co(NH3)6]2+
[Cr(OH)6]3[CuCl4]2Other metals also form complex ions - it isn't something that only transition
metals do. Transition metals do, however, form a very wide range of
complex ions.

http://www.chemguide.co.uk/inorganic/transition/features.html (11 of 15)30/12/2004 11:34:17

introducing transition metals

Note: You will find much more about complex ions by


following this link. It will take you to a part of the site dealing
exclusively with complex ions.
If you do follow the link, use the BACK button on your browser
(or the History file or Go menu) if you want to return to this
page again.

The formation of coloured compounds


Some common examples
The diagrams show aproximate colours for some common transition metal
complex ions.

http://www.chemguide.co.uk/inorganic/transition/features.html (12 of 15)30/12/2004 11:34:17

introducing transition metals

You will find these and others discussed if you follow links to individual
metals from the transition metal menu (link at the bottom of the page).
Alternatively, you could explore the complex ions menu (follow the link in
the help box which has just disappeared off the top of the screen).
The origin of colour in the transition metal ions
When white light passes through a solution of one of these ions, or is
reflected off it, some colours in the light are absorbed. The colour you see
is how your eye perceives what is left.
Attaching ligands to a metal ion has an effect on the energies of the d
orbitals. Light is absorbed as electrons move between one d orbital and
another. This is explained in detail on another page.

Note: You will find a detailed explanation of the origin of


colour in complex ions and the factors which cause it to change
by following this link. That page is on the part of the site
dealing with complex ions.
Use the BACK button on your browser if you want to return to
this page again.

Catalytic activity
Transition metals and their compounds are often good catalysts. A few of
the more obvious cases are mentioned below, but you will find catalysis
explored in detail elsewhere on the site (follow the link after the examples).
Transition metals and their compounds function as catalysts either
because of their ability to change oxidation state or, in the case of the
metals, to adsorb other substances on to their surface and activate them in
the process. All this is expored in the main catalysis section.

http://www.chemguide.co.uk/inorganic/transition/features.html (13 of 15)30/12/2004 11:34:17

introducing transition metals

Transition metals as catalysts


Iron in the Haber Process
The Haber Process combines hydrogen and nitrogen to make ammonia
using an iron catalyst.

Nickel in the hydrogenation of C=C bonds


This reaction is at the heart of the manufacture of margarine from
vegetable oils.
However, the simplest example is the reaction between ethene and
hydrogen in the presence of a nickel catalyst.

Transition metal compounds as catalysts


Vanadium(V) oxide in the Contact Process
At the heart of the Contact Process is a reaction which converts sulphur
dioxide into sulphur trioxide. Sulphur dioxide gas is passed together with
air (as a source of oxygen) over a solid vanadium(V) oxide catalyst.

Iron ions in the reaction between persulphate ions and iodide ions
Persulphate ions (peroxodisulphate ions), S2O82-, are very powerful
oxidising agents. Iodide ions are very easily oxidised to iodine. And yet the
reaction between them in solution in water is very slow.
The reaction is catalysed by the presence of either iron(II) or iron(III) ions.

http://www.chemguide.co.uk/inorganic/transition/features.html (14 of 15)30/12/2004 11:34:17

introducing transition metals

Note: You will find detailed explanations of these reactions in


the catalysis section of the site. You could usefully start with
the types of catalysts page.
Use the BACK button on your browser to return quickly to this
page.

Where would you like to go now?


To the transition metal menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/inorganic/transition/features.html (15 of 15)30/12/2004 11:34:17

vanadium

VANADIUM

This page looks at some aspects of vanadium chemistry required for UK


A' level exams. That includes the use of vanadium(V) oxide as a catalyst
in the Contact Process, and converting between the various vanadium
oxidation states.

Vanadium(V) oxide as a catalyst in the Contact Process


The overall reaction
During the Contact Process for manufacturing sulphuric acid, sulphur
dioxide has to be converted into sulphur trioxide. This is done by passing
sulphur dioxide and oxygen over a solid vanadium(V) oxide catalyst.

Note: The equation is written with the half in it to make the


explanation below tidier. You may well be familiar with the
equation written as twice that shown, but the present version
is perfectly acceptable. It is also shown as a one-way rather
than a reversible reaction to avoid complicating things.

How the reaction works


This is a good example of the ability of transition metals and their
compounds to act as catalysts because of their ability to change their
oxidation state (oxidation number).

http://www.chemguide.co.uk/inorganic/transition/vanadium.html (1 of 14)30/12/2004 11:34:28

vanadium

Note: If you aren't sure about oxidation states, you really


ought to follow this link before you go on. You will need to be
confident about oxidation states throughout the whole of the
present page.
Use the BACK button on your browser to return to this page.

The sulphur dioxide is oxidised to sulphur trioxide by the vanadium(V)


oxide. In the process, the vanadium(V) oxide is reduced to vanadium(IV)
oxide.

The vanadium(IV) oxide is then re-oxidised by the oxygen.

Although the catalyst has been temporarily changed during the reaction,
at the end it is chemically the same as it started.

Note: If you want more detail about the Contact Process, you
will find a full description of the conditions used and the
reasons for them by following this link.
If you want to find out more about catalysis, this link will take
you to an introductory page about catalysts with further links
to other catalysis pages if you want them.
Use the BACK button on your browser (or the History file of
Go menu) if you want to return to this page later.

http://www.chemguide.co.uk/inorganic/transition/vanadium.html (2 of 14)30/12/2004 11:34:28

vanadium

Vanadium's oxidation states


Vanadium has oxidation states in its compounds of +5, +4, +3 and +2.
This section looks at ways of changing between them. It starts with a bit of
description, and then goes on to look at the reactions in terms of standard
redox potentials (standard electrode potentials).
Observing the changes in the lab
Reducing vanadium(V) in stages to vanadium(II)
The usual source of vanadium in the +5 oxidation state is ammonium
metavanadate, NH4VO3. This isn't very soluble in water and is usually first
dissolved in sodium hydroxide solution.
The solution can be reduced using zinc and an acid - either hydrochloric
acid or sulphuric acid, usually using moderately concentrated acid.
The exact vanadium ion present in the solution is very complicated, and
varies with the pH of the solution. The reaction is done under acidic
conditions when the main ion present is VO2+ - called the dioxovanadium
(V) ion.

Note: The ion is usually written as VO2+, but is more


accurately [VO2(H2O)4]+.

http://www.chemguide.co.uk/inorganic/transition/vanadium.html (3 of 14)30/12/2004 11:34:28

vanadium

If you do the reaction in a small flask, it is normally stoppered with some


cotton wool. This allows hydrogen (produced from a side reaction
between the zinc and acid) to escape. At the same time it stops much air
from entering. This prevents re-oxidation of the lower oxidation states of
vanadium (particularly the +2 state) by oxygen in the air.
The reaction is usually warmed so that the changes happen in a
reasonable time.
The reduction is shown in two stages. Some individual important colours
are shown, but the process is one continuous change from start to finish.
The reduction from +5 to +4

It is important to notice that the green colour you see isn't actually another
oxidation state. it is just a mixture of the original yellow of the +5 state and
the blue of the +4.
Be very careful with the formulae of the two vanadium ions - they are very
easy to confuse!

http://www.chemguide.co.uk/inorganic/transition/vanadium.html (4 of 14)30/12/2004 11:34:28

vanadium

Note: Just like the VO2+ ion, the VO2+ ion will have water
molecules attached to it as well - [VO(H2O)5]2+. We usually
use the simpler form.

The reduction from +4 to +2


The colour changes just continue.

The reason for the inverted commas around the vanadium(III) ion is that
this is almost certainly a simplification. The exact nature of the complex
ion will depend on which acid you use in the reduction process. The
simplification is probably reasonable at this level.

Note: If you use hydrochloric acid, you get a ligand exchange


reaction to give [V(H2O)4Cl2]+. This causes the green colour in
the vanadium(III) solution.
I am uncertain about the sulphuric acid case. One source says
that with sulphuric acid, you actually get the [V(H2O)6]3+ ion
which is a dull grey-blue colour. However, when I checked this
in the lab before writing this, I got exactly the same green
colour with both acids.
One possibility is that you get another ligand exchange
reaction with sulphate ions to give [V(H2O)5(SO4)]+, but I
http://www.chemguide.co.uk/inorganic/transition/vanadium.html (5 of 14)30/12/2004 11:34:28

vanadium

haven't been able to confirm this.


There is, however, a very similar case in chromium chemistry
which is discussed on the page about ligand exchange
reactions.
If you follow this link, use the BACK button on your browser to
return to this page.

Re-oxidation of the vanadium(II)


The vanadium(II) ion is very easily oxidised. If you remove the cotton wool
from the flask and pour some solution into a test tube, it turns green
because of its contact with oxygen in the air. It is oxidised back to
vanadium(III).

Note: It only changes this quickly if the solution is still warm.


In the cold, the change is quite a lot slower.

If it is allowed to stand for a long time, the solution eventually turns blue
as the air oxidises it back to the vanadium(IV) state - VO2+ ions.
Adding nitric acid (a reasonably powerful oxidising agent) to the original
vanadium(II) solution also produces blue VO2+ ions. The vanadium(II) is
again oxidised back to vanadium(IV).

http://www.chemguide.co.uk/inorganic/transition/vanadium.html (6 of 14)30/12/2004 11:34:28

vanadium

Important! There is no point in reading the rest of this page


unless you are familiar with redox potentials (electrode
potentials).
This link will take you to a page explaining how to use redox
potentials in the present context. However, you should be
aware that it is the final page in a linked sequence. That page
will give you the opportunity to start at the beginning if you
need to. You would be strongly advised to do that unless you
are really confident! It will take you some time, but at the end
you should really understand how to use redox potentials.
If you think you know about E values, it would probably be
quicker to read the rest of this current page and then come
back to this link if you need to.
If you follow this link, use the BACK button on your browser
(or the History file or the Go menu) to return quickly to this
page.

Explaining the changes in terms of redox potentials (electrode


potentials)
Using zinc as the reducing agent
The first stage of the series of reductions
Let's look at the first stage of the reduction - from VO2+ to VO2+. The
redox potential for the vanadium half-reaction is given by:

The corresponding equilibrium for the zinc is:

http://www.chemguide.co.uk/inorganic/transition/vanadium.html (7 of 14)30/12/2004 11:34:28

vanadium

The simple principle is that if you couple two of these half-reactions


together, the one with the more positive E value will move to the
right; the one with the more negative (or less positive) E moves to
the left.

Note: If you have learnt to use E values using some more


complicated method (and there are lots of unnecessarily
complicated methods around!), it would pay you to follow the
link above to find out how to do it all much more easily.

So . . . if you mix together zinc and VO2+ ions in the presence of acid to
provide the H+ ions:

That converts the two equilibria into two one-way reactions. You can write
these down and combine them to give the ionic equation for the reaction if
you want to.

http://www.chemguide.co.uk/inorganic/transition/vanadium.html (8 of 14)30/12/2004 11:34:28

vanadium

Note: If you aren't happy about building up ionic equations


like this, you could usefully follow this link.
Use the BACK button on your browser to return quickly to this
page.

The other stages of the reaction


Here are the E values for all the steps of the reduction from vanadium(V)
to vanadium(II):

. . . and here is the zinc value again:

Remember that for the vanadium reactions to move to the right (which is
what we want), their E values must be more positive than whatever you
are reacting them with.
In other words, for the reactions to work, zinc must always have the more
negative value - and that's the case.

http://www.chemguide.co.uk/inorganic/transition/vanadium.html (9 of 14)30/12/2004 11:34:28

vanadium

Zinc can reduce the vanadium through each of these steps to give the
vanadium(II) ion.
Using other reducing agents
Suppose you replaced zinc as the reducing agent by tin. How far would
the set of reductions go this time?
Here are the E values again:

. . . and here is the tin value:

In order for each reduction to happen, the vanadium reaction has to have
the more positive E value because we want it to go to the right. That
means that the tin must have the more negative value.
In the first vanadium equation (from +5 to +4), the tin value is more
negative. That works OK.
In the second vanadium equation (from +4 to +3), the tin value is again
the more negative. That works as well.
But in the final vanadium reaction (from +3 to +2), tin no longer has the
more negative E value. Tin won't reduce vanadium(III) to vanadium(II).

http://www.chemguide.co.uk/inorganic/transition/vanadium.html (10 of 14)30/12/2004 11:34:28

vanadium

Important! If this doesn't seem pretty obvious to you, then


you don't really understand about redox potentials.
This link will take you to the redox potential menu. Spend
some time by starting at the beginning of the sequence of
pages you will find there.
If you follow this link, use the BACK button on your browser
(or the History file or the Go menu) if you want to return to this
page later.

Re-oxidation of the vanadium(II)


The vanadium(II) oxidation state is easily oxidised back to vanadium(III) or even higher.
Oxidation by hydrogen ions
You will remember that the original reduction we talked about was carried
out using zinc and an acid in a flask stoppered with a piece of cotton wool
to keep the air out. Air will rapidly oxidise the vanadium(II) ions - but so
also will the hydrogen ions present in the solution!
The vanadium(II) solution is only stable as long as you keep the air out,
and in the presence of the zinc. The zinc is necessary to keep the
vanadium reduced.
What happens if the zinc isn't there? Look at these E values:

The reaction with the more negative E value goes to the left; the reaction
with the more positive (or less negative) one to the right.
That means that the vanadium(II) ions will be oxidised to vanadium(III)

http://www.chemguide.co.uk/inorganic/transition/vanadium.html (11 of 14)30/12/2004 11:34:28

vanadium

ions, and the hydrogen ions reduced to hydrogen.


Will the oxidation go any further - for example, to the vanadium(IV) state?
Have a look at the E values and decide:

In order for the vanadium equilibrium to move to the left, it would have to
have the more negative E value. It hasn't got the more negative E value
and so the reaction doesn't happen.
Oxidation by nitric acid
In a similar sort of way, you can work out how far nitric acid will oxidise
the vanadium(II).
Here's the first step:

The vanadium reaction has the more negative E value and so will move
to the left; the nitric acid reaction moves to the right.
Nitric acid will oxidise vanadium(II) to vanadium(III).
The second stage involves these E values:

The nitric acid again has the more positive E value and so moves to the
right. The more negative (less positive) vanadium reaction moves to the
left.

http://www.chemguide.co.uk/inorganic/transition/vanadium.html (12 of 14)30/12/2004 11:34:28

vanadium

Nitric acid will certainly oxidise vanadium(III) to vanadium(IV).


Will it go all the way to vanadium(V)?

No, it won't! For the vanadium reaction to move to the left to form the
dioxovanadium(V) ion, it would have to have the more negative (less
positive) E value. It hasn't got a less positive value, and so the reaction
doesn't happen.

Note: There are several possible half-reactions involving the


nitric acid with a variety of E values. Two of these actually
have E values more positive than +1.00 and so, in principle,
nitric acid could seem to be able to oxidise vanadium(IV) to
vanadium(V) - but involving different products from the nitric
acid.
In practice, if you do this reaction in the lab, the solution turns
blue - producing the vanadium(IV) state. Just because the E
values tell you that a reaction is possible, you can't assume
that it will actually happen. There may be very large activation
energy barriers involved, causing the reaction to be infinitely
slow!

http://www.chemguide.co.uk/inorganic/transition/vanadium.html (13 of 14)30/12/2004 11:34:28

vanadium

You can work out the effect of any other oxidising agent on the lower
oxidation states of vanadium in exactly the same way. But don't assume
that because the E values show that a reaction is possible, it will
necessarily happen.

Where would you like to go now?


To the transition metal menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/inorganic/transition/vanadium.html (14 of 14)30/12/2004 11:34:28

chromium

CHROMIUM

This page looks at some aspects of chromium chemistry required for UK A' level
exams. It includes: reactions of chromium(III) ions in solution (summarised from
elsewhere on the site); the interconversion of the various oxidation states of
chromium; the chromate(VI)-dichromate(VI) equilibrium; and the use of dichromate
(VI) ions as an oxidising agent (including titrations).
The first part of this page is a summary of the reactions of chromium(III) ions in
solution. You will find links to other pages where these reactions are discussed in
more detail.

Reactions of chromium(III) ions in solution


The simplest ion that chromium forms in solution is the hexaaquachromium(III) ion
- [Cr(H2O)6]3+.

Note: If you aren't happy about complex ions (including the way they
are bonded and named), it would pay you to follow this link and explore
the first couple of pages in the complex ions menu before you go on.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/inorganic/transition/chromium.html (1 of 16)30/12/2004 11:35:13

chromium

The acidity of the hexaaqua ions


In common with the other 3+ ions, the hexaaquachromium(III) ion is fairly acidic with a pH for typical solutions in the 2 - 3 range.
The ion reacts with water molecules in the solution. A hydrogen ion is lost from
one of the ligand water molecules:

The complex ion is acting as an acid by donating a hydrogen ion to water


molecules in the solution. The water is, of course, acting as a base by accepting
the hydrogen ion.
Because of the confusing presence of water from two different sources (the
ligands and the solution), it is easier to simplify this:

However, if you write it like this, remember that the hydrogen ion isn't just falling
off the complex ion. It is being pulled off by a water molecule in the solution.
Whenever you write "H+(aq)" what you really mean is a hydroxonium ion, H3O+.

Note: You will find the full reasons for the acidity of hexaaqua ions if
you follow this link. You only need to read the beginning of that page
which concentrates on explaining the acidity of the hexaaquairon(III)
ion. What is said applies equally to the chromium-containing ion.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/inorganic/transition/chromium.html (2 of 16)30/12/2004 11:35:13

chromium

Ligand exchange reactions involving chloride or sulphate ions


The hexaaquachromium(III) ion is a "difficult to describe" violet-blue-grey colour.
However, when it is produced during a reaction in a test tube, it is often green.
We nearly always describe the green ion as being Cr3+(aq) - implying the
hexaaquachromium(III) ion. That's actually an over-simplification.
What happens is that one or more of the ligand water molecules get replaced by a
negative ion in the solution - typically sulphate or chloride.
Replacement of the water by sulphate ions
You can do this simply by warming some chromium(III) sulphate solution.

One of the water molecules is replaced by a sulphate ion. Notice the change in
the charge on the ion. Two of the positive charges are cancelled by the presence
of the two negative charges on the sulphate ion.
Replacement of the water by chloride ions
In the presence of chloride ions (for example with chromium(III) chloride), the
most commonly observed colour is green. This happens when two of the water
molecules are replaced by chloride ions to give the tetraaquadichlorochromium(III)
ion - [Cr(H2O)4Cl2]+.
Once again, notice that replacing water molecules by chloride ions changes the
charge on the ion.

http://www.chemguide.co.uk/inorganic/transition/chromium.html (3 of 16)30/12/2004 11:35:13

chromium

Note: You will find an extensive discussion of ligand exchange


reactions if you follow this link.
Use the BACK button on your browser to return to this page.

Reactions of hexaaquachromium(III) ions with hydroxide ions


Hydroxide ions (from, say, sodium hydroxide solution) remove hydrogen ions from
the water ligands attached to the chromium ion.
Once a hydrogen ion has been removed from three of the water molecules, you
are left with a complex with no charge - a neutral complex. This is insoluble in
water and a precipitate is formed.

Note: The colour coding is to show that this isn't a ligand exchange
reaction. The oxygens which were originally attached to the chromium
are still attached in the neutral complex.

But the process doesn't stop there. More hydrogen ions are removed to give ions
like [Cr(H2O)2(OH)4]- and [Cr(OH)6]3-.
For example:

The precipitate redissolves because these ions are soluble in water.


In the test-tube, the colour changes are:

http://www.chemguide.co.uk/inorganic/transition/chromium.html (4 of 16)30/12/2004 11:35:13

chromium

Note: You will find the reactions between hexaaqua ions and
hydroxide ions discussed in detail if you follow this link.
Use the BACK button on your browser to return to this page.

Reactions of hexaaquachromium(III) ions with ammonia solution


The ammonia acts as both a base and a ligand. With a small amount of ammonia,
hydrogen ions are pulled off the hexaaqua ion exactly as in the hydroxide ion case
to give the same neutral complex.

That precipitate dissolves to some extent if you add an excess of ammonia


(especially if it is concentrated). The ammonia replaces water as a ligand to give
hexaamminechromium(III) ions.

http://www.chemguide.co.uk/inorganic/transition/chromium.html (5 of 16)30/12/2004 11:35:13

chromium

Note: You might wonder why this second equation is given starting
from the original hexaaqua ion rather than the neutral complex.
Explaining why the precipitate redissolves is quite complicated. You
will find the explanation in full (although by reference to the
corresponding copper case) on the page about the reactions between
hexaaqua ions and ammonia solution.
Use the BACK button on your browser to return to this page.

The colour changes are:

Reactions of hexaaquachromium(III) ions with carbonate ions


If you add sodium carbonate solution to a solution of hexaaquachromium(III) ions,
you get exactly the same precipitate as if you added sodium hydroxide solution or
ammonia solution.
This time, it is the carbonate ions which remove hydrogen ions from the hexaaqua
ion and produce the neutral complex.
Depending on the proportions of carbonate ions to hexaaqua ions, you will get
either hydrogencarbonate ions formed or carbon dioxide gas from the reaction
http://www.chemguide.co.uk/inorganic/transition/chromium.html (6 of 16)30/12/2004 11:35:13

chromium

between the hydrogen ions and carbonate ions. The more usually quoted
equation shows the formation of carbon dioxide.

Apart from the carbon dioxide, there is nothing new in this reaction:

Note: You will find the reactions between hexaaqua ions and
carbonate ions discussed in detail if you follow this link.
Use the BACK button on your browser to return to this page.

The oxidation of chromium(III) to chromium(VI)


An excess of sodium hydroxide solution is added to a solution of the
hexaaquachromium(III) ions to produce a solution of green hexahydroxochromate
(III) ions.

http://www.chemguide.co.uk/inorganic/transition/chromium.html (7 of 16)30/12/2004 11:35:13

chromium

This is then oxidised by warming it with hydrogen peroxide solution. You


eventually get a bright yellow solution containing chromate(VI) ions.

The equation for the oxidation stage is:

Note: Although it is still a complex ion, you don't write square brackets
around the chromate(VI) ion - any more than you would around a
sulphate or carbonate ion.
If you want to know how to work out this equation , follow this link.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/inorganic/transition/chromium.html (8 of 16)30/12/2004 11:35:13

chromium

Some chromium(VI) chemistry


The chromate(VI)-dichromate(VI) equilibrium
You are probably more familiar with the orange dichromate(VI) ion, Cr2O72-, than
the yellow chromate(VI) ion, CrO42-.
Changing between them is easy:
If you add dilute sulphuric acid to the yellow solution it turns orange. If you add
sodium hydroxide solution to the orange solution it turns yellow.

Note: If you had just produced the yellow chromate(VI) ions by


oxidising chromium(III) ions using hydrogen peroxide, you can't convert
them into dichromate(VI) ions without taking a precaution first.
In the presence of acid, dichromate(VI) ions react with any hydrogen
peroxide which is left in the solution from the original reaction. To
prevent this, you heat the solution for some time to decompose the
hydrogen peroxide into water and oxygen before adding the acid.

http://www.chemguide.co.uk/inorganic/transition/chromium.html (9 of 16)30/12/2004 11:35:13

chromium

The equilibrium reaction at the heart of the interconversion is:

If you add extra hydrogen ions to this, the equilibrium shifts to the right. This is
consistent with Le Chatelier's Principle.

Note: If you aren't familiar with Le Chatelier's Principle, you should


follow this link and read the first part of that page about the effect of
concentration on position of equilibrium.
Use the BACK button on your browser to return to this page.

If you add hydroxide ions, these react with the hydrogen ions. The equilibrium tips
to the left to replace them.

http://www.chemguide.co.uk/inorganic/transition/chromium.html (10 of 16)30/12/2004 11:35:13

chromium

The reduction of dichromate(VI) ions with zinc and an acid


Dichromate(VI) ions (for example, in potassium dichromate(VI) solution) can be
reduced to chromium(III) ions and then to chromium(II) ions using zinc and either
dilute sulphuric acid or hydrochloric acid.
Hydrogen is produced from a side reaction between the zinc and acid. This must
be allowed to escape, but you need to keep air out of the reaction. Oxygen in the
air rapidly re-oxidises chromium(II) to chromium(III).
An easy way of doing this is to put a bit of cotton wool in the top of the flask (or
test-tube) that you are using. This allows the hydrogen to escape, but stops most
of the air getting in against the flow of the hydrogen.

The reason for the inverted commas around the chromium(III) ion is that this is a
simplification. The exact nature of the complex ion will depend on which acid you
use in the reduction process. This has already been discussed towards the top of
the page.

Note: To re-read this use this link.

http://www.chemguide.co.uk/inorganic/transition/chromium.html (11 of 16)30/12/2004 11:35:13

chromium

The equations for the two stages of the reaction are:


For the reduction from +6 to +3:

For the reduction from +3 to +2:

Note: If you don't know how to work out equations like this, you can
find out how to do it on the page about writing ionic equations for redox
reactions.
Use the BACK button on your browser to return to this page.

Using potassium dichromate(VI) as an oxidising agent in organic chemistry


Potassium dichromate(VI) solution acidified with dilute sulphuric acid is commonly
used as an oxidising agent in organic chemistry. It is a reasonably strong oxidising
agent without being so powerful that it takes the whole of the organic molecule to
pieces! (Potassium manganate(VII) solution has some tendency to do that!)
It is used to:

oxidise secondary alcohols to ketones;

oxidise primary alcohols to aldehydes;

oxidise primary alcohols to carboxylic acids.

For example, with ethanol (a primary alcohol), you can get either ethanal (an
aldehyde) or ethanoic acid (a carboxylic acid) depending on the conditions.

If the alcohol is in excess, and you distil off the aldehyde as soon as it is
formed, you get ethanal as the main product.

http://www.chemguide.co.uk/inorganic/transition/chromium.html (12 of 16)30/12/2004 11:35:13

chromium

If the oxidising agent is in excess, and you don't allow the product to
escape - for example, by heating the mixture under reflux (heating the flask
with a condenser placed vertically in the neck) - you get ethanoic acid.

In organic chemistry, these equations are often simplified to concentrate on what


is happening to the organic molecules. For example, the last two could be written:

The oxygen written in square brackets just means "oxygen from an oxidising
agent".

Note: These are not a proper substitute for real equations. Only use
them if your examiners are happy with them. Check your syllabus and
look at past papers and mark schemes.
If you don't have these things, UK A' level students can find out how to
get hold of them by going to the syllabuses page.

Using potassium dichromate(VI) as an oxidising agent in titrations


Potassium dichromate(VI) is often used to estimate the concentration of iron(II)
ions in solution. It serves as an alternative to using potassium manganate(VII)
solution.

http://www.chemguide.co.uk/inorganic/transition/chromium.html (13 of 16)30/12/2004 11:35:13

chromium

Note: Potassium manganate(VII) titrations are described fully on the


page about manganese chemistry.

In practice
There are advantages and disadvantages in using potassium dichromate(VI).
Advantages

Potassium dichromate(VI) can be used as a primary standard. That means


that it can be made up to give a stable solution of accurately known
concentration. That isn't true of potassium manganate(VII).
Potassium dichromate(VI) can be used in the presence of chloride ions (as
long as the chloride ions aren't present in very high concentration).
Potassium manganate(VII) oxidises chloride ions to chlorine; potassium
dichromate(VI) isn't quite a strong enough oxidising agent to do this. That
means that you don't get unwanted side reactions with the potassium
dichromate(VI) soution.

Disadvantage

The main disadvantage lies in the colour change. Potassium manganate


(VII) titrations are self-indicating. As you run the potassium manganate(VII)
solution into the reaction, the solution becomes colourless. As soon as you
add as much as one drop too much, the solution becomes pink - and you
know you have reached the end point.
Unfortunately potassium dichromate(VI) solution turns green as you run it
into the reaction, and there is no way you could possibly detect the colour
change when you have one drop of excess orange solution in a strongly
coloured green solution.
With potassium dichromate(VI) solution you have to use a separate
indicator, known as a redox indicator. These change colour in the presence
of an oxidising agent.

http://www.chemguide.co.uk/inorganic/transition/chromium.html (14 of 16)30/12/2004 11:35:13

chromium

There are several such indicators - such as diphenylamine sulphonate. This


gives a violet-blue colour in the presence of excess potassium dichromate
(VI) solution. However, the colour is made difficult by the strong green also
present. The end point of a potassium dichromate(VI) titration isn't as easy
to see as the end point of a potassium manganate(VII) one.
The calculation
The half-equation for the dichromate(VI) ion is:

. . . and for the iron(II) ions is:

Combining these gives:

You can see that the reacting proportions are 1 mole of dichromate(VI) ions to 6
moles of iron(II) ions.
Once you have established that, the titration calculation is going to be just like any
other one.

Note: If you aren't very good at doing titration calculations, you might
be interested in my chemistry calculations book.

http://www.chemguide.co.uk/inorganic/transition/chromium.html (15 of 16)30/12/2004 11:35:13

chromium

Where would you like to go now?


To the transition metal menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/inorganic/transition/chromium.html (16 of 16)30/12/2004 11:35:13

manganese

MANGANESE

This page looks at some aspects of manganese chemistry required for


UK A' level exams. It includes: two simple reactions of manganese(II)
ions in solution (summarised from elsewhere on the site), and the use of
potassium manganate(VII) (potassium permanganate) as an oxidising
agent - including its use in titrations.

Reactions of manganese(II) ions in solution


The simplest ion that manganese forms in solution is the
hexaaquamanganese(II) ion - [Mn(H2O)6]2+.

Note: If you aren't happy about complex ions (including the


way they are bonded and named), it would pay you to follow
this link and explore the first couple of pages in the complex
ions menu before you go on.
Use the BACK button on your browser to return to this page.

The reaction of hexaaquamanganese(II) ions with hydroxide ions


Hydroxide ions (from, say, sodium hydroxide solution) remove hydrogen
ions from the water ligands attached to the manganese ion.
Once a hydrogen ion has been removed from two of the water
molecules, you are left with a complex with no charge - a neutral
complex. This is insoluble in water and a precipitate is formed.

http://www.chemguide.co.uk/inorganic/transition/manganese.html (1 of 8)30/12/2004 11:35:23

manganese

Note: The colour coding is to show that this isn't a ligand


exchange reaction. The oxygens which were originally
attached to the manganese are still attached in the neutral
complex.

In the test-tube, the colour changes are:

I have shown the original solution as very pale pink (the palest I can
produce!), but in fact it is virtually colourless. The pale brown precipitate
is oxidised to darker brown manganese(III) oxide in contact with oxygen
from the air.

Note: You will find the reactions between hexaaqua ions and
hydroxide ions discussed in detail if you follow this link.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/inorganic/transition/manganese.html (2 of 8)30/12/2004 11:35:23

manganese

The reaction of hexaaquamanganese(II) ions with ammonia solution


Ammonia can act as both a base and a ligand. In this case, at usual lab
concentrations, it simply acts as a base - removing hydrogen ions from
the aqua complex.

Again, I have shown the original solution as the palest pink I can
produce, but in fact it is virtually colourless. The pale brown precipitate is
oxidised to darker brown manganese(III) oxide in contact with oxygen
from the air.
There is no observable difference in appearance between this reaction
and the last one.

Note: You will find the reactions between hexaaqua ions and
ammonia solution discussed in detail if you follow this link.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/inorganic/transition/manganese.html (3 of 8)30/12/2004 11:35:23

manganese

Some potassium manganate(VII) chemistry


Potassium manganate(VII) (potassium permanganate) is a powerful
oxidising agent.
Using potassium manganate(VII) as an oxidising agent in organic
chemistry
Potassium manganate(VII) is usually used in neutral or alkaline solution
in organic chemistry. Acidified potassium manganate(VII) tends to be a
rather destructively strong oxidising agent, breaking carbon-carbon
bonds.
The potassium manganate(VII) solution is usually made mildly alkaline
with sodium carbonate solution, and the typical colour changes are:

In testing for a C=C double bond


Potassium manganate(VII) oxidises carbon-carbon double bonds, and so
goes through the colour changes above.
Ethene, for example, is oxidised to ethane-1,2-diol.

http://www.chemguide.co.uk/inorganic/transition/manganese.html (4 of 8)30/12/2004 11:35:23

manganese

The oxygen in square brackets is taken to mean "oxygen from an


oxidising agent". This abbreviated form of the equation is most commonly
used in organic chemistry. You are very unlikely to have to write the
complete ionic equation for this reaction at this level.
To be honest, this isn't a good test for a carbon-carbon double bond,
because anything which is even mildly reducing would have the same
effect on the potassium manganate(VII) solution.
You could, however, use this reaction simply as a means of making the
diol.
In the oxidation of aromatic side chains
Alkaline potassium manganate(VII) solution oxidises any hydrocarbon
side chain attached to a benzene ring back to a single -COOH group.
Prolonged heating is necessary.
For example:

In the case of the ethyl side chain, you will also get carbon dioxide. With
longer side chains, you will get all sorts of mixtures of other products but in each case, the main product will be benzoic acid.

Using potassium manganate(VII) as an oxidising agent in titrations

http://www.chemguide.co.uk/inorganic/transition/manganese.html (5 of 8)30/12/2004 11:35:23

manganese

Background
Potassium manganate(VII) solution is used to find the concentration of all
sorts of reducing agents. It is always used in acidic solution.
For example, it oxidises

iron(II) ions to iron(III) ions

hydrogen peroxide solution to oxygen

ethanedioic acid to carbon dioxide (This reaction has to be done


hot.)

sulphite ions (sulphate(IV) ions) to sulphate ions (sulphate(VI) ions)

In each case, the half-equation for the manganate(VII) ions in acidic


solution is:

These equations can be combined to give you an overall ionic equation


for each possible reaction. That, of course, also gives you the reacting
proportions.
For example, when the equations are combined, you find that 1 mole of
MnO4- ions react with 5 moles of Fe2+ ions. Having got that information,
the titration calculations are just like any other ones.

http://www.chemguide.co.uk/inorganic/transition/manganese.html (6 of 8)30/12/2004 11:35:23

manganese

Note: If you don't know how to work out half-equations and


combine them into equations like this, you can find out how to
do it on the page about writing ionic equations for redox
reactions.
If you aren't very good at titration calculations, you might be
interested in my chemistry calculations book.

Doing the titration


The potassium manganate(VII) solution always goes into the burette, and
the other solution in the flask is acidified with dilute sulphuric acid.
As the potassium manganate(VII) solution is run into the flask it becomes
colourless. The end point is the first faint trace of permanent pink in the
solution showing that there is a tiny excess of manganate(VII) ions
present.
Problems with the use of potassium manganate(VII) solution
There are two things you need to be aware of:

Potassium manganate(VII) can't be used in titrations in the


presence of ions like chloride or bromide which it oxidises. An
unknown amount of the potassium manganate(VII) would be used
in side reactions, and so the titration result would be inaccurate.
This is why you don't acidify the solution with hydrochloric acid.

Potassium manganate(VII) isn't a primary standard. That means


that it can't be made up to give a stable solution of accurately
known concentration.
It is so strongly coloured that it is impossible to see when all the
crystals you have used have dissolved, and over a period of time it
oxidises the water it is dissolved in to oxygen.

http://www.chemguide.co.uk/inorganic/transition/manganese.html (7 of 8)30/12/2004 11:35:23

manganese

Bottles of potassium manganate(VII) solution usually have a brown


precipitate around the top. This is manganese(IV) oxide - and is
produced when the manganate(VII) ions react with the water.
You have to make up a solution which is approximately what you
want, and then standardise it by doing a titration. This is often
done with ethanedioic acid solution, because this is a primary
standard.

Where would you like to go now?


To the transition metal menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/inorganic/transition/manganese.html (8 of 8)30/12/2004 11:35:23

iron

IRON

This page looks at some aspects of iron chemistry required for UK A'
level exams (summarised from elsewhere on the site). You will find links
to pages where full details and explanations are given.

Iron and its ions as catalysts


Iron as catalyst in the Haber Process
The Haber Process combines nitrogen and hydrogen into ammonia. The
nitrogen comes from the air and the hydrogen is obtained mainly from
natural gas (methane). Iron is used as a catalyst.

Note: You can find a full discussion about the Haber


Process by following this link.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/inorganic/transition/iron.html (1 of 12)30/12/2004 11:35:35

iron

Iron ions as a catalyst in the reaction between persulphate ions and


iodide ions
The reaction between persulphate ions (peroxodisulphate ions), S2O82-,
and iodide ions in solution can be catalysed using either iron(II) or iron
(III) ions.
The overall equation for the reaction is:

For the sake of argument, we'll take the catalyst to be iron(II) ions. The
reaction happens in two stages.

If you use iron(III) ions, the second of these reactions happens first.
This is a good example of the use of transition metal compounds as
catalysts because of their ability to change oxidation state.

Note: This reaction is explained in more detail on a page


about types of catalysis.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/inorganic/transition/iron.html (2 of 12)30/12/2004 11:35:35

iron

Reactions of iron ions in solution


The simplest ions in solution are:

the hexaaquairon(II) ion - [Fe(H2O)6]2+.

the hexaaquairon(III) ion - [Fe(H2O)6]3+.

They are both acidic, but the iron(III) ion is more strongly acidic.

Note: If you aren't happy about complex ions (including the


way they are bonded and named), it would pay you to follow
this link and explore the first couple of pages in the complex
ions menu before you go on.
The acidity of the complex ions is discussed in detail on
another page.
Use the BACK button on your browser to return to this page.

Reactions of the iron ions with hydroxide ions


Hydroxide ions (from, say, sodium hydroxide solution) remove hydrogen
ions from the water ligands attached to the iron ions.
When enough hydrogen ions have been removed, you are left with a
complex with no charge - a neutral complex. This is insoluble in water
and a precipitate is formed.
In the iron(II) case:

In the iron(III) case:


http://www.chemguide.co.uk/inorganic/transition/iron.html (3 of 12)30/12/2004 11:35:35

iron

Note: The colour coding is to show that these aren't ligand


exchange reactions. The oxygens which were originally
attached to the iron are still attached in the neutral complexes.

In the test-tube, the colour changes are:


In the iron(II) case:

Iron is very easily oxidised under alkaline conditions. Oxygen in the air
oxidises the iron(II) hydroxide precipitate to iron(III) hydroxide especially
around the top of the tube. The darkening of the precipitate comes from
the same effect.
In the iron(III) case:

http://www.chemguide.co.uk/inorganic/transition/iron.html (4 of 12)30/12/2004 11:35:35

iron

Note: You will find the reactions between hexaaqua ions and
hydroxide ions discussed in detail if you follow this link.
Use the BACK button on your browser to return to this page.

Reactions of the iron ions with ammonia solution


Ammonia can act as both a base and a ligand. In these cases, it simply
acts as a base - removing hydrogen ions from the aqua complex.
In the iron(II) case:

http://www.chemguide.co.uk/inorganic/transition/iron.html (5 of 12)30/12/2004 11:35:35

iron

The appearance is just the same as in when you add sodium hydroxide
solution. The precipitate again changes colour as the iron(II) hydroxide
complex is oxidised by the air to iron(III) hydroxide.
In the iron(III) case:

The reaction looks just the same as when you add sodium hydroxide
solution.

Note: You will find the reactions between hexaaqua ions and
ammonia solution discussed in detail if you follow this link.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/inorganic/transition/iron.html (6 of 12)30/12/2004 11:35:35

iron

Reactions of the iron ions with carbonate ions


There is an important difference here between the behaviour of iron(II)
and iron(III) ions.
Iron(II) ions and carbonate ions
You simply get a precipitate of what you can think of as iron(II) carbonate.

Note: Only one of the UK A' level Exam Boards wants this,
and this is the simplification that they make. In fact, the
precipitate is better described as a basic carbonate with a
formula of the type xFeCO3,yFe(OH)2,zH2O.

Iron(III) ions and carbonate ions


The hexaaquairon(III) ion is sufficiently acidic to react with the weakly
basic carbonate ion.
If you add sodium carbonate solution to a solution of hexaaquairon(III)
ions, you get exactly the same precipitate as if you added sodium
hydroxide solution or ammonia solution.
This time, it is the carbonate ions which remove hydrogen ions from the
hexaaqua ion and produce the neutral complex.

http://www.chemguide.co.uk/inorganic/transition/iron.html (7 of 12)30/12/2004 11:35:35

iron

Depending on the proportions of carbonate ions to hexaaqua ions, you


will get either hydrogencarbonate ions formed or carbon dioxide gas from
the reaction between the hydrogen ions and carbonate ions. The more
usually quoted equation shows the formation of carbon dioxide.

Apart from the carbon dioxide, there is nothing new in this reaction:

Note: You will find the reactions between hexaaqua ions and
carbonate ions discussed in detail if you follow this link.
Use the BACK button on your browser to return to this page.

Testing for iron(III) ions with thiocyanate ions


This provides an extremely sensitive test for iron(III) ions in solution.
If you add thiocyanate ions, SCN-, (from, say, sodium or potassium or
ammonium thiocyanate solution) to a solution containing iron(III) ions,
you get an intense blood red solution containing the ion [Fe(SCN)(H2O)5]
2+.

http://www.chemguide.co.uk/inorganic/transition/iron.html (8 of 12)30/12/2004 11:35:35

iron

Note: I haven't been able to suggest properly the intense


rich red of the thiocyano complex. Imagine blood!
This is a ligand exchange reaction. You can find out more
about these by following this link.
Use the BACK button on your browser to return quickly to this
page.

Finding the concentration of iron(II) ions in solution by titration


You can find the concentration of iron(II) ions in solution by titrating with
either potassium manganate(VII) solution or potassium dichromate(VI)
solution. The reactions are done in the presence of dilute sulphuric acid.
In either case, you would pipette a known volume of solution containing
the iron(II) ions into a flask, and add a roughly equal volume of dilute
sulphuric acid. What happens next depends on whether you are using
potassium manganate(VII) solution or potassium dichromate(VI) solution.
Using potassium manganate(VII) solution
The potassium manganate(VII) solution is run in from a burette. At first, it
turns colourless as it reacts. The end point is the first trace of permanent
pink in the solution showing a tiny excess of manganate(VII) ions.
The manganate(VII) ions oxidise iron(II) to iron(III) ions. The two halfhttp://www.chemguide.co.uk/inorganic/transition/iron.html (9 of 12)30/12/2004 11:35:35

iron

equations for the reaction are:

These combine to give the ionic equation for the reaction:

Note: If you don't know how to work out half-equations and


combine them into equations like this, you can find out how to
do it on the page about writing ionic equations for redox
reactions.
Use the BACK button on your browser to return to this page
later.

The complete equation shows that 1 mole of manganate(VII) ions react


with 5 moles of iron(II) ions. Having got that information, the titration
calculations are just like any other ones.

Note: If you aren't very good at titration calculations, you


might be interested in my chemistry calculations book.
There is more about potassium manganate(VII) titrations on
the page about manganese chemistry.

http://www.chemguide.co.uk/inorganic/transition/iron.html (10 of 12)30/12/2004 11:35:35

iron

Using potassium dichromate(VI) solution


Potassium dichromate(VI) solution turns green as it reacts with the iron
(II) ions, and there is no way you could possibly detect the colour change
when you have one drop of excess orange solution in a strongly coloured
green solution.
With potassium dichromate(VI) solution you have to use a separate
indicator, known as a redox indicator. These change colour in the
presence of an oxidising agent.
There are several such indicators - such as diphenylamine sulphonate.
This gives a violet-blue colour in the presence of excess potassium
dichromate(VI) solution.
The two half-equations are:

Combining these gives:

You can see that the reacting proportions are 1 mole of dichromate(VI)
ions to 6 moles of iron(II) ions.
Once you have established that, the titration calculation is again going to
be just like any other one.

Note: There is more about potassium dichromate(VI)


titrations on the page about chromium chemistry.

http://www.chemguide.co.uk/inorganic/transition/iron.html (11 of 12)30/12/2004 11:35:35

iron

Where would you like to go now?


To the transition metal menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/inorganic/transition/iron.html (12 of 12)30/12/2004 11:35:35

copper

COPPER

This page looks at some aspects of copper chemistry required for UK A'
level exams. The chemistry of copper(II) is mainly summarised from
elsewhere on the site, with links available to more detailed explanations.
The page also covers some simple copper(I) chemistry.

Reactions of copper(II) ions in solution


The simplest ion that copper forms in solution is the typical blue
hexaaquacopper(II) ion - [Cu(H2O)6]2+.

Note: If you aren't happy about complex ions (including the


way they are bonded and named), it would pay you to follow
this link and explore the first couple of pages in the complex
ions menu before you go on.
Use the BACK button on your browser to return to this page.

The reaction of hexaaquacopper(II) ions with hydroxide ions


Hydroxide ions (from, say, sodium hydroxide solution) remove hydrogen
ions from the water ligands attached to the copper ion.
Once a hydrogen ion has been removed from two of the water
molecules, you are left with a complex with no charge - a neutral
complex. This is insoluble in water and a precipitate is formed.

http://www.chemguide.co.uk/inorganic/transition/copper.html (1 of 10)30/12/2004 11:35:49

copper

Note: The colour coding is to show that this isn't a ligand


exchange reaction. The oxygens which were originally
attached to the copper are still attached in the neutral
complex.

In the test-tube, the colour change is:

Note: You will find the reactions between hexaaqua ions and
hydroxide ions discussed in detail if you follow this link.
Use the BACK button on your browser to return to this page.

Reactions of hexaaquacopper(II) ions with ammonia solution


The ammonia acts as both a base and a ligand. With a small amount of
ammonia, hydrogen ions are pulled off the hexaaqua ion exactly as in the
hydroxide ion case to give the same neutral complex.

That precipitate dissolves if you add an excess of ammonia.


The ammonia replaces water as a ligand to give
tetraamminediaquacopper(II) ions. Notice that only 4 of the 6 water
http://www.chemguide.co.uk/inorganic/transition/copper.html (2 of 10)30/12/2004 11:35:49

copper

molecules are replaced.

Note: You might wonder why this second equation is given


starting from the original hexaaqua ion rather than the neutral
complex. Explaining why the precipitate redissolves is quite
complicated. You will find the explanation in full on the page
about the reactions between hexaaqua ions and ammonia
solution.
Use the BACK button on your browser to return to this page.

The colour changes are:

The reaction of hexaaquacopper(II) ions with carbonate ions


You simply get a precipitate of what you can think of as copper(II)
carbonate.

http://www.chemguide.co.uk/inorganic/transition/copper.html (3 of 10)30/12/2004 11:35:49

copper

Note: Only one of the UK A' level Exam Boards wants this,
and this is the simplification that they make. In fact, the
precipitate is better described as a basic carbonate with a
formula of the type xCuCO3,yCu(OH)2,zH2O.
You will find the reactions between hexaaqua ions and
carbonate ions discussed in detail if you follow this link.
Use the BACK button on your browser to return to this page.

A ligand exchange reaction involving chloride ions


If you add concentrated hydrochloric acid to a solution containing
hexaaquacopper(II) ions, the six water molecules are replaced by four
chloride ions.

http://www.chemguide.co.uk/inorganic/transition/copper.html (4 of 10)30/12/2004 11:35:49

copper

The reaction taking place is reversible.

Because the reaction is reversible, you get a mixture of colours due to


both of the complex ions.
You may find the colour of the tetrachlorocuprate(II) ion variously
described as olive-green or yellow. If you add water to the green solution,
it returns to the blue colour.

Note: You will find this reaction explored in more detail on a


page about ligand exchange reactions.
Use the BACK button on your browser to return to this page.

The reaction of hexaaquacopper(II) ions with iodide ions


The simple reaction
Copper(II) ions oxidise iodide ions to iodine, and in the process are
themselves reduced to copper(I) iodide.

http://www.chemguide.co.uk/inorganic/transition/copper.html (5 of 10)30/12/2004 11:35:49

copper

Note: If you need to, you can find out more about oxidation
states by following this link.
Use the BACK button on your browser to return to this page.

The initial mucky brown mixture separates into an off-white precipitate of


copper(I) iodide under an iodine solution.

Note: I've had several attempts at getting this colour right,


but my software keeps changing it slightly as I adapt the
diagram for Web use. The colour of the iodine solution
actually varies quite a lot depending on its concentration, so
I'm not too bothered about this!

http://www.chemguide.co.uk/inorganic/transition/copper.html (6 of 10)30/12/2004 11:35:49

copper

Using this reaction to find the concentration of copper(II) ions in


solution
If you pipette a known volume of a solution containing copper(II) ions into
a flask, and then add an excess of potassium iodide solution, you get the
reaction we have just described.

You can find the amount of iodine liberated by titration with sodium
thiosulphate solution.

As the sodium thiosulphate solution is run in from a burette, the colour of


the iodine fades. When it is almost all gone, you add some starch
solution. This reacts reversibly with iodine to give a deep blue starchiodine complex which is much easier to see.
You add the last few drops of the sodium thiosulphate solution slowly
until the blue colour disappears.
If you trace the reacting proportions through the two equations, you will
find that for every 2 moles of copper(II) ions you had to start with, you
need 2 moles of sodium thiosulphate solution.
If you know the concentration of the sodium thiosulphate solution, it is
easy to calculate the concentration of the copper(II) ions.

Note: If you aren't very good at titration calculations, you


might be interested in my chemistry calculations book. This
particular example is worked through in detail on pages 6667.

http://www.chemguide.co.uk/inorganic/transition/copper.html (7 of 10)30/12/2004 11:35:49

copper

Some essential copper(I) chemistry


The disproportionation of copper(I) ions in solution
Copper(I) chemistry is limited by a reaction which occurs involving simple
copper(I) ions in solution. This is a good example of disproportionation
- a reaction in which something oxidises and reduces itself.
Copper(I) ions in solution disproportionate to give copper(II) ions and a
precipitate of copper.
The reaction is:

Any attempt to produce a simple copper(I) compound in solution results


in this happening.
For example, if you react copper(I) oxide with hot dilute sulphuric acid,
you might expect to get a solution of copper(I) sulphate and water
produced. In fact you get a brown precipitate of copper and a blue
solution of copper(II) sulphate because of the disproportionation reaction.

Stabalising the copper(I) oxidation state


Insoluble copper(I) compounds
We've already seen that copper(I) iodide is produced as an off-white
precipitate if you add potassium iodide solution to a solution containing
copper(II) ions. The copper(I) iodide is virtually insoluble in water, and so

http://www.chemguide.co.uk/inorganic/transition/copper.html (8 of 10)30/12/2004 11:35:49

copper

the disproportionation reaction doesn't happen.


Similarly copper(I) chloride can be produced as a white precipitate
(reaction described below). Provided this is separated from the solution
and dried as quickly as possible, it remains white. In contact with water,
though, it slowly turns blue as copper(II) ions are formed.
The disproportionation reaction only occurs with simple copper(I) ions in
solution.
Copper(I) complexes
Forming copper(I) complexes (other than the one with water as a ligand)
also stabalises the copper(I) oxidation state.
For example, both [Cu(NH3)2]+ and [CuCl2]- are copper(I) complexes
which don't disproportionate.
The chlorine-containing complex is formed if copper(I) oxide is dissolved
in concentrated hydrochloric acid.
You can think of this happening in two stages. First, you get copper(I)
chloride formed:

But in the presence of excess chloride ions from the HCl, this reacts to
give a stable, soluble copper(I) complex.

You can get the white precipitate of copper(I) chloride (mentioned above)
by adding water to this solution. This reverses the last reaction by
stripping off the extra chloride ion.

Where would you like to go now?


To the transition metal menu . . .

http://www.chemguide.co.uk/inorganic/transition/copper.html (9 of 10)30/12/2004 11:35:49

copper

To the Inorganic Chemistry menu . . .


To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/inorganic/transition/copper.html (10 of 10)30/12/2004 11:35:49

substitution in complex ions - stability constants

COMPLEX METAL IONS - STABILITY CONSTANTS

This page explains what is meant by a stability constant for a complex ion,
and goes on to look at how its size is governed in part by the entropy
change during a ligand exchange reaction.

Note: If you aren't happy about what a ligand exchange


reaction is, it would be useful to have a quick look at this link
before you go on. There is no need to worry about the detail,
but you should read enough to get a feel for what we are
talking about.

What is a stability constant?


Replacing water by ammonia around copper(II) ions
If you add ammonia solution to a solution containing hexaaquacopper(II)
ions, [Cu(H2O)6]2+, four of the water molecules are eventually replaced by
ammonia molecules to give [Cu(NH3)4(H2O)2]2+.
This can be written as an equilibrium reaction to show the overall effect:

In fact, the water molecules get replaced one at a time, and so this is
made up of a series of part-reactions:

http://www.chemguide.co.uk/inorganic/complexions/stabconst.html (1 of 11)30/12/2004 11:36:05

substitution in complex ions - stability constants

Although this can look a bit daunting at first sight, all that is happening is
that first you have one, then two, then three, then four water molecules in
total replaced by ammonias.
Individual stability constants
Let's take a closer look at the first of these equilibria:

Like any other equilibrium, this one has an equilibrium constant, Kc except that in this case, we call it a stability constant. Because this is the
first water molecule to be replaced, we call it K1.

Warning! If you aren't happy about how to write expressions


for equilibrium constants, it is essential to follow this link
before you go any further.
Use the BACK button on your browser to return quickly to this
page.

http://www.chemguide.co.uk/inorganic/complexions/stabconst.html (2 of 11)30/12/2004 11:36:05

substitution in complex ions - stability constants

Here is the equation again:

K1 is given by this expression:

There are two points of possible confusion here - one minor, one more
important!

First, the square brackets have changed their meaning! Square


brackets are often used to keep everything in a complex ion
together and tidy. Here, they have reverted to their other use, which
implies concentrations in mol dm-3.
In order to avoid complete confusion, the square brackets keeping
the complexes together have been removed entirely.

More importantly, if you compare the equilibrium constant


expression with the equation, you will see that the water on the righthand side hasn't been included. That is normal practice with these
expressions.
Because everything is dissolved in water, the water is present as a
huge excess. Generating a little bit more during the reaction is
going to make no effective difference to the total concentration of
the water in terms of moles of water per dm3.
The concentration of the water is approximately constant. The
equilibrium constant is defined so that you avoid having an extra
unnecessary constant in the expression.

With that out of the way, let's go back to where we were - but introduce a
value for K1:

http://www.chemguide.co.uk/inorganic/complexions/stabconst.html (3 of 11)30/12/2004 11:36:05

substitution in complex ions - stability constants

The value of the equilibrium constant is fairly large, suggesting that there
is a strong tendency to form the ion containing an ammonia molecule.
A high value of a stability constant shows that the ion is easily formed.
Each of the other equilibria above also has its own stability constant, K2,
K3 and K4. For example, K2 is given by:

The ion with two ammonias is even more stable than the ion with one
ammonia.
You could keep plugging away at this and come up with the following table
of stability constants:

ion

Kn

value (mol-1 dm3)

log Kn

[Cu(NH3)(H2O)5]2+

K1

1.78 x 104

4.25

[Cu(NH3)2(H2O)4]2+

K2

4.07 x 103

3.61

[Cu(NH3)3(H2O)3]2+

K3

9.55 x 102

2.98

http://www.chemguide.co.uk/inorganic/complexions/stabconst.html (4 of 11)30/12/2004 11:36:05

substitution in complex ions - stability constants

[Cu(NH3)4(H2O)2]2+

K4

1.74 x 102

2.24

You will often find these values quoted as log K1 or whatever. All this does
is tidy the numbers up so that you can see the patterns more easily.
The ions keep on getting more stable as you replace up to 4 water
molecules, but notice that the equilibrium constants are gradually getting
less big as you replace more and more waters. This is common with
individual stability constants.
Overall stability constants
The overall stability constant is simply the equilibrium constant for the total
reaction:

It is given by this expression:

You can see that overall this is a very large equilibrium constant, implying
a high tendency for the ammonias to replace the waters. The "log" value is
13.1.
This overall value is found by multiplying together all the individual values
of K1, K2 and so on. To find out why that works, you will need a big bit of
paper and some patience!
Write down expressions for all the individual values (the first two are done
for you above), and then multiply those expressions together. You will find
that all the terms for the intermediate ions cancel out to leave you with the
expression for the overall stability constant.

http://www.chemguide.co.uk/inorganic/complexions/stabconst.html (5 of 11)30/12/2004 11:36:05

substitution in complex ions - stability constants

Summary
Whether you are looking at the replacement of individual water molecules
or an overall reaction producing the final complex ion, a stability constant
is simply the equilibrium constant for the reaction you are looking at.
The larger the value of the stability constant, the further the reaction lies to
the right. That implies that complex ions with large stability constants are
more stable than ones with smaller ones.
Stability constants tend to be very large numbers. In order to simplify the
numbers a "log" scale is often used. Because of the way this works, a
difference of 1 in the log value comes from a 10 times difference in the
stability constant. A difference of 2 comes from a 100 (in other words, 102)
times difference in stability constant - and so on.

Stability constants and entropy - the chelate effect


What is the chelate effect?
This is an effect which happens when you replace water (or other simple
ligands) around the central metal ion by multidentate ligands like 1,2diaminoethane (often abbreviated to "en") or EDTA.

Important! If words like multidentate ligand or EDTA don't


mean anything to you, then it is essential that you read the
introductory page on complex ions. The second half of that
page deals with multidentate ligands.
Use the BACK button on your browser to return quickly to this
page.

http://www.chemguide.co.uk/inorganic/complexions/stabconst.html (6 of 11)30/12/2004 11:36:05

substitution in complex ions - stability constants

Compare what happens if you replace two water molecules around a [Cu
(H2O)6]2+ ion with either 2 ammonia molecules or one molecule of 1,2diaminoethane.

This second structure is known as a chelate from a Greek word meaning


a crab's claw. You can picture the copper ion as being nipped by the claw
of the 1,2-diaminoethane molecule.
Chelates are much more stable than complex ions formed from simple
monodentate ligands. The overall stability constants for the two ions are:

ion

log K

[Cu(NH3)2(H2O)4]2+

7.86

[Cu(H2O)4(en)]2+

10.6

The reaction with the 1,2-diaminoethane could eventually go on to


produce a complex ion [Cu(en)3]2+. Simplifying the structure of this:

http://www.chemguide.co.uk/inorganic/complexions/stabconst.html (7 of 11)30/12/2004 11:36:05

substitution in complex ions - stability constants

The overall stability constant for this (as log K) is 18.7.


Another copper-based chelate comes from the reaction with EDTA.

This also has a high stability constant - log K is 18.8.


However many examples you take, you always find that a chelate (a
complex ion involving multidentate ligands) is more stable than ions with
only unidentate ligands. This is known as the chelate effect.

http://www.chemguide.co.uk/inorganic/complexions/stabconst.html (8 of 11)30/12/2004 11:36:05

substitution in complex ions - stability constants

The reason for the chelate effect


If you compare the two equilibria below, the one with the 1,2diaminoethane ("en") has the higher equilibrium (stability) constant (for
values, see above).

The enthalpy changes of the two reactions are fairly similar. You might
expect this because in each case you are breaking two bonds between
copper and oxygen atoms and replacing them by two bonds between
copper and nitrogen atoms.

Note: This is an over-simplification. You have to consider


interactions between the various species and water molecules
(from the solvent) as well.

If the enthalpy changes are similar, what causes the difference in the
extent to which the two reactions happen?
You need to think about the entropy change during each reaction.
Entropy is most easily thought of as a measure of disorder. Any change
which increases the amount of disorder increases the tendency of a
reaction to happen.

Note: If you want a more formal approach to entropy, I'm


afraid you will have to look elsewhere. You will find a gentle
introduction to the mathematical side of entropy in my
chemistry calculations book.

http://www.chemguide.co.uk/inorganic/complexions/stabconst.html (9 of 11)30/12/2004 11:36:05

substitution in complex ions - stability constants

If you look again at the two equiilbria, you might notice that the 1,2diaminoethane equilibrium does lead to an increase in the disorder of the
system (an increase in its entropy). There are only two species on the lefthand side of the equation, but three on the right.

You can obviously get more disorder out of three species than out of only
two.
Compare that with the other equilibrium. In this case, there is no change in
the total number of species before and after reaction, and so no useful
contribution to an increase in entropy.

In the case of the complex with EDTA, the increase in entropy is very
pronounced.

Here, we are increasing the number of species present from two on the
left-hand side to seven on the right. You can get a major amount of
increase in disorder by making this change.
Reversing this last change is going to be far more difficult in entropy
terms. You would have to move from a highly disordered state to a much
more ordered one. That isn't so likely to happen, and so the copper-EDTA
complex is very stable.

Summary
Complexes involving multidentate ligands are more stable than those with
only unidentate ligands in them.
The underlying reason for this is that each multidentate ligand displaces
more than one water molecule. This leads to an increase in the number of
species present in the system, and therefore an increase in entropy.

http://www.chemguide.co.uk/inorganic/complexions/stabconst.html (10 of 11)30/12/2004 11:36:05

substitution in complex ions - stability constants

An increase in entropy makes the formation of the chelated complex more


favourable.

Where would you like to go now?


To the complex ion menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/inorganic/complexions/stabconst.html (11 of 11)30/12/2004 11:36:05

Reactions of the Group 2 elements with water

REACTIONS OF THE GROUP 2 ELEMENTS WITH


WATER

This page looks at the reactions of the Group 2 elements - beryllium,


magnesium, calcium, strontium and barium - with water (or steam). It
uses these reactions to explore the trend in reactivity in Group 2.

The Facts
Beryllium
Beryllium has no reaction with water or steam even at red heat.
Magnesium
Magnesium burns in steam to produce magnesium oxide and hydrogen.

Very clean magnesium has a very slight reaction with cold water. The
reaction soon stops because the magnesium hydroxide formed is almost
insoluble in water and forms a barrier on the magnesium preventing
further reaction.

Note: As a general rule, if a metal reacts with steam, the


metal oxide is formed. If it reacts with cold water, you get the
metal hydroxide.

http://www.chemguide.co.uk/inorganic/group2/reacth2o.html (1 of 6)30/12/2004 11:36:12

Reactions of the Group 2 elements with water

Calcium, strontium and barium


These all react with cold water with increasing vigour to give the metal
hydroxide and hydrogen. Strontium and barium have reactivities similar
to lithium in Group 1 of the Periodic Table.
The equation for the reactions of any of these metals would be:

The hydroxides aren't very soluble, but they get more soluble as you go
down the Group. The calcium hydroxide formed shows up mainly as a
white precipitate (although some does dissolve). You get less precipitate
as you go down the Group because more of the hydroxide dissolves in
the water.
Summary of the trend in reactivity
The Group 2 metals become more reactive towards water as you go
down the Group.

Explaining the trend in reactivity


Looking at the enthalpy changes for the reactions
The enthalpy change of a reaction is a measure of the amount of heat
absorbed or evolved when the reaction takes place. An enthalpy change
is negative if heat is evolved, and positive if it is absorbed. That's really
all you need to know for this section!

Note: If you aren't happy about calculating enthalpy


changes, you might be interested in my book on A'level
calculations.

http://www.chemguide.co.uk/inorganic/group2/reacth2o.html (2 of 6)30/12/2004 11:36:12

Reactions of the Group 2 elements with water

If you calculate the enthalpy change for the possible reactions between
beryllium or magnesium and steam, you come up with these answers:

Notice that both possible reactions are strongly exothermic, giving out
almost identical amounts of heat. However, only the magnesium reaction
actually happens. The explanation for the different reactivities must lie
somewhere else.
Similarly, if you calculate the enthalpy changes for the reactions between
calcium, strontium or barium and cold water, you again find that the
amount of heat evolved in each case is almost exactly the same - in this
case, about -430 kJ mol-1.
The reason for the increase in reactivity must again lie elsewhere.
Looking at the activation energies for the reactions
The activation energy for a reaction is the minimum amount of energy
which is needed in order for the reaction to take place. It doesn't matter
how exothermic the reaction would be once it got started - if there is a
high activation energy barrier, the reaction will take place very slowly, if
at all.
When Group 2 metals react to form oxides or hydroxides, metal ions are
formed.

http://www.chemguide.co.uk/inorganic/group2/reacth2o.html (3 of 6)30/12/2004 11:36:12

Reactions of the Group 2 elements with water

Note: This is a simplification in the case of beryllium.


Beryllium oxide isn't fully ionic. There isn't enough
electronegativity difference between the beryllium and
oxygen for the beryllium to lose control of the bonding pair of
electrons and form ions. The approach we are taking here is
in line with the sort of answer that you would be expected to
give at A'level. Thinking about beryllium as an entirely
different case would make this argument unnecessarily
complicated.

The formation of the ions from the original metal involves various stages
all of which require the input of energy - contributing to the activation
energy of the reaction. These stages involve the input of:

the atomisation energy of the metal. This is the energy needed to


break the bonds holding the atoms together in the metallic lattice.
the first + second ionisation energies. These are necessary to
convert the metal atoms into ions with a 2+ charge.

After this, there will be a number of steps which give out heat again leading to the formation of the products, and overall exothermic reactions.
The graph shows the effect of these important energy-absorbing stages
as you go down Group 2.

http://www.chemguide.co.uk/inorganic/group2/reacth2o.html (4 of 6)30/12/2004 11:36:12

Reactions of the Group 2 elements with water

Notice that the ionisation energies dominate this - particularly the second
ionisation energies. Ionisation energies fall as you go down the Group.
Because it gets easier to form the ions, the reactions will happen more
quickly.

Note: If you are unhappy about the changes in ionisation


energy as you go down Group 2 you should follow this link.
You will find a further link to a wider discussion of ionisation
energy if you need it.

Summarising the reason for the increase in reactivity as you go


down the Group
The reactions become easier as the energy needed to form positive ions
falls. This is mainly due to a decrease in ionisation energy as you go
down the Group. This leads to lower activation energies, and therefore
faster reactions.

Where would you like to go now?


To the Group 2 menu . . .
http://www.chemguide.co.uk/inorganic/group2/reacth2o.html (5 of 6)30/12/2004 11:36:12

Reactions of the Group 2 elements with water

To the Inorganic Chemistry menu . . .


To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/inorganic/group2/reacth2o.html (6 of 6)30/12/2004 11:36:12

Reactions of the Group 2 elements with air or oxygen

REACTIONS OF THE GROUP 2 ELEMENTS WITH


AIR OR OXYGEN

This page looks at the reactions of the Group 2 elements - beryllium,


magnesium, calcium, strontium and barium - with air or oxygen. It
explains why it is difficult to observe many tidy patterns.

The Facts
The reactions with oxygen
Formation of simple oxides
On the whole, the metals burn in oxygen to form a simple metal oxide.
Beryllium is reluctant to burn unless it is in the form of dust or powder.
Beryllium has a very strong (but very thin) layer of beryllium oxide on its
surface, and this prevents any new oxygen getting at the underlying
beryllium to react with it.

"X" in the equation can represent any of the metals in the Group.
It is almost impossible to find any trend in the way the metals react with
oxygen. It would be quite untrue to say that they burn more vigorously as
you go down the Group.
To be able to make any sensible comparison, you would have to have
pieces of metal which were all equally free of oxide coating, with exactly
the same surface area and shape, exactly the same flow of oxygen
around them, and heated to exactly the same extent to get them started.
It can't be done!

http://www.chemguide.co.uk/inorganic/group2/reacto2.html (1 of 9)30/12/2004 11:36:21

Reactions of the Group 2 elements with air or oxygen

Note: One of the UK Exam Boards (OCR) implies in their


syllabus that you should be able to state a trend and then
explain it in terms of ionisation energy differences. You can
do this with the reactions with water (or steam), and you
might like to follow this link if you haven't already been there.
Trying to account for a non-existent trend in the reactions
with oxygen is just silly!

What they metals look like when they burn is a bit problematical!

Beryllium: I can't find a reference anywhere (text books or internet)


to the colour of the flame that beryllium burns with. My best guess
would be the same sort of silvery sparkles that magnesium or
aluminium powder burn with if they are scattered into a flame - but
I don't know that for sure.
Magnesium, of course, burns with a typical intense white flame.
Calcium is quite reluctant to start burning, but then bursts
dramatically into flame, burning with an intense white flame with a
tinge of red at the end.
Strontium: I have never seen strontium burn. I suspect it might be
rather like calcium, but with a much richer red tinge to the flame.
Barium: I have only seen this burn on video, and although the
accompanying description talked about a pale green flame, the
flame appeared to be white with some pale green tinges. It wasn't
noticeably any more dramatic than the familiar magnesium flame.

http://www.chemguide.co.uk/inorganic/group2/reacto2.html (2 of 9)30/12/2004 11:36:21

Reactions of the Group 2 elements with air or oxygen

Note: If anyone reading this has actually seen beryllium,


strontium or barium burning in oxygen in the lab (not on
video), would they please contact me via the link on the
about this site page.
I know what the colours are when you do a flame test with
compounds of strontium and barium. What I need to know is
what colours you get if you burn a piece of the metal - that's
not necessarily the same thing.

Formation of peroxides
Strontium and barium will also react with oxygen to form strontium or
barium peroxide.
Strontium forms this if it is heated in oxygen under high pressures, but
barium forms barium peroxide just on normal heating in oxygen. Mixtures
of barium oxide and barium peroxide will be produced.

The strontium equation would look just the same.

The reactions with air


The reactions of the Group 2 metals with air rather than oxygen is
complicated by the fact that they all react with nitrogen to produce
nitrides. In each case, you will get a mixture of the metal oxide and the
metal nitride.
The general equation for the Group is:

The familiar white ash you get when you burn magnesium ribbon in air is
a mixture of magnesium oxide and magnesium nitride (despite what you
http://www.chemguide.co.uk/inorganic/group2/reacto2.html (3 of 9)30/12/2004 11:36:21

Reactions of the Group 2 elements with air or oxygen

might have been told when you were first learning Chemistry!).

The Explanations
Trying to pick out patterns in the way the metals burn
There are no simple patterns. It would be tempting to say that the
reactions get more vigorous as you go down the Group, but it isn't true.
The overall amount of heat evolved when one mole of oxide is produced
from the metal and oxygen shows no simple pattern:

If anything, there is a slight tendency for the amount of heat evolved to


get less as you go down the Group.
But how reactive a metal seems to be depends on how fast the reaction
happens - not the overall amount of heat evolved. The speed is
controlled by factors like the presence of surface coatings on the metal
and the size of the activation energy.
You could argue that the activation energy will fall as you go down the
Group and that will make the reaction go faster. The activation energy
will fall because the ionisation energies of the metals fall.

http://www.chemguide.co.uk/inorganic/group2/reacto2.html (4 of 9)30/12/2004 11:36:21

Reactions of the Group 2 elements with air or oxygen

Note: This has been argued through in detail on the page


about the reactions of these metals with water (or steam). If
you need to know about the reactions with oxygen, you will
almost certainly need to know about the reactions with water
as well.

In this case, though, the effect of the fall in the activation energy is
masked by other factors - for example, the presence of existing oxide
layers on the metals, and the impossibility of controlling precisely how
much heat you are supplying to the metal in order to get it to start
burning.

Note: It is interesting to look at what happens if you heat a


very reactive metal like potassium in air. The potassium melts
at a low temperature and almost instantly turns into a pool of
molten potassium oxide. The activation energy is so low that
the reaction happens very quickly at quite a low temperature.
There is often no trace of flame. It can be fairly boring!
Magnesium, on the other hand, has to be heated to quite a
high temperature before it will start to react. The activation
energy is much higher. There are also problems with surface
coatings. It is then so hot that it produces the typical intense
white flame.
It would obviously be totally misleading to say that
magnesium is more reactive than potassium on the evidence
of the bright flame. You haven't had to heat them by the
same amount to get the reactions happening.

http://www.chemguide.co.uk/inorganic/group2/reacto2.html (5 of 9)30/12/2004 11:36:21

Reactions of the Group 2 elements with air or oxygen

Why do some metals form peroxides on heating in oxygen?


Beryllium, magnesium and calcium don't form peroxides when heated in
oxygen, but strontium and barium do. There is an increase in the
tendency to form the peroxide as you go down the Group.
The peroxide ion, O22- looks llike this:

The covalent bond between the two oxygen atoms is relatively weak.
Now imagine bringing a small 2+ ion close to the peroxide ion. Electrons
in the peroxide ion will be strongly attracted towards the positive ion. This
is then well on the way to forming a simple oxide ion if the right-hand
oxygen atom (as drawn below) breaks off.

We say that the positive ion polarises the negative ion. This works best
if the positive ion is small and highly charged - if it has a high charge
density.

http://www.chemguide.co.uk/inorganic/group2/reacto2.html (6 of 9)30/12/2004 11:36:21

Reactions of the Group 2 elements with air or oxygen

Note: A high charge density simply means that you have a


lot of charge packed into a small volume.

Ions of the metals at the top of the Group have such a high charge
density (because they are so small) that any peroxide ion near them falls
to pieces to give an oxide and oxygen. As you go down the Group and
the positive ions get bigger, they don't have so much effect on the
peroxide ion.
Barium peroxide can form because the barium ion is so large that it
doesn't have such a devastating effect on the peroxide ions as the
metals further up the Group.

Why do these metals form nitrides on heating in air?


Nitrogen is often thought of as being fairly unreactive, and yet all these
metals combine with it to produce nitrides, X3N2, containing X2+ and N3ions.
Nitrogen is fairly unreactive because of the very large amount of energy
needed to break the triple bond joining the two atoms in the nitrogen
molecule, N2.
When something like magnesium nitride forms, you have to supply all the
energy needed to form the magnesium ions as well as breaking the
nitrogen-nitrogen bonds and then forming N3- ions. All of these
processes absorb energy.
This energy has to be recovered from somewhere to give an overall
exothermic reaction - if the energy can't be recovered, the overall change
will be endothermic and won't happen.

http://www.chemguide.co.uk/inorganic/group2/reacto2.html (7 of 9)30/12/2004 11:36:21

Reactions of the Group 2 elements with air or oxygen

Note: This is a bit of a simplification! In order to find out


whether a reaction is feasible, you have to consider free
energy changes and not just whether the reaction is
exothermic or endothermic. If you don't know anything about
free energy changes, don't worry about it. The simplification
is valid in this particular case.

Energy is evolved when the ions come together to produce the crystal
lattice. This energy is known as lattice energy or lattice enthalpy.
The size of the lattice energy depends on the attractions between the
ions. The lattice energy is greatest if the ions are small and highly
charged - the ions will be close together with very strong attractions. In
the whole of Group 2, the attractions between the 2+ metal ions and the
3- nitride ions are big enough to produce very high lattice energies.
When the crystal lattices form, so much energy is released that it more
than compensates for the energy needed to produce the various ions in
the first place. The excess energy evolved makes the overall process
exothermic.
This is in contrast to what happens in Group 1 of the Periodic Table
(lithium, sodium, potassium, rubidium and caesium). Their ions only carry
one positive charge, and so the lattice energies of their nitrides will be
much less.
Lithium is the only metal in Group 1 to form a nitride. Lithium has by far
the smallest ion in the Group, and so lithium nitride has the largest lattice
energy of any possible Group 1 nitride. Only in lithium's case is enough
energy released to compensate for the energy needed to ionise the
metal and the nitrogen - and so produce an exothermic reaction overall.
In all the other cases in Group 1, the overall reaction would be
endothermic. Those reactions don't happen, and the nitrides of sodium
and the rest aren't formed.

Where would you like to go now?


http://www.chemguide.co.uk/inorganic/group2/reacto2.html (8 of 9)30/12/2004 11:36:21

Reactions of the Group 2 elements with air or oxygen

To the Group 2 menu . . .


To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/inorganic/group2/reacto2.html (9 of 9)30/12/2004 11:36:21

Solubility of the hydroxides, sulphates and carbonates of the Group 2 elements in water

SOLUBILITY OF THE HYDROXIDES,


SULPHATES AND CARBONATES OF THE
GROUP 2 ELEMENTS IN WATER

This page looks at the solubility in water of the hydroxides, sulphates and
carbonates of the Group 2 elements - beryllium, magnesium, calcium,
strontium and barium. Although it describes the trends, there isn't any
attempt to explain them on this page - for reasons discussed later.
You will find that there aren't any figures given for any of the solubilities.
There are major discrepancies between the figures given by two
common A'level Data Books (Nuffield Advanced Science Book of Data,
and Chemistry Data Book by Stark and Wallace). There are also
important inconsistencies within the books (one set of figures doesn't
agree with those which can be calculated from another set). I haven't
been able to find data which I am sure is correct, and therefore prefer not
to give any.

The Facts
Solubility of the hydroxides

The hydroxides become more soluble as you go down the Group.

This is a trend which holds for the whole Group, and applies whichever
set of data you choose.
Some examples may help you to remember the trend:
Magnesium hydroxide appears to be insoluble in water. However, if you
shake it with water, filter it and test the pH of the solution, you find that it
is slightly alkaline. This shows that there are more hydroxide ions in the
solution than there were in the original water. Some magnesium
hydroxide must have dissolved.
Calcium hydroxide solution is used as "lime water". 1 litre of pure water
http://www.chemguide.co.uk/inorganic/group2/solubility.html (1 of 4)30/12/2004 11:36:24

Solubility of the hydroxides, sulphates and carbonates of the Group 2 elements in water

will dissolve about 1 gram of calcium hydroxide at room temperature.


Barium hydroxide is soluble enough to be able to produce a solution
with a concentration of around 0.1 mol dm-3 at room temperature.

Solubility of the sulphates

The sulphates become less soluble as you go down the Group.

The simple trend is true provided you include hydrated beryllium sulphate
in it, but not if the beryllium sulphate is anhydrous.
The Nuffield Data Book quotes anyhydrous beryllium sulphate, BeSO4,
as insoluble (I haven't been able to confirm this from any other source),
whereas the hydrated form, BeSO4.4H2O is soluble. (The Data Books
agree on this - giving a figure of about 39 g dissolving in 100 g of water
at room temperature.)
Figures for magnesium sulphate and calcium sulphate also vary
depending on whether the salt is hydrated or not, but nothing like so
dramatically.
Two common examples may help you to remember the trend:
You are probably familiar with the reaction between magnesium and
dilute sulphuric acid to give lots of hydrogen and a colourless solution of
magnesium sulphate. Notice that you get a solution, not a precipitate.
The magnesium sulphate is obviously soluble.
You may also remember that barium sulphate is formed as a white
precipitate during the test for sulphate ions in solution. The ready
formation of a precipitate shows that the barium sulphate must be pretty
insoluble. In fact, 1 litre of water will only dissolve about 2 mg of barium
sulphate at room temperature.
Solubility of the carbonates

The carbonates tend to become less soluble as you go down the


Group.

http://www.chemguide.co.uk/inorganic/group2/solubility.html (2 of 4)30/12/2004 11:36:24

Solubility of the hydroxides, sulphates and carbonates of the Group 2 elements in water

None of the carbonates is anything more than very sparingly soluble.


Magnesium carbonate (the most soluble one I have data for) is soluble to
the extent of about 0.02 g per 100 g of water at room temperature.
I can't find any data for beryllium carbonate, but it tends to react with
water and so that might confuse the trend.
The trend to lower solubility is, however, broken at the bottom of the
Group. Barium carbonate is slightly more soluble than strontium sulphate.
There are no simple examples which might help you to remember the
carbonate trend.

What - no explanations?
Before I started to write this page, I thought I understood the trends in
solubility patterns including the explanations for them. The more I have
dug around to try to find reliable data, and the more time I have spent
thinking about it, the less I'm sure that it is possible to come up with any
simple explanation of the solubility patterns.

Note: If you are interested in the reasons why I am unwilling


to give the usual over-simplified explanations, I have
described some of the problems as I see them on a separate
page. That page also includes an attempt at a better
explanation. Unless your syllabus specifically asks for
explanations of these trends, you would be better off ignoring
this follow-up page!

http://www.chemguide.co.uk/inorganic/group2/solubility.html (3 of 4)30/12/2004 11:36:24

Solubility of the hydroxides, sulphates and carbonates of the Group 2 elements in water

Where would you like to go now?


To an attempt to explain these trends . . .
To the Group 2 menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/inorganic/group2/solubility.html (4 of 4)30/12/2004 11:36:24

Problems in explaining the solubility of Group 2 compounds

EXPLANATIONS FOR THE TRENDS IN


SOLUBILITY OF SOME GROUP 2 COMPOUNDS

This page looks at the usual explanations for the solubility patterns in the
hydroxides, sulphates and carbonates of Group 2. It goes on to look at
my misgivings about these. This page is probably best avoided unless
your syllabus specifically asks for these explanations!

Warning: If you have come straight to this page via a search


engine, you should be aware that it is a follow-up to another
page describing the solubility patterns. You should read that
page before going on with this one.

The usual explanations


Enthalpy changes during the process
The usual explanation is in terms of the enthalpy changes which occur
when an ionic compound dissolves in water. Energy has to be supplied
to break up the lattice of ions, and energy is released when these ions
form bonds of one sort or another with water molecules.
As you go down a Group, the energy needed to break up the lattice falls
as the positive ions get bigger. The bigger the ions, the more distance
there is between them, and the weaker the forces holding them together.
Again as the positive ions get bigger, the energy released as the ions
bond to water molecules (their hydration enthalpies) falls as well. Bigger
ions aren't so strongly attracted to the water molecules.
Since both of these important enthalpy terms fall as you go down the
Group, what matters in deciding whether the change becomes more
endothermic or more exothermic overall is how fast they fall relative to
http://www.chemguide.co.uk/inorganic/group2/problems.html (1 of 12)30/12/2004 11:36:31

Problems in explaining the solubility of Group 2 compounds

each other.
A sample calculation to make this clear
Here is an example of the sort of calculations you might do to work out
the enthalpy change of solution for sodium chloride and potassium
chloride.

Note: You may wonder why I haven't used examples directly


relevant to the hydroxides, sulphates or carbonates of Group
2. I can't find any lattice enthalpy data for them, or hydration
enthalpies for the sulphate or carbonate ions!
The data used comes from Chemistry Data Book by Stark
and Wallace. The Nuffield Data Book doesn't have any
hydration enthalpy values.

In this case, we are defining lattice enthalpy as the heat needed to


convert 1 mole of crystal in its standard state into separate gaseous ions
- an endothermic change.

Note: Lattice enthalpy is more usually defined in the


opposite direction - as the heat evolved when the lattice is
formed from its gaseous ions. Defining it in the endothermic
direction makes the argument a bit easier in this particular
case. This endothermic change involving the break up of the
lattice is properly called "lattice dissociation enthalpy",
although the simpler term "lattice enthalpy" is still often used
for it.

http://www.chemguide.co.uk/inorganic/group2/problems.html (2 of 12)30/12/2004 11:36:31

Problems in explaining the solubility of Group 2 compounds

You can see that the enthalpy of solution changes from NaCl to KCl
because the lattice enthalpy and hydration enthalpy of the positive ion fall
by different amounts.

If the hydration enthalpy falls faster than the lattice enthalpy (as in
this case), the net effect is that the overall change becomes more
endothermic (or less exothermic in other possible cases where the
total enthalpy change turns out to be negative).
If the lattice enthalpy falls faster than the hydration enthalpy, the
opposite happens - the change will become less endothermic (or
more exothermic).

What controls the relative rate of fall of the two terms?


It turns out that the main factor is the size of the negative ion.

For small negative ions like hydroxide, the lattice enthalpy falls
faster than the hydration enthalpy of the positive ions. That means
that the enthalpy of solution will become less positive (or more
negative).
For large negative ions like sulphate or carbonate, the hydration
enthalpy of the positive ions falls faster than the lattice enthalpy. In
this case, the enthalpy of solution will become more positive (or
less negative).

Why the difference? Lattice enthalpy is governed by several factors http://www.chemguide.co.uk/inorganic/group2/problems.html (3 of 12)30/12/2004 11:36:31

Problems in explaining the solubility of Group 2 compounds

including the distance between the negative and positive ions. Where
you have a big negative ion, this inter-ionic distance is largely controlled
by the size of that negative ion. Changes in the size of the positive ion
don't make as great a percentage difference to the inter-ionic distance as
they would if the negative ion was small.
With sulphates, for example, the percentage increase in the inter-ionic
distance as you go from magnesium to calcium sulphate isn't as great as
it would be with a smaller negative ion like hydroxide.
Since the percentage increase in inter-ionic distance isn't very great, the
change in the lattice enthalpy won't be very great either.
The relationship between enthalpy of solution and solubility
The assumption is made that the more endothermic (or less exothermic)
the enthalpy of solution is, the less soluble the compound.
So sulphates and carbonates become less soluble as you go down the
Group; hydroxides become more soluble.

Problems with the usual explanations


Problems with the data
Unfortunately, the enthalpy of solution values for the Group 1 chlorides
as calculated above don't agree with the values given in the same Data
Book:

Calculated value
(kJ mol-1)

Data book value


(kJ mol-1)

NaCl

+1

+3.9

KCl

+15

+17.2

http://www.chemguide.co.uk/inorganic/group2/problems.html (4 of 12)30/12/2004 11:36:31

Problems in explaining the solubility of Group 2 compounds

RbCl

+10

+16.7

CsCl

+5

+17.9

The discrepancies are enough to disrupt any pattern (such as there is!).
The reasons for the discrepancies lie in the way the numbers are
calculated. These are very small numbers worked out from much larger
ones. Small uncertainties in those large numbers will cause large swings
in the answers.
For example, if each of the numbers in the calculations we did earlier on
this page was out by just 5 kJ, each answer could vary by +/- 15 kJ completely disrupting the patterns!
You can't therefore reliably use the data available to calculate the trends
you want with sufficient accuracy to make sense.

Problems correlating enthalpy data with the facts


The solubilities of the Group 1 chlorides (in moles of solute saturating
100 g of water at 298 K) compared with their enthalpies of solution are:

Enthalpy of
solution
(kJ mol-1)

Solubility
(mol /100 g of water)

NaCl

+3.9

0.615

KCl

+17.2

0.481

RbCl

+16.7

0.781

CsCl

+17.9

1.13

http://www.chemguide.co.uk/inorganic/group2/problems.html (5 of 12)30/12/2004 11:36:31

Problems in explaining the solubility of Group 2 compounds

There is no obvious relationship connecting the relative movements of


these solubility values with the enthalpy of solution figures. It would be
quite untrue to say that the more endothermic the change, the less
soluble the compound!

Problems in relating the sign of the enthalpy change to solubility


The table above illustrates this problem, but it gets worse! Taking the
sign of enthalpy of solution at face value, you get some bizarre results.
The enthalpy of solution figures for the Group 2 carbonates are: (source:
Chemistry Data Book by Stark and Wallace; values in kJ mol-1)

MgCO3

-25.3

CaCO3

-12.3

SrCO3

-3.4

BaCO3

+4.2

Sodium chloride and the other Group 1 chlorides dissolve despite the
fact that their enthalpies of solution are positive, and yet magnesium
carbonate (and most of the other Group 2 carbonates) are very sparingly
soluble, but have exothermic enthalpies of solution. You might have
expected exactly the opposite to happen.
However if you ignore the comparison with the Group 1 chlorides, you
could argue that the figures get progressively less exothermic, and at
barium carbonate become endothermic. That would seem to support the
decrease in solubility as you go down the Group quite nicely.
Unfortunately, if you look at the solubility data, the trend is broken at the
bottom of the Group. Barium carbonate is more soluble than strontium
carbonate! Although figures from my two data sources differ in detail,
they agree on this.
http://www.chemguide.co.uk/inorganic/group2/problems.html (6 of 12)30/12/2004 11:36:31

Problems in explaining the solubility of Group 2 compounds

Clearly, trying to correlate solubility simply with the enthalpy change of


solution doesn't work.

Introducing entropy changes


The basic explanation
To get around the problem of many compounds dissolving freely in water
despite the fact that their enthalpies of solution are endothermic you
have to introduce the concept of entropy change.
The only way of making sense of entropy without getting bogged down in
some serious maths is to think of it as a measure of the amount of
disorder in a system. Entropy is given the symbol S. If a system becomes
more disordered, then its entropy increases.
In order to see whether a change is possible or not, you have to think
about a combination of the enthalpy change and the entropy change.
These can be combined mathematically to give an important term known
as free energy change.

As an approximation, for a reaction to happen, the free energy change


must be negative.
What happens if the enthalpy change is positive - as for example when
sodium chloride dissolves in water (+3.9 kJ mol-1, using the values in one
of the tables above)?

http://www.chemguide.co.uk/inorganic/group2/problems.html (7 of 12)30/12/2004 11:36:31

Problems in explaining the solubility of Group 2 compounds

As long as the entropy change is positive enough, it is possible to get a


negative value for free energy change. In the sodium chloride case, you
don't have to have very much increase in entropy to outweigh the small
enthalpy change of +3.9 kJ mol-1.

So . . . does the entropy increase when sodium chloride dissolve in


water? Yes, it does!
Originally, the sodium and chloride ions were arranged in a very tidy way
in the crystal lattice - their entropy was low. (Remember that entropy is a
measure of disorder.) When you dissolve the crystal in water, the entropy
increases as the ions and water molecules become completely jumbled
up - they become much more disordered than they were originally.
This is where the explanation usually stops, but to stop at this point is
very misleading because it won't explain all the facts!

What's wrong with this explanation?


The most obvious thing that's wrong is that it won't explain why some
compounds (like magnesium carbonate, and most of the other Group 2
carbonates) don't dissolve in water even though their enthalpies of
solution are mainly negative.
In these cases, the entropy of the system must fall when the compounds
dissolve in water - in other words, the solution in water is more ordered
than the original crystal and water!

http://www.chemguide.co.uk/inorganic/group2/problems.html (8 of 12)30/12/2004 11:36:31

Problems in explaining the solubility of Group 2 compounds

This happens because the water molecules become more ordered when
the compound dissolves in them. Instead of milling around pretty much at
random, they become attracted to the ions present and arranged around
them.
This is particularly effective if the ions are small and highly charged - and
so the effect is greatest for the positive ions at the top of the Group, and
gets less as you go down. It is also much more important in Group 2 than
in Group 1 where the ions only carry one positive charge.
That means that you have two entropy effects to consider. There is the
increase in disorder as the crystal lattice breaks up, but a corresponding
increase in order in the water - which varies depending on the sizes and
charges of the ions present.
What a nightmare!
To explain this properly, you need to think about the way lattice enthalpy
changes as you go down the Group, the way that hydration enthalpies
change, and the way that entropy changes. The way those changes
happen will vary from one type of compound to another.
Can we explain everything now?
No - at least not easily! For example, although it might be possible to
account for the lack of pattern in the solubilities of the Group 1 chlorides
(and also the bromides) by a mathematical application of these effects,
trying to do it in general terms defeats me completely!

http://www.chemguide.co.uk/inorganic/group2/problems.html (9 of 12)30/12/2004 11:36:31

Problems in explaining the solubility of Group 2 compounds

You could, however, make a reasonable suggestion as to why the


solubility trend in the carbonates is broken at barium. (Don't expect the
explanation to be instantly understandable though!)
Remember that the solubility of the carbonates falls as you go down
Group 2, apart from an increase as you go from strontium to barium
carbonate.
The general fall is because hydration enthalpies are falling faster than
lattice enthalpies. Remember that where you have a big negative ion, its
size dominates the inter-ionic distance and so doesn't allow the lattice
enthalpy to change much. This gives the enthalpy of solution values
we've already looked at (values in kJ mol-1):

MgCO3

-25.3

CaCO3

-12.3

SrCO3

-3.4

BaCO3

+4.2

But the entropy change will also be varying as you go down the Group.
At the top, where you have small 2+ ions, the overall entropy change in
the system must be negative - the system as a whole becomes more
ordered when the compound dissolves because of the way the water
molecules become organised around the positive ions. That negative
entropy change is going to be enough to wipe out the effect of the
exothermic enthalpy of solution.

Note: You need to think about the effect of the positiveness


and negativeness of these values on the size of the free
energy change. Use the equation further up this page.

http://www.chemguide.co.uk/inorganic/group2/problems.html (10 of 12)30/12/2004 11:36:31

Problems in explaining the solubility of Group 2 compounds

Towards the bottom of the Group, this effect changes. The bigger ions
have less organising effect on the water molecules. The entropy change
is becoming less negative (or perhaps even at this stage, positive).
That's going to tend to make the compounds more soluble.
The overall effect is a complex balance between the way the enthalpy of
solution varies and the way the entropy change of solution alters. At
barium carbonate, the effect of increasing entropy must be enough to
make it more soluble than strontium carbonate.

A personal comment: I see it as quite wrong that this


should be discussed at A' level. The usual explanations are
over-simplistic and potentially misleading - especially for
anyone who might go on to do Chemistry at degree level.
Fortunately most syllabuses have enough sense not to ask
for these explanations!
Science progresses by offering theories which have to
explain all the facts. Where a fact won't fit a theory, the theory
has to be modified, or even discarded. What we seem to be
doing here is presenting students with an inadequate theory
and then ignoring all the facts which don't fit it. I see that as
quite dangerous. It would be much better not to discuss this
at all at this level, rather than to give students a false view of
the way science works.

Where would you like to go now?


Return to the page outlining trends in solubility . . .
To the Group 2 menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

http://www.chemguide.co.uk/inorganic/group2/problems.html (11 of 12)30/12/2004 11:36:31

Problems in explaining the solubility of Group 2 compounds

Jim Clark 2002

http://www.chemguide.co.uk/inorganic/group2/problems.html (12 of 12)30/12/2004 11:36:31

Thermal decomposition of the Group 2 carbonates and nitrates

THERMAL STABILITY OF THE GROUP 2


CARBONATES AND NITRATES

This page looks at the effect of heat on the carbonates and nitrates of the
Group 2 elements - beryllium, magnesium, calcium, strontium and
barium. It describes and explains how the thermal stability of the
compounds changes as you go down the Group.

The Facts
The effect of heat on the Group 2 carbonates
All the carbonates in this Group undergo thermal decomposition to give
the metal oxide and carbon dioxide. Thermal decomposition is the term
given to splitting up a compound by heating it.
If "X" represents any one of the elements:

As you go down the Group, the carbonates have to be heated more


strongly before they will decompose.

The carbonates become more stable to heat as you go down the


Group.

The effect of heat on the Group 2 nitrates


All the nitrates in this Group undergo thermal decomposition to give the
metal oxide, nitrogen dioxide and oxygen.
Again, if "X" represents any one of the elements:

As you go down the Group, the nitrates also have to be heated more
http://www.chemguide.co.uk/inorganic/group2/thermstab.html (1 of 10)30/12/2004 11:36:44

Thermal decomposition of the Group 2 carbonates and nitrates

strongly before they will decompose.

The nitrates also become more stable to heat as you go down the
Group.

Summary
Both carbonates and nitrates become more thermally stable as you go
down the Group. The ones lower down have to be heated more strongly
than those at the top before they will decompose.

Explanations
This page offers two different ways of looking at the problem. You need
to find out which of these your examiners are likely to expect from you so
that you don't get involved in more difficult things than you actually need.
You should look at your syllabus, and past exam papers - together with
their mark schemes

Note: If you haven't got copies of your syllabus and past


papers follow this link to find out how to get hold of them.

Detailed explanations are given for the carbonates because the diagrams
are easier to draw, and their equations are also easier. Exactly the same
arguments apply to the nitrates.

Explaining the trend in terms of the polarising ability of the positive


ion
A small 2+ ion has a lot of charge packed into a small volume of space. It
has a high charge density and will have a marked distorting effect on any
negative ions which happen to be near it.
A bigger 2+ ion has the same charge spread over a larger volume of
http://www.chemguide.co.uk/inorganic/group2/thermstab.html (2 of 10)30/12/2004 11:36:44

Thermal decomposition of the Group 2 carbonates and nitrates

space. Its charge density will be lower, and it will cause less distortion to
nearby negative ions.
The structure of the carbonate ion
If you worked out the structure of a carbonate ion using "dots-andcrosses" or some similar method, you would probably come up with:

This shows two single carbon-oxygen bonds and one double one, with
two of the oxygens each carrying a negative charge. Unfortunately, in
real carbonate ions all the bonds are identical, and the charges are
spread out over the whole ion - although concentrated on the oxygen
atoms. We say that the charges are delocalised.
This is a rather more complicated version of the bonding you might have
come across in benzene or in ions like ethanoate. For the purposes of
this topic, you don't need to understand how this bonding has come
about.

Note: If you are interested, you could follow these links to


benzene or to organic acids. Either of these links is likely to
involve you in a fairly time-consuming detour!

http://www.chemguide.co.uk/inorganic/group2/thermstab.html (3 of 10)30/12/2004 11:36:44

Thermal decomposition of the Group 2 carbonates and nitrates

The next diagram shows the delocalised electrons. The shading is


intended to show that there is a greater chance of finding them around
the oxygen atoms than near the carbon.

Polarising the carbonate ion


Now imagine what happens when this ion is placed next to a positive ion.
The positive ion attracts the delocalised electrons in the carbonate ion
towards itself. The carbonate ion becomes polarised.

If this is heated, the carbon dioxide breaks free to leave the metal oxide.
How much you need to heat the carbonate before that happens depends
on how polarised the ion was. if it is highly polarised, you need less heat
than if it is only slightly polarised.
The smaller the positive ion is, the higher the charge density, and the
greater effect it will have on the carbonate ion. As the positive ions get
bigger as you go down the Group, they have less effect on the carbonate
ions near them. To compensate for that, you have to heat the compound
more in order to persuade the carbon dioxide to break free and leave the
metal oxide.
http://www.chemguide.co.uk/inorganic/group2/thermstab.html (4 of 10)30/12/2004 11:36:44

Thermal decomposition of the Group 2 carbonates and nitrates

In other words, as you go down the Group, the carbonates become more
thermally stable.
What about the nitrates?
The argument is exactly the same here. The small positive ions at the top
of the Group polarise the nitrate ions more than the larger positive ions at
the bottom. Drawing diagrams to show this happening is much more
difficult because the process has interactions involving more than one
nitrate ion. You wouldn't be expected to attempt to draw this in an exam.

Explaining the trend in terms of the energetics of the process


Looking at the enthalpy changes
If you calculate the enthalpy changes for the decompostion of the various
carbonates, you find that all the changes are quite strongly endothermic.
That implies that the reactions are likely to have to be heated constantly
to make them happen.
The calculated enthalpy changes (in kJ mol-1) are given in the table.
Figures to calculate the beryllium carbonate value weren't available.
Remember that the reaction we are talking about is:

MgCO3

+117

CaCO3

+178

SrCO3

+235

BaCO3

+267

You can see that the reactions become more endothermic as you go

http://www.chemguide.co.uk/inorganic/group2/thermstab.html (5 of 10)30/12/2004 11:36:44

Thermal decomposition of the Group 2 carbonates and nitrates

down the Group. That's entirely what you would expect as the
carbonates become more thermally stable. You have to supply
increasing amounts of heat energy to make them decompose.
Explaining the enthalpy changes
Here's where things start to get difficult! If you aren't familiar with Hess's
Law cycles (or with Born-Haber cycles) and with lattice enthalpies (lattice
energies), you aren't going to understand the next bit. Don't waste your
time looking at it.

Note: If you should know about these cycles, but don't really
understand them, you might be interested in my book on
A'level calculations. What you need to know and be able to
do is covered in a lot of detail.

Using an enthalpy cycle


You can dig around to find the underlying causes of the increasingly
endothermic changes as you go down the Group by drawing an enthalpy
cycle involving the lattice enthalpies of the metal carbonates and the
metal oxides.
Confusingly, there are two ways of defining lattice enthalpy. In order to
make the argument mathematically simpler, during the rest of this page I
am going to use the less common version (as far as the A' level
syllabuses are concerned):
Lattice enthalpy is the heat needed to split one mole of crystal in its
standard state into its separate gaseous ions. For example, for
magnesium oxide, it is the heat needed to carry out 1 mole of this
change:

http://www.chemguide.co.uk/inorganic/group2/thermstab.html (6 of 10)30/12/2004 11:36:44

Thermal decomposition of the Group 2 carbonates and nitrates

Note: Lattice enthalpy is more usually defined as the heat


evolved when 1 mole of crystal is formed from its gaseous
ions. In that case, the lattice enthalpy for magnesium oxide
would be -3889 kJ mol-1. The term we are using here should
more accurately be called the "lattice dissociation enthalpy".

The cycle we are interested in looks like this:

You can apply Hess's Law to this, and find two routes which will have an
equal enthalpy change because they start and end in the same places.

http://www.chemguide.co.uk/inorganic/group2/thermstab.html (7 of 10)30/12/2004 11:36:44

Thermal decomposition of the Group 2 carbonates and nitrates

For reasons we will look at shortly, the lattice enthalpies of both the
oxides and carbonates fall as you go down the Group. But they don't fall
at the same rate.
The oxide lattice enthalpy falls faster than the carbonate one. If you think
carefully about what happens to the value of the overall enthalpy change
of the decomposition reaction, you will see that it gradually becomes
more positive as you go down the Group.

Explaining the relative falls in lattice enthalpy


The size of the lattice enthalpy is governed by several factors, one of
which is the distance between the centres of the positive and negative
ions in the lattice. Forces of attraction are greatest if the distances
between the ions are small. If the attractions are large, then a lot of
energy will have to be used to separate the ions - the lattice enthalpy will
be large.
http://www.chemguide.co.uk/inorganic/group2/thermstab.html (8 of 10)30/12/2004 11:36:44

Thermal decomposition of the Group 2 carbonates and nitrates

The lattice enthalpies of both carbonates and oxides fall as you go down
the Group because the positive ions are getting bigger. The inter-ionic
distances are increasing and so the attractions become weaker.

ionic radius (nm)


Mg2+

0.065

Ca2+

0.099

O2-

0.140

CO32-

The lattice enthalpies fall at different rates because of the different sizes
of the two negative ions - oxide and carbonate. The oxide ion is relatively
small for a negative ion (0.140 nm), whereas the carbonate ion is large
(no figure available).
In the oxides, when you go from magnesium oxide to calcium oxide, for
example, the inter-ionic distance increases from 0.205 nm (0.140 +
0.065) to 0.239 nm (0.140 + 0.099) - an increase of about 17%.
In the carbonates, the inter-ionic distance is dominated by the much
larger carbonate ion. Although the inter-ionic distance will increase by the
same amount as you go from magnesium carbonate to calcium
carbonate, as a percentage of the total distance the increase will be
much less.
Some made-up figures show this clearly.
I can't find a value for the radius of a carbonate ion, and so can't use real
figures. For the sake of argument, suppose that the carbonate ion radius
was 0.3 nm. The inter-ionic distances in the two cases we are talking
about would increase from 0.365 nm to 0.399 nm - an increase of only
about 9%.

http://www.chemguide.co.uk/inorganic/group2/thermstab.html (9 of 10)30/12/2004 11:36:44

Thermal decomposition of the Group 2 carbonates and nitrates

The rates at which the two lattice energies fall as you go down the Group
depends on the percentage change as you go from one compound to the
next. On that basis, the oxide lattice enthalpies are bound to fall faster
than those of the carbonates.
What about the nitrates?
The nitrate ion is bigger than an oxide ion, and so its radius tends to
dominate the inter-ionic distance. The lattice enthalpy of the oxide will
again fall faster than the nitrate. if you constructed a cycle like that further
up the page, the same arguments would apply.

Where would you like to go now?


To the Group 2 menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/inorganic/group2/thermstab.html (10 of 10)30/12/2004 11:36:44

Chemistry of beryllium untypical of Group 2

SOME BERYLLIUM CHEMISTRY UNTYPICAL OF


GROUP 2

This page describes and explains three examples from beryllium


chemistry where it behaves differently from the rest of Group 2.

Beryllium chloride is covalent


The facts
Physical properties
Beryllium chloride, BeCl2, melts at 405C and boils at 520C. That
compares with 714C and 1412C for magnesium chloride.
Notice how much dramatically lower the boiling point of beryllium chloride
is compared with magnesium chloride. The much higher boiling point of
magnesium chloride is what you might expect from the strong forces
between the positive and negative ions present.
Because its boiling point is much lower, it follows that beryllium chloride
can't contain ions - it must be covalent. On the other hand, the melting
point is quite high for such a small covalent molecule. There must be
something more complicated going on!
Reaction with water
Beryllium chloride reacts vigorously and exothermically with water with
the evolution of acidic, steamy hydrogen chloride gas. This is typical of
covalent chlorides.
In the first instance, it reacts to give hydrated beryllium ions, [Be(H2O)4]2
+,

and chloride ions.

But the hydrated beryllium ions (called tetraaquaberyllium ions) are quite

http://www.chemguide.co.uk/inorganic/group2/beryllium.html (1 of 12)30/12/2004 11:36:54

Chemistry of beryllium untypical of Group 2

strongly acidic. The small beryllium ion at the centre attracts the
electrons in the bonds towards itself, and that makes the hydrogen atoms
in the water even more positive than they usually are. If the solution is
hot and concentrated (as it is likely to be if you add water to solid
beryllium chloride - a very exothermic reaction), chloride ions can remove
one or more of these hydrogen ions to produce hydrogen chloride gas.
All the other ionic chlorides in Group 2 dissolve in water without any
obvious reaction.

Note: There is actually a very small amount of reaction


between anhydrous magnesium chloride and water, although
you wouldn't notice it. Magnesium chloride isn't quite as
purely ionic as we sometimes pretend!
Follow this link if you are interested in exploring the naming of
complex ions.
. . . or this one for detailed explanations of why complex ions
similar to the beryllium one are acidic.

The structure of beryllium chloride


As a gas . . .
Beryllium chloride, BeCl2, is a linear molecule with all three atoms in a
straight line. Showing only the outer electrons:

http://www.chemguide.co.uk/inorganic/group2/beryllium.html (2 of 12)30/12/2004 11:36:54

Chemistry of beryllium untypical of Group 2

Beryllium chloride is known as an electron-deficient compound


because it has the two empty orbitals at the bonding level.
As a solid . . .
If it had this same simple structure as a solid, you would expect the
melting point to be much lower than it actually is. It is a very small
molecule, and so the intermolecular attractions would be expected to be
fairly weak.

Note: If you aren't sure about intermolecular forces (Van der


Waals forces) you could follow this link to find out more about
them.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/inorganic/group2/beryllium.html (3 of 12)30/12/2004 11:36:54

Chemistry of beryllium untypical of Group 2

In the solid, the BeCl2 molecules polymerise to make long chains. They
do this by forming coordinate bonds (dative covalent bonds) between
lone pairs on chlorine atoms and adjacent beryllium atoms. The diagram
shows a simple dimer - the start of the polymerisation process.

Note: This is exactly the same as the coordinate bonding in


aluminium chloride. If you aren't happy about coordinate
(dative covalent) bonding, it would pay you to look at this
page before you go on.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/inorganic/group2/beryllium.html (4 of 12)30/12/2004 11:36:54

Chemistry of beryllium untypical of Group 2

You can see from the last diagram that the beryllium atoms are still
electron deficient. The process continues. The next diagram shows the
coordinate bonds in the conventional way using arrows. The arrow goes
from the atom which is supplying the pair of electrons to the atom with
the empty orbital.

Make sure that you fully understand how this diagram relates to the
dimer shown in the previous diagram.

Why isn't beryllium chloride ionic?


Beryllium has quite a high electronegativity compared with the rest of the
Group. That means that it attracts a bonding pair of electrons towards
itself more strongly than magnesium and the rest do.
In order for an ionic bond to form, the beryllium has to let go of its
electrons. It is too electronegative to do that.

Note: The trends in electronegativity in Group 2 are


discussed on another page. That page looks at the way the
electons are arranged in the beryllium-chlorine bond
compared with the magnesium-chlorine bond.
Use the BACK button on your browser to return to this page or come back via the Group 2 menu.

http://www.chemguide.co.uk/inorganic/group2/beryllium.html (5 of 12)30/12/2004 11:36:54

Chemistry of beryllium untypical of Group 2

Beryllium forms 4-coordinated complex ions


Some simple background
Although beryllium doesn't normally form simple ions, Be2+, it does form
ions in solution. In these, the beryllium ion becomes attached to four
water molecules to give a complex ion with the formula [Be(H2O)4]2+.
The ion is said to be 4-coordinated, or to have a coordination number
of 4, because there are four water molecules arranged around the
central beryllium.
Many hydrated metal ions are 6-coordinated. For example, magnesium
ions in solution exist as [Mg(H2O)6]2+.
The water molecules in these ions are attached to the central metal ion
via coordinate bonds (dative covalent bonds). One of the lone pairs on
each water molecule is used to form a bond with an empty orbital in the
metal ion.
Each time one of these bonds is formed, energy is released, and the ion
becomes more stable. It would seem logical for the metal ion to form as
many bonds like this as it possibly can.

Note: If you aren't happy about coordinate bonding you must


follow this link before you go on. You will find the bonding in
hydrated metal ions discussed in some detail on that page.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/inorganic/group2/beryllium.html (6 of 12)30/12/2004 11:36:54

Chemistry of beryllium untypical of Group 2

Why does beryllium only attach four water molecules?


The hydration of beryllium
The problem is that there has to be somewhere that the lone pairs on the
water molecules can attach to. Beryllium has the electronic structure
1s22s2. It is helpful to draw this as an "electrons-in-boxes" diagram:

Note: If you aren't happy about orbitals you really ought to


follow this link before you go on. You may want to explore
further in that part of the site as well. Unless you understand
exactly what this electrons-in-boxes diagram is about, you
won't be able to make sense of what is coming up next.

When beryllium forms a 2+ ion it loses the 2 electrons in the 2s orbital.


That leaves the 2-level completely empty.
The 2-level orbitals reorganise themselves (hybridise) to make four equal
orbitals, each of which can accept a lone pair of electrons from a water
molecule. In the next diagram the 1s electrons have been left out. They
aren't relevant to the bonding.

http://www.chemguide.co.uk/inorganic/group2/beryllium.html (7 of 12)30/12/2004 11:36:54

Chemistry of beryllium untypical of Group 2

Each water molecule, of course, has two lone pairs of electrons. Only
one of them is shown to avoid cluttering the diagram.
Notice that once four water molecules have bonded in this way, there
isn't any more space available at the bonding level. All the empty orbitals
from the original beryllium ion are being used.
The water molecules arrange themselves to get as far apart as possible which is pointing towards the corners of a tetrahedron. The ion therefore
has a tetrahedral shape.
The hydration of magnesium
You might think that magnesium would behave just the same, but at the
3-level there are 3d orbitals available as well as 3s and 3p.
When the magnesium ion is formed, it leaves empty 3s, 3p and 3d
orbitals. When that ion is hydrated, it uses the 3s orbital, all three of the
3p orbitals and two of the 3d orbitals. These are reorganised to leave a
total of six empty orbitals which are then used for bonding.

http://www.chemguide.co.uk/inorganic/group2/beryllium.html (8 of 12)30/12/2004 11:36:54

Chemistry of beryllium untypical of Group 2

Why does magnesium stop at attaching six waters? Why doesn't it use
the remaining 3d orbitals as well? You can't physically fit more than six
water molecules around the magnesium - they take up too much room.
What about the other ions in Group 2?
As the ions get bigger, there is less tendency for them to form proper
coordinate bonds with water molecules. The ions become so big that
they aren't sufficiently attractive to the lone pairs on the water molecules
to form formal bonds - instead the water molecules tend to cluster more
loosely around the positive ions.
Where they do form coordinate bonds with the water, however, they will
be 6-coordinated just like the magnesium.

Beryllium hydroxide is amphoteric


Amphoteric means that it can react with both acids and bases to form
salts.
The other Group 2 hydroxides
The other hydroxides of the Group 2 metals are all basic. They react with
acids to form salts. For example:

Calcium hydroxide reacts with dilute hydrochloric acid to give calcium


http://www.chemguide.co.uk/inorganic/group2/beryllium.html (9 of 12)30/12/2004 11:36:54

Chemistry of beryllium untypical of Group 2

chloride and water.


Beryllium hydroxide
Beryllium hydroxide reacts with acids, forming solutions of beryllium
salts. For example:

But it also reacts with bases such as sodium hydroxide solution.


Beryllium hydroxide reacts with the sodium hydroxide to give a colourless
solution of sodium tetrahydroxoberyllate.

This contains the complex ion, [Be(OH)4]2-. The name describes this ion.
Tetra means four; hydroxo refers to the OH groups; beryllate shows that
the beryllium is present in a negative ion. The "ate" ending always shows
that the ion is negative.
A simple explanation of what is happening
You need to think about where the beryllium hydroxide came from in the
first place. It would probably have been made by adding sodium
hydroxide solution to a solution of a beryllium salt like beryllium sulphate.
Remember that beryllium ions in solution exist as the hydrated ion, [Be
(H2O)4]2+. The beryllium has such a strongly polarising effect on the
water molecules that hydrogen ions are very easily removed from them.
The sodium hydroxide solution contains hydroxide ions which are
powerful bases. If you add just the right amount of sodium hydroxide
solution, you get a precipitate of what is normally called "beryllium
hydroxide" - but which is a shade more complicated than that!

The product (other than water) is a neutral complex, and it is covalently


bonded. All that has happened to the original complex ion is that two
hydrogen ions have been removed from the water molecules.

http://www.chemguide.co.uk/inorganic/group2/beryllium.html (10 of 12)30/12/2004 11:36:54

Chemistry of beryllium untypical of Group 2

You get a precipitate of the neutral complex because of the lack of


charge on it. There isn't enough attraction between this neutral complex
and water molecules to bring it into solution.
What happens if you add an acid to this?
The hydrogen ions that were originally removed are simply replaced. The
precipitate dissolves as the original hydrated beryllium ion is re-formed.

What happens if you add a base?


Adding more hydroxide ions to the neutral complex pulls more hydrogen
ions off the water molecules to give the tetrahydroxoberyllate ion:

The beryllium hydroxide dissolves because the neutral complex is


converted into an ion which will be sufficiently attracted to water
molecules.
Why doesn't this happen with, for example, calcium hydroxide?
Calcium hydroxide is truly ionic - and contains simple hydroxide ions,
OH-. These react with hydrogen ions from an acid to form water - and so
the hydroxide reacts with acids.
However, there isn't any equivalent to the neutral complex. Adding more
hydroxide ions from a base has no effect because they haven't got
anything to react with.

Note: This has been simplified to bring it into line with the
sort of treatment you will meet for the acid-base behaviour of
transition metal hydroxides. In particular, the structure of
beryllium hydroxide is probably even more complicated than
has been suggested above!

http://www.chemguide.co.uk/inorganic/group2/beryllium.html (11 of 12)30/12/2004 11:36:54

Chemistry of beryllium untypical of Group 2

Where would you like to go now?


To the Group 2 menu . . .
To the Inorganic Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002

http://www.chemguide.co.uk/inorganic/group2/beryllium.html (12 of 12)30/12/2004 11:36:54

amines as bases

AMINES AS BASES

This page looks at the reactions of amines as bases. Their basic


properties include the reactions with dilute acids, water and copper(II)
ions.
It only deals with amines where the functional group is not attached
directly to a benzene ring. Aromatic amines such as phenylamine
(aniline) are much weaker bases than the amines discussed on this page
and are dealt with separately on a page specifically about phenylamine. If
you are interested in phenylamine, read this page first and then follow the
link at the bottom.

The basic properties of amines


We are going to have to use two different definitions of the term "base" in
this page.
A base is

a substance which combines with hydrogen ions. This is the


Bronsted-Lowry theory.
an electron pair donor. This is the Lewis theory.

Note: If you aren't familiar with either of these terms, you


should follow this link to a page about theories of acids and
bases.
Use the BACK button on your browser to return to this page
when you are confident about these terms.

http://www.chemguide.co.uk/organicprops/amines/base.html (1 of 8)30/12/2004 11:37:16

amines as bases

The easiest way of looking at the basic properties of amines is to think of


an amine as a modified ammonia molecule. In an amine, one or more of
the hydrogen atoms in ammonia has been replaced by a hydrocarbon
group.
Replacing the hydrogens still leaves the lone pair on the nitrogen
unchanged - and it is the lone pair on the nitrogen that gives ammonia its
basic properties. Amines will therefore behave much the same as
ammonia in all cases where the lone pair is involved.

The reactions of amines with acids


These are most easily considered using the Bronsted-Lowry theory of
acids and bases - the base is a hydrogen ion acceptor. We'll do a straight
comparison between amines and the familiar ammonia reactions.
A reminder about the ammonia reactions
Ammonia reacts with acids to produce ammonium ions. The ammonia
molecule picks up a hydrogen ion from the acid and attaches it to the
lone pair on the nitrogen.

If the reaction is in solution in water (using a dilute acid), the ammonia


takes a hydrogen ion (a proton) from a hydroxonium ion. (Remember that
hydrogen ions present in solutions of acids in water are carried on water
molecules as hydroxonium ions, H3O+.)

If the acid was hydrochloric acid, for example, you would end up with a
http://www.chemguide.co.uk/organicprops/amines/base.html (2 of 8)30/12/2004 11:37:16

amines as bases

solution containing ammonium chloride - the chloride ions, of course,


coming from the hydrochloric acid.
You could also write this last equation as:

. . . but if you do it this way, you must include the state symbols. If you
write H+ on its own, it implies an unattached hydrogen ion - a proton.
Such things don't exist on their own in solution in water.
If the reaction is happening in the gas state, the ammonia accepts a
proton directly from the hydrogen chloride:

This time you produce clouds of white solid ammonium chloride.


The corresponding reactions with amines
The nitrogen lone pair behaves exactly the same. The fact that one (or
more) of the hydrogens in the ammonia has been replaced by a
hydrocarbon group makes no difference.
For example, with ethylamine:
If the reaction is done in solution, the amine takes a hydrogen ion from a
hydroxonium ion and forms an ethylammonium ion.

Or:

The solution would contain ethylammonium chloride or sulphate or


whatever.
Alternatively, the amine will react with hydrogen chloride in the gas state
to produce the same sort of white smoke as ammonia did - but this time
of ethylammonium chloride.
http://www.chemguide.co.uk/organicprops/amines/base.html (3 of 8)30/12/2004 11:37:16

amines as bases

These examples have involved a primary amine. It would make no real


difference if you used a secondary or tertiary one. The equations would
just look more complicated.
The product ions from diethylamine and triethylamine would be
diethylammonium ions and triethylammonium ions respectively.

The reactions of amines with water


Again, it is easiest to use the Bronsted-Lowry theory and, again, it is
useful to do a straight comparison with ammonia.
A reminder about the ammonia reaction with water
Ammonia is a weak base and takes a hydrogen ion from a water
molecule to produce ammonium ions and hydroxide ions.
However, the ammonia is only a weak base, and doesn't hang on to the
hydrogen ion very successfully. The reaction is reversible, with the great
majority of the ammonia at any one time present as free ammonia rather
than ammonium ions.

The presence of the hydroxide ions from this reaction makes the solution
alkaline.
The corresponding reaction with amines
The amine still contains the nitrogen lone pair, and does exactly the
same thing.

http://www.chemguide.co.uk/organicprops/amines/base.html (4 of 8)30/12/2004 11:37:16

amines as bases

For example, with ethylamine, you get ethylammonium ions and


hydroxide ions produced.

There is, however, a difference in the position of equilibrium. Amines are


usually stronger bases than ammonia. (There are exceptions to this,
though - particularly if the amine group is attached directly to a benzene
ring.)

Note: If you want to explore some of the reasons for the


relative strengths of ammonia and the amines as bases you
could follow this link.
UK A level syllabuses are only concerned with the relative
strengths of ammonia and the primary amines, so that is all
you will find on that page.
If you choose to follow this link, use the BACK button on your
browser to return to this page.

The reactions of amines with copper(II) ions


Just like ammonia, amines react with copper(II) ions in two separate
stages. In the first step, we can go on using the Bronsted-Lowry theory
(that a base is a hydrogen ion acceptor). The second stage of the
reaction can only be explained in terms of the Lewis theory (that a base
is an electron pair donor).
The reaction between ammonia and copper(II) ions
Copper(II) sulphate solution, for example, contains the blue
hexaaquacopper(II) ion - [Cu(H2O)6]2+.
In the first stage of the reaction, the ammonia acts as a Bronsted-Lowry
base. With a small amount of ammonia solution, hydrogen ions are
pulled off two water molecules in the hexaaqua ion.

http://www.chemguide.co.uk/organicprops/amines/base.html (5 of 8)30/12/2004 11:37:16

amines as bases

This produces a neutral complex - one carrying no charge. If you remove


two positively charged hydrogen ions from a 2+ ion, then obviously there
isn't going to be any charge left on the ion.
Because of the lack of charge, the neutral complex isn't soluble in water,
and so you get a pale blue precipitate.

This precipitate is often written as Cu(OH)2 and called copper(II)


hydroxide. The reaction is reversible because ammonia is only a weak
base.
That precipitate dissolves if you add an excess of ammonia solution,
giving a deep blue solution.
The ammonia replaces four of the water molecules around the copper to
give tetraamminediaquacopper(II) ions. The ammonia uses its lone pair
to form a co-ordinate covalent bond (dative covalent bond) with the
copper. It is acting as an electron pair donor - a Lewis base.

Note: You might wonder why this second equation is given


starting from the original hexaaqua ion rather than the neutral
complex. Explaining why the precipitate redissolves is quite
complicated. You will find a lot more explanations about the
reactions between hexaaqua ions and ammonia solution in
the inorganic section of this site if you are interested.
(Important: The inorganic section describes ammonia acting
as a ligand in the second stage of the reaction. It is acting as
a ligand because it has a lone pair of electrons - in other
words, because it is a Lewis base.)
Use the BACK button (or HISTORY file or GO menu) on your
browser to return to this page.

http://www.chemguide.co.uk/organicprops/amines/base.html (6 of 8)30/12/2004 11:37:16

amines as bases

The colour changes are:

The corresponding reaction with amines


The small primary amines behave in exactly the same way as ammonia.
There will, however, be slight differences in the shades of blue that you
get during the reactions.
Taking methylamine as an example:
With a small amount of methylamine solution you will get a pale blue
precipitate of the same neutral complex as with ammonia. All that is
happening is that the methylamine is pulling hydrogen ions off the
attached water molecules.

With more methylamine solution the precipitate redissolves to give a


deep blue solution - just as in the ammonia case. The amine replaces
four of the water molecules around the copper.

As the amines get bigger and more bulky, the formula of the final product
may change - simply because it is impossible to fit four large amine
molecules and two water molecules around the copper atom.
http://www.chemguide.co.uk/organicprops/amines/base.html (7 of 8)30/12/2004 11:37:16

amines as bases

Where would you like to go now?


To the section on phenylamine
To the amines menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/amines/base.html (8 of 8)30/12/2004 11:37:16

phenylamine (aniline) as an amine

PHENYLAMINE AS A PRIMARY AMINE

This page looks at reactions of phenylamine (also known as aniline or


aminobenzene) where it behaves as a fairly straightforward primary amine.
It explains why phenylamine is a weaker base than other primary amines,
and summarises its reactions with acyl chlorides (acid chlorides), acid
anhydrides and halogenoalkanes (haloalkanes or alkyl halides).
Before you read each section on this page, you should follow the link to
the corresponding page about aliphatic amines (those not based on
benzene rings). In most cases, the reactions are the same, and this page
only really looks in detail at the differences in the phenylamine case.

Phenylamine as a base

Important: First read the page about the basic properties of


amines to give you all the detailed explanations about amines
in general.
Use the BACK button on your browser to return to this page
when you have done that.

http://www.chemguide.co.uk/organicprops/aniline/amine.html (1 of 10)30/12/2004 11:37:28

phenylamine (aniline) as an amine

Amines are bases because the lone pair of electrons on the nitrogen atom
can accept a hydrogen ion - in other words, for exactly the same reason
that ammonia is a base.
With phenylamine, the only difference is that it is a much weaker base than
ammonia or an amine like ethylamine - for reasons that we will explore
later.

The reaction of phenylamine with acids


Phenylamine reacts with acids like hydrochloric acid in exactly the same
way as any other amine. Despite the fact that the phenylamine is only a
very weak base, with a strong acid like hydrochloric acid the reaction is
completely straightforward.
Phenylamine is only very slightly soluble in water, but dissolves freely in
dilute hydrochloric acid. A solution of a salt is formed - phenylammonium
chloride.
If you just want to show the formation of the salt, you could write:

. . . or if you want to emphasise the fact that the phenylamine is acting as a


base, you could most simply use:

Note: I have deliberately written these equations without


stressing the benzene ring to emphasise the similarity between
the phenylamine reaction and that of other amines.

http://www.chemguide.co.uk/organicprops/aniline/amine.html (2 of 10)30/12/2004 11:37:28

phenylamine (aniline) as an amine

Getting the phenylamine back from its salt


To get the phenylamine back from the phenylammonium ion present in the
salt, all you have to do is to take the hydrogen ion away again. You can do
that by adding any stronger base.
Normally, you would choose sodium hydroxide solution.

The phenylamine is formed first as an off-white emulsion - tiny droplets of


phenylamine scattered throughout the water. This then settles out to give
an oily bottom layer of phenylamine under the aqueous layer.

The reaction of phenylamine with water


This is where it is possible to tell that phenylamine is a much weaker base
than ammonia and the aliphatic amines like methylamine and ethylamine.
Phenylamine reacts reversibly with water to give phenylammonium ions
and hydroxide ions.

The position of equilibrium lies well to the left of the corresponding


ammonia or aliphatic amine equilibria - which means that not many
hydroxide ions are formed in the solution.
The effect of this is that the pH of a solution of phenylamine will be quite a
bit lower than a solution of ammonia or one of the aliphatic amines of the
same concentration. For example, a 0.1 M phenylamine solution has a pH
of about 9 compared to a pH of about 11 for 0.1 M ammonia solution.
Why is phenylamine such a weak base?
Amines are bases because they pick up hydrogen ions on the lone pair on
the nitrogen atom. In phenylamine, the attractiveness of the lone pair is
lessened because of the way it interacts with the ring electrons.

http://www.chemguide.co.uk/organicprops/aniline/amine.html (3 of 10)30/12/2004 11:37:28

phenylamine (aniline) as an amine

The lone pair on the nitrogen touches the delocalised ring electrons . . .

. . . and becomes delocalised with them:

Note: If you don't understand these diagrams, you will need to


read about bonding in benzene in order to make sense of them.
If you follow this link, you may have to explore several other
pages before you are ready to come back here again. Use the
BACK button (or HISTORY file or GO menu) on your browser
to return to this page.

http://www.chemguide.co.uk/organicprops/aniline/amine.html (4 of 10)30/12/2004 11:37:28

phenylamine (aniline) as an amine

That means that the lone pair is no longer fully available to combine with
hydrogen ions. The nitrogen is still the most electronegative atom in the
molecule, and so the delocalised electrons will be attracted towards it, but
the electron density around the nitrogen is nothing like it is in, say, an
ammonia molecule.
The other problem is that if the lone pair is used to join to a hydrogen ion, it
is no longer available to contribute to the delocalisation. That means that
the delocalisation would have to be disrupted if the phenylamine acts as a
base. Delocalisation makes molecules more stable, and so disrupting the
delocalisation costs energy and won't happen easily.
Taken together - the lack of intense charge around the nitrogen, and the
need to break some delocalisation - means that phenylamine is a very
weak base indeed.

The acylation of phenylamine


The reactions with acyl chlorides and with acid anhydrides
These are reactions in which the phenylamine acts as a nucleophile. There
is no essential difference between these reactions and the same reactions
involving any other primary amine. You will find a summary of the
reactions below, but all the detailed explanations are on other pages.

Important: For detailed explanations, read the pages about


the reactions of acyl chlorides and reactions of acid anhydrides
with nitrogen compounds. It might pay you to read the acid
anhydride page first because this also summarises the acyl
chloride reactions. It could save you having to read the acyl
chloride page at all.
If you want the mechanism for the reaction with an acyl
chloride, you will find a link from the acyl chloride page.
Mechanisms for acid anhydrides aren't required by any UK A
level syllabus, so aren't covered anywhere on the site.
Take your time over this - it isn't easy stuff! It might be a good

http://www.chemguide.co.uk/organicprops/aniline/amine.html (5 of 10)30/12/2004 11:37:28

phenylamine (aniline) as an amine

idea to check your syllabus before you get too bogged down.
Follow this link if you are a UK A level student and haven't got
a copy of your syllabus.
Use the BACK button (or HISTORY file or GO menu) on your
browser to return to this page when you have done all this.

We'll take ethanoyl chloride as a typical acyl chloride, and ethanoic


anhydride as a typical acid anhydride. The important product of the
reaction of phenylamine with either of these is the same.
Phenylamine reacts vigorously in the cold with ethanoyl chloride to give a
mixture of solid products - ideally white, but usually stained brownish. A
mixture of N-phenylethanamide (old name: acetanilide) and
phenylammonium chloride is formed.
The overall equation for the reaction is:

With ethanoic anhydride, heat is needed. In this case, the products are a
mixture of N-phenylethanamide and phenylammonium ethanoate.

Note: Don't be confused by the fact that the two salts


(phenylammonium chloride and phenylammonium ethanoate)
are written differently. Ethanoates are always written with the
ethanoic acid bit first. That means having to reverse the way
you write the phenylammonium ion to keep the charges close
together in the formula.

http://www.chemguide.co.uk/organicprops/aniline/amine.html (6 of 10)30/12/2004 11:37:28

phenylamine (aniline) as an amine

The main product molecule (the N-phenylethanamide) is often drawn


looking like this:

If you stop and think about it, this is obviously the same molecule as in the
equation above, but it stresses the phenylamine part of it much more.
Looking at it this way, notice that one of the hydrogens of the -NH2 group
has been replaced by an acyl group - an alkyl group attached to a carbonoxygen double bond.
You can say that the phenylamine has been acylated or has undergone
acylation.
Because of the nature of this particular acyl group, it is also described as
ethanoylation. The hydrogen is being replaced by an ethanoyl group,
CH3CO-.

Going beyond A level: This is a useful reaction in the case of


phenylamine. The -NH2 group in phenylamine has major
effects on the properties of the benzene ring. It makes it much
more reactive than in benzene itself, and also makes it prone
to oxidation.
For example, if you wanted to nitrate phenylamine, you would
get several nitro groups substituted around the ring and a lot of
unwanted oxidation because nitric acid is a strong oxidising
agent.
To just substitute one nitro group and stop the oxidation, you
first ethanoylate the -NH2 group which reduces its effect on the
ring considerably. After nitration, you can remove the ethanoyl
group by simply warming with sodium hydroxide solution. The
N-substituted amide is hydrolysed just like any other amide
http://www.chemguide.co.uk/organicprops/aniline/amine.html (7 of 10)30/12/2004 11:37:28

phenylamine (aniline) as an amine

under alkaline conditions.


If you are interested, the alkaline hydrolysis of amides is
covered on another page.

The reaction of phenylamine with halogenoalkanes


This is another reaction of phenylamine as a nucleophile, and again there
is no essential difference between its reactions and those of aliphatic
amines.

Important: You should read the first half of the page about
amines as nucleophiles to give you explanations about amines
in general - and possibly also the further link on that page back
to reactions of halogenoalkanes for even more detail.
Again, take your time over this. It isn't easy, particularly if you
have to go beyond the formation of the secondary amine.
Use the BACK button on your browser to return to this page
when you have done that.

http://www.chemguide.co.uk/organicprops/aniline/amine.html (8 of 10)30/12/2004 11:37:28

phenylamine (aniline) as an amine

Taking bromoethane as a typical halogenoalkane, the reaction with


phenylamine happens in the same series of complicated steps as with any
other amine.
We'll just look at the first step.
On heating, the bromoethane and phenylamine react to give a mixture of a
salt of a secondary amine and some free secondary amine. In this case,
you would first get N-ethylphenylammonium bromide:

. . . but this would instantly be followed by a reversible reaction in which


some unreacted phenylamine would take a hydrogen ion from the salt to
give some free secondary amine: N-ethylphenylamine.

The reaction wouldn't stop there. You will get further reactions to produce
a tertiary amine and its salt, and eventually a quaternary ammonium
compound. If you want to explore this further, refer to the last link just up
the page, and trace the sequence of equations through using phenylamine
rather than ethylamine.

Note: You are so unlikely to be asked to do this for UK A level


purposes that I am not going to go through it. All it would do is
make chemistry look scary! If you want to trace the whole
reaction through, make sure that you understand what is going
on at each stage by following up the leads I have given you to
other pages on the site - and then work it out for yourself. If you
can do it - well done! If you can't, don't worry about it - life's too
short!

http://www.chemguide.co.uk/organicprops/aniline/amine.html (9 of 10)30/12/2004 11:37:28

phenylamine (aniline) as an amine

Where would you like to go now?


To the phenylamine menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/aniline/amine.html (10 of 10)30/12/2004 11:37:28

the hydrolysis of amides

HYDROLYSING AMIDES

This page describes the hydrolysis of amides under both acidic and
alkaline conditions. It also describes the use of alkaline hydrolysis in
testing for amides.

The hydrolysis of amides


What is hydrolysis?
Technically, hydrolysis is a reaction with water. That is exactly what
happens when amides are hydrolysed in the presence of dilute acids
such as dilute hydrochloric acid. The acid acts as a catalyst for the
reaction between the amide and water.
The alkaline hydrolysis of amides actually involves reaction with
hydroxide ions, but the result is similar enough that it is still classed as
hydrolysis.

Hydrolysis under acidic conditions


Taking ethanamide as a typical amide:
If ethanamide is heated with a dilute acid (such as dilute hydrochloric
acid), ethanoic acid is formed together with ammonium ions. So, if you
were using hydrochloric acid, the final solution would contain ammonium
chloride and ethanoic acid.

http://www.chemguide.co.uk/organicprops/amides/hydrolysis.html (1 of 3)30/12/2004 11:37:33

the hydrolysis of amides

Note: You might argue that because the hydrochloric acid is


changed during the reaction, it isn't acting as a catalyst. In
fact, it is doing two things. It is acting as a catalyst in a
reaction between the amide and water which would produce
ammonium ethanoate (containing ammonium ions and
ethanoate ions). It is secondly reacting with those ethanoate
ions to make ethanoic acid.

Hydrolysis under alkaline conditions


Again, taking ethanamide as a typical amide:
If ethanamide is heated with sodium hydroxide solution, ammonia gas is
given off and you are left with a solution containing sodium ethanoate.

Using alkaline hydrolysis to test for an amide


If you add sodium hydroxide solution to an unknown organic compound,
and it gives off ammonia on heating (but not immediately in the cold),
then it is an amide.
You can recognise the ammonia by smell and because it turns red litmus
paper blue.
The possible confusion using this test is with ammonium salts.
Ammonium salts also produce ammonia with sodium hydroxide solution,
but in this case there is always enough ammonia produced in the cold for
the smell to be immediately obvious.

http://www.chemguide.co.uk/organicprops/amides/hydrolysis.html (2 of 3)30/12/2004 11:37:33

the hydrolysis of amides

Note: This test is OK for UK A level purposes, but there are


other things which also give off ammonia on heating with
sodium hydroxide solution - for example, nitriles (but you
won't come across them in a practical situation at this level)
and imides (but they are beyond the scope of courses at this
level).

Where would you like to go now?


To the amides menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/amides/hydrolysis.html (3 of 3)30/12/2004 11:37:33

Amides Menu

Understanding Chemistry

AMIDES MENU

Background . . .
An introduction to amides including their physical properties.
Preparation of amides . . .
Their preparation from carboxylic acids, acyl chlorides and acid
anhydrides.
The hydrolysis of amides . . .
The hydrolysis of amides using acids or alkalis (including the test
for an amide).
Other reactions of amides . . .
The lack of base character in amides, the dehydration of amides to
make nitriles, and the Hofmann degradation of amides to make
primary amines with one less carbon atom.
Polyamides . . .
A summary of the chemistry of simple polyamides like nylon and
Kevlar, including their formation, structure, hydrolysis and uses.

Go to menu of other organic compounds . . .

http://www.chemguide.co.uk/organicprops/amidemenu.html (1 of 2)30/12/2004 11:37:34

Amides Menu

Go to Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/amidemenu.html (2 of 2)30/12/2004 11:37:34

an introduction to amides

INTRODUCING AMIDES

This page explains what amides are and looks at their simple physical
properties such as solubility and melting points.

What are amides?


Amides are derived from carboxylic acids. A carboxylic acid contains the COOH group, and in an amide the -OH part of that group is replaced by
an -NH2 group.
So . . . amides contain the -CONH2 group.
Some simple amides
The most commonly discussed amide is ethanamide, CH3CONH2 (old
name: acetamide).

The three simplest amides are:

HCONH2

methanamide

CH3CONH2

ethanamide

CH3CH2CONH2

propanamide

Notice that in each case, the name is derived from the acid by replacing

http://www.chemguide.co.uk/organicprops/amides/background.html (1 of 4)30/12/2004 11:37:37

an introduction to amides

the "oic acid" ending by "amide".


If the chain was branched, the carbon in the -CONH2 group counts as
the number 1 carbon atom. For example:

Physical properties
Melting points
Methanamide is a liquid at room temperature (melting point: 3C), but the
other amides are solid.
For example, ethanamide forms colourless deliquescent crystals with a
melting point of 82C. A deliquescent substance is one which picks up
water from the atmosphere and dissolves in it. Ethanamide crystals
nearly always look wet.

Note: Ethanamide is said to smell of mice. In fact, the smell


is due to an impurity in the ethanamide called Nmethylethanamide, CH3CONHCH3, where one of the
hydrogens in the -NH2 group has been replaced by a methyl
group.

http://www.chemguide.co.uk/organicprops/amides/background.html (2 of 4)30/12/2004 11:37:37

an introduction to amides

The melting points of the amides are high for the size of the molecules
because they can form hydrogen bonds. The hydrogen atoms in the NH2 group are sufficiently positive to form a hydrogen bond with a lone
pair on the oxygen atom of another molecule.

As you can see, there is the potential for lots of hydrogen bonds to be
formed. Each molecule has two slightly positive hydrogen atoms and two
lone pairs on the oxygen atom.
These hydrogen bonds need a reasonable amount of energy to break,
and so the melting points of the amides are quite high.

Note: If you aren't happy about hydrogen bonding then you


really ought to follow this link before you go on.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/amides/background.html (3 of 4)30/12/2004 11:37:37

an introduction to amides

Solubility in water
The small amides are soluble in water because they have the ability to
hydrogen bond with the water molecules.
It needs energy to break the hydrogen bonds between amide molecules
and between water molecules before they can mix - but enough energy
is released again when the new hydrogen bonds are set up to allow this
to happen.

Where would you like to go now?


To the amides menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/amides/background.html (4 of 4)30/12/2004 11:37:37

the preparation of amides

MAKING AMIDES

This page describes the preparation of amides from carboxylic acids, acyl
chlorides (acid chlorides) and acid anhydrides.

Making amides from carboxylic acids


Summary of the process
The carboxylic acid is first converted into an ammonium salt which then
produces an amide on heating.
The ammonium salt is formed by adding solid ammonium carbonate to an
excess of the acid.
For example, ammonium ethanoate is made by adding ammonium
carbonate to an excess of ethanoic acid.

When the reaction is complete, the mixture is heated and the ammonium
salt dehydrates producing ethanamide.

The excess of ethanoic acid is there to prevent dissociation of the


ammonium salt before it dehydrates.
Ammonium salts tend to split into ammonia and the parent acid on heating,
recombining on cooling. If dissociation happened in this case, the
ammonia would escape from the reaction mixture and be lost. You couldn't
get any recombination.
The dissociation is reversible:

http://www.chemguide.co.uk/organicprops/amides/preparation.html (1 of 5)30/12/2004 11:37:47

the preparation of amides

The presence of the excess ethanoic acid helps to prevent this from
happening by moving the position of equilibrium to the left.

Note: To understand why this is, you need to know about Le


Chatelier's Principle and about the effect of changes of
concentration on the position of equilibrium.
If you choose to follow this link, use the BACK button on your
browser to return to this page.

Some brief practical details


The ammonium carbonate is added slowly to concentrated ethanoic acid
and the reaction is left until all production of carbon dioxide stops.
It is then heated under reflux for half an hour for the dehydration to take
place.
The mixture is distilled at about 170C to remove excess ethanoic acid and
water - leaving almost pure ethanamide in the flask.
Further purification stages are beyond the scope of this site.

Making amides from acyl chlorides


Acyl chlorides (also known as acid chlorides) have the general formula
RCOCl. The chlorine atom is very easily replaced by other things. For
example, it is easily replaced by an -NH2 group to make an amide.
To make ethanamide from ethanoyl chloride, you normally add the
ethanoyl chloride to a concentrated solution of ammonia in water. There is
a very violent reaction producing lots of white smoke - a mixture of solid
ammonium chloride and ethanamide. Some of the mixture remains
dissolved in water as a colourless solution.

http://www.chemguide.co.uk/organicprops/amides/preparation.html (2 of 5)30/12/2004 11:37:47

the preparation of amides

You can think of the reaction as happening in two stages.


In the first stage, the ammonia reacts with the ethanoyl chloride to give
ethanamide and hydrogen chloride gas.

Then the hydrogen chloride produced reacts with excess ammonia to give
ammonium chloride.

. . . and you can combine all this together to give one overall equation:

Note: If you are want the mechanism for this reaction, you will
find it by following this link to another part of the site dealing
with nucleophilic addition-elimination reactions.
You can find more about the reactions of acyl chlorides with
nitrogen compounds by following this link. This will also give
you information about the preparation of N-substituted amides
by the reaction between acyl chlorides and primary amines.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/amides/preparation.html (3 of 5)30/12/2004 11:37:47

the preparation of amides

Making amides from acid anhydrides


An acid anhydride is what you get if you remove a molecule of water from
two carboxylic acid -COOH groups.
For example, if you took two ethanoic acid molecules and removed a
molecule of water between them you would get the acid anhydride,
ethanoic anhydride (old name: acetic anhydride).

For equation purposes, ethanoic anhydride is often written as (CH3CO)2O.


The reactions of acid anhydrides are rather like those of acyl chlorides
except that during their reactions, a molecule of carboxylic acid is
produced rather than the HCl formed when an acyl chloride reacts.
If ethanoic anhydride is added to concentrated ammonia solution,
ethanamide is formed together with ammonium ethanoate. Again, the
reaction happens in two stages.
In the first stage, ethanamide is formed together with ethanoic acid.

Then the ethanoic acid produced reacts with excess ammonia to give
ammonium ethanoate.

. . . and you can combine all this together to give one overall equation:

http://www.chemguide.co.uk/organicprops/amides/preparation.html (4 of 5)30/12/2004 11:37:47

the preparation of amides

You need to follow this through really carefully, because the two products
of the reaction overall can look confusingly similar.

Note: You can find more about the reactions of acid


anhydrides with nitrogen compounds by following this link. This
will also give you information about the preparation of Nsubstituted amides by the reaction between acid anhydrides
and primary amines.
Use the BACK button on your browser to return to this page.

Where would you like to go now?


To the amides menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/amides/preparation.html (5 of 5)30/12/2004 11:37:47

other reactions of amides

OTHER REACTIONS OF AMIDES

This page explains the reason for the lack of basic character in amides,
and describes their dehydration to give nitriles, and their reaction with
bromine and sodium hydroxide solution to form primary amines with one
less carbon atom (the Hofmann degradation).

Note: The hydrolysis of amides is described on a separate


page.
If you choose to follow this link, use the BACK button on your
browser to return to this page.

The lack of base character in amides


Unusually for compounds containing the -NH2 group, amides are neutral.
This section explains why -NH2 groups are usually basic and why amides
are different.
The usual basic character of the -NH2 group
Simple compounds containing an -NH2 group such as ammonia, NH3, or
a primary amine like methylamine, CH3NH2, are weak bases. A primary
amine is a compound where the -NH2 group is attached to a hydrocarbon
group.
The active lone pair of electrons on the nitrogen atom in ammonia can
combine with a hydrogen ion (a proton) from some other source - in other
words it acts as a base.
With a compound like methylamine, all that has happened is that one of
the hydrogen atoms attached to the nitrogen has been replaced by a
methyl group. It doesn't make a huge amount of difference to the lone
http://www.chemguide.co.uk/organicprops/amides/other.html (1 of 7)30/12/2004 11:37:52

other reactions of amides

pair and so ammonia and methylamine behave similarly.

Note: The reasons that these are bases and the differences
between them (because there are slight differences) are
explored in some detail on a page about organic bases. It
would be useful to read this page before you go on because
it is relevant to what is coming next.
If you follow this link, use the BACK button on your browser
to return to this page.

For example, if you dissolve these compounds in water, the nitrogen lone
pair takes a hydrogen ion from a water molecule - and equilibria like
these are set up:

Notice that the reactions are reversible. In both cases the positions of
equilibrium lie well to the left. These compounds are weak bases
because they don't hang on to the incoming hydrogen ion very well.

http://www.chemguide.co.uk/organicprops/amides/other.html (2 of 7)30/12/2004 11:37:52

other reactions of amides

Both ammonia and the amines are alkaline in solution because of the
presence of the hydroxide ions, and both of them turn red litmus blue.

Why doesn't something similar happen with amides?


Amides are neutral to litmus and have virtually no basic character at all despite having the -NH2 group. Their tendency to attract hydrogen ions is
so slight that it can be ignored for most purposes.

Note: If you haven't already done so, follow the link


mentioned above to the page about organic bases, and read
the bit about phenylamine. It is directly relevant to what's next.
Use the BACK button on your browser to return to this page.

We need to look at the bonding in the -CONH2 group.


Like any other double bond, a carbon-oxygen double bond is made up of
two different parts. One electron pair is found on the line between the two
nuclei - this is known as a sigma bond. The other electron pair is found
above and below the plane of the molecule in a pi bond.
A pi bond is made by sideways overlap between p orbitals on the carbon
and the oxygen.
In an amide, the lone pair on the nitrogen atom ends up almost parallel to
these p orbitals, and overlaps with them as they form the pi bond.

http://www.chemguide.co.uk/organicprops/amides/other.html (3 of 7)30/12/2004 11:37:52

other reactions of amides

The result of this is that the nitrogen lone pair becomes delocalised - in
other words it is no longer found located on the nitrogen atom, but the
electrons from it are spread out over the whole of that part of the
molecule.
This has two effects which prevent the lone pair accepting hydrogen ions
and acting as a base:

Because the lone pair is no longer located on a single atom as an


intensely negative region of space, it isn't anything like as
attractive for a nearby hydrogen ion.
Delocalisation makes molecules more stable. For the nitrogen to
reclaim its lone pair and join to a hydrogen ion, the delocalisation
would have to be broken, and that will cost energy.

http://www.chemguide.co.uk/organicprops/amides/other.html (4 of 7)30/12/2004 11:37:52

other reactions of amides

Note: If you want to look in more detail at the bonding in the


carbon-oxygen double bond, you could follow this link.
If you do choose to follow this link, it will probably take you to
several other pages before you are ready to come back here
again. Use the BACK button (or HISTORY file or GO menu)
on your browser to return to this page later.

The dehydration of amides


Amides are dehydrated by heating a solid mixture of the amide and
phosphorus(V) oxide, P4O10.
Water is removed from the amide group to leave a nitrile group, -CN. The
liquid nitrile is collected by simple distillation.
For example, with ethanamide, you will get ethanenitrile.

Note: This is a just a flow scheme rather than a proper


equation. I haven't been able to find a single example of the
use of the full equation for this reaction. In fact the
phosphorus(V) oxide reacts with the water to produce
mixtures of phosphorus-containing acids.

http://www.chemguide.co.uk/organicprops/amides/other.html (5 of 7)30/12/2004 11:37:52

other reactions of amides

The Hofmann Degradation


The Hofmann degradation is a reaction between an amide and a mixture
of bromine and sodium hydroxide solution. Heat is needed.
The net effect of the reaction is a loss of the -CO- part of the amide
group. You get a primary amine with one less carbon atom than the
original amide had.
The general case would be (as a flow scheme):

If you started with ethanamide, you would get methylamine. The full
equation for the reaction is:

The Hofmann degradation is used as a way of cutting a single carbon


atom out of a chain.

Where would you like to go now?


To the amides menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

http://www.chemguide.co.uk/organicprops/amides/other.html (6 of 7)30/12/2004 11:37:52

other reactions of amides

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/amides/other.html (7 of 7)30/12/2004 11:37:52

polyamides - nylon and Kevlar

POLYAMIDES

This page looks at the structures, formation, hydrolysis and uses of the
polyamides, nylon and Kevlar.

What are polyamides?


Polyamides are polymers where the repeating units are held together by
amide links.
An amide group has the formula - CONH2. An amide link has this
structure:

In an amide itself, of course, the bond on the right is attached to a


hydrogen atom.

Note: If you know any biology or biochemistry, you may


have come across this called a peptide group. If you are
interested in the presence of this group in proteins, you could
follow this link.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/amides/polyamides.html (1 of 8)30/12/2004 11:38:03

polyamides - nylon and Kevlar

Nylon
In nylon, the repeating units contain chains of carbon atoms. (That is
different from Kevlar, where the repeating units contain benzene rings see below.) There are various different types of nylon depending on the
nature of those chains.
Nylon-6,6
Nylon-6,6 is made from two monomers each of which contain 6 carbon
atoms - hence its name.
One of the monomers is a 6 carbon acid with a -COOH group at each
end - hexanedioic acid.

Note: When you count the carbons don't forget to include


the ones in the -COOH groups.

The other monomer is a 6 carbon chain with an amino group, -NH2, at


each end. This is 1,6-diaminohexane (also known as hexane-1,6diamine).

When these two compounds polymerise, the amine and acid groups
combine, each time with the loss of a molecule of water. This is known
as condensation polymerisation.
Condensation polymerisation is the formation of a polymer involving the
loss of a small molecule. In this case, the molecule is water, but in other
cases different small molecules might be lost.
The diagram shows the loss of water between two of the monomers:

http://www.chemguide.co.uk/organicprops/amides/polyamides.html (2 of 8)30/12/2004 11:38:03

polyamides - nylon and Kevlar

This keeps on happening, and so you get a chain which looks like this:

Note: This chain is much easier to work out than to


remember. Learn the structures of the monomers, and then
practice writing them down and removing water from them as
shown above.

Nylon-6
If you are doing UK A level, you are unlikely to need the structure of
nylon-6. I am including it to show that it is possible to get a polyamide
from a single monomer.
Nylon-6 is made from a monomer called caprolactam.

Notice that this already contains an amide link. When this molecule
polymerises, the ring opens, and the molecules join up in a continuous
chain.

http://www.chemguide.co.uk/organicprops/amides/polyamides.html (3 of 8)30/12/2004 11:38:03

polyamides - nylon and Kevlar

Kevlar
Kevlar is similar in structure to nylon-6,6 except that instead of the amide
links joining chains of carbon atoms together, they join benzene rings.
The two monomers are benzene-1,4-dicarboxylic acid and 1,4diaminobenzene.

If you line these up and remove water between the -COOH and -NH2
groups in the same way as we did with nylon-6,6, you get the structure of
Kevlar:

Making nylon-6,6
Making nylon-6,6 industrially
Nylon-6,6 is made by polymerising hexanedioic acid and 1,6diaminohexane exactly as shown further up the page.
Because the acid is acidic and the amine is basic, they first react
together to form a salt. That is then converted into nylon-6,6 by heating it
under pressure at 350C.

http://www.chemguide.co.uk/organicprops/amides/polyamides.html (4 of 8)30/12/2004 11:38:03

polyamides - nylon and Kevlar

The two monomers can both be made from cyclohexane.

Oxidation of the cyclohexane opens the ring of carbon atoms and


produces a -COOH group at each end. That gives you the
hexanedioic acid.
Some of that can then be converted into the 1,6-diaminohexane.

The acid is treated with ammonia to produce the ammonium salt.

The ammonium salt is heated to 350C in the presence of


hydrogen and a nickel catalyst. This both dehydrates the salt and
reduces it to the 1,6-diaminohexane.

Making nylon-6,6 in the lab


In the lab, it is easy to make nylon-6,6 at room temperature using an acyl
chloride (acid chloride) rather than an acid.
The 1,6-diaminohexane is used just as before, but hexanedioyl dichloride
is used instead of hexanedioic acid.

If you compare the next diagram with the diagram further up the page for
the formation of nylon-6,6, you will see that the only difference is that
molecules of HCl are lost rather than molecules of water.

http://www.chemguide.co.uk/organicprops/amides/polyamides.html (5 of 8)30/12/2004 11:38:03

polyamides - nylon and Kevlar

Note: If you are interested, the reaction between acyl


chlorides and amines is explored on another page.
If you follow this link, use the BACK button on your browser
to return to this page.

In the lab, this reaction is the basis for the nylon rope trick.
You make a solution of the hexanedioyl dichloride in an organic solvent,
and a solution of 1,6-diaminohexane in water. You carefully float one
solution on top of the other in a small beaker, taking care to get as little
mixing as possible.
Nylon-6,6 forms at the boundary between the two solutions. If you pick
up the boundary layer with a pair of tweezers, you can pull out an
amazingly long tube of nylon from the beaker.

http://www.chemguide.co.uk/organicprops/amides/polyamides.html (6 of 8)30/12/2004 11:38:03

polyamides - nylon and Kevlar

Note: The exact details of this experiment depend on which


organic solvent you use for the hexanedioyl dichloride. This
will affect which layer is at the top and which layer at the
bottom, depending on the densities of the two solutions. For
specific details, you should find this experiment in most
organic practical books at this level, or a Google search for
"nylon rope trick" will throw up lots of variations.
You will find a Google search box at the bottom of the Main
Menu. Don't forget to set it to search the whole web and not
just this site.

Hydrolysis of polyamides
Simple amides are easily hydrolysed by reaction with dilute acids or
alkalis.
Polyamides are fairly readily attacked by strong acids, but are much
more resistant to alkaline hydrolysis. Hydrolysis is faster at higher
temperatures. Hydrolysis by water alone is so slow as to be completely
unimportant. Kevlar is rather more resistant to hydrolysis than nylon is.
If you spill something like dilute sulphuric acid on a fabric made from
nylon, the amide linkages are broken. The long chains break and you
can eventually end up with the original monomers - hexanedioic acid and
1,6-diaminohexane.
Because you produce small molecules rather than the original polymer,
the fibres are destroyed, and you end up with a hole!

Note: Hydrolysis of amides is covered in detail on another


page in this section.

http://www.chemguide.co.uk/organicprops/amides/polyamides.html (7 of 8)30/12/2004 11:38:03

polyamides - nylon and Kevlar

Uses of polyamides
Nylon
Apart from obvious uses in textiles for clothing and carpets, a lot of nylon
is used to make tyre cords - the inner structure of a vehicle tyre
underneath the rubber.
The fibres are also used in ropes, and nylon can be cast into solid
shapes for cogs and bearings in machines, for example.
Kevlar
Kevlar is a very strong material - about five times as strong as steel,
weight for weight. It is used in bulletproof vests, in composites for boat
construction, in lightweight mountaineering ropes, and for lightweight skis
and racquets - amongst many other things.

Where would you like to go now?


To the amides menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/amides/polyamides.html (8 of 8)30/12/2004 11:38:03

amines as nucleophiles

AMINES AS NUCLEOPHILES

This page summarises the reactions of amines as nucleophiles. This


includes their reactions with halogenoalkanes (haloalkanes or alkyl
halides), with acyl chlorides (acid chlorides) and with acid anhydrides.
All of these reactions are dealt with in detail on other pages and you will
find links to all of these.

Note: If you are mainly interested in phenylamine (aniline),


its reactions are summarised in another section. You will find
a link at the bottom of this page. It might be a good idea to
read the current page first, though.

The nucleophilic properties of amines


Why do amines act as nucleophiles?
A nucleophile is something which is attracted to, and then attacks, a
positive or slightly positive part of another molecule or ion.
All amines contain an active lone pair of electrons on the very
electronegative nitrogen atom. It is these electrons which are attracted to
positive parts of other molecules or ions.

The reactions of primary amines with halogenoalkanes


You get a complicated series of reactions on heating to give a mixture of
products - probably one of the most confusing sets of reactions you will
meet at this level. The products of the reactions include secondary and
tertiary amines and their salts, and quaternary ammonium salts.
Making secondary amines and their salts

http://www.chemguide.co.uk/organicprops/amines/nucleophile.html (1 of 5)30/12/2004 11:38:12

amines as nucleophiles

In the first stage of the reaction, you get the salt of a secondary amine
formed. For example if you started with ethylamine and bromoethane,
you would get diethylammonium bromide

In the presence of excess ethylamine in the mixture, there is the


possibility of a reversible reaction. The ethylamine removes a hydrogen
from the diethylammonium ion to give free diethylamine - a secondary
amine.

Making tertiary amines and their salts


But it doesn't stop here! The diethylamine also reacts with bromoethane in the same two stages as before. This is where the reaction would start
if you reacted a secondary amine with a halogenoalkane.
In the first stage, you get triethylammonium bromide.

There is again the possibility of a reversible reaction between this salt


and excess ethylamine in the mixture.

http://www.chemguide.co.uk/organicprops/amines/nucleophile.html (2 of 5)30/12/2004 11:38:12

amines as nucleophiles

The ethylamine removes a hydrogen ion from the triethylammonium ion


to leave a tertiary amine - triethylamine.
Making a quaternary ammonium salt
The final stage! The triethylamine reacts with bromoethane to give
tetraethylammonium bromide - a quaternary ammonium salt (one in
which all four hydrogens have been replaced by alkyl groups).

This time there isn't any hydrogen left on the nitrogen to be removed. The
reaction stops here.

Note: You will find this reaction explored in the page about
the reactions between halogenoalkanes and ammonia although in that case starting from bromoethane and
ammonia. It would be useful to read that in order to compare
the ammonia reactions with the amine reactions. The only
real difference is that the hydrogen ions are removed by an
amine in the case we are currently looking at rather than
ammonia on the other page.
You will find a further link to the mechanisms for these
reactions on that page. If you follow that link it will take you to
quite a number of pages exploring these reactions. You will
need to spend some time on these if you really want to
understand this. However, understanding the reactions in
terms of their mechanisms means that you can work out what
happens if you need to rather than trying to remember all this
(a fairly pointless and soul-destroying exercise!).
Use the BACK button (or HISTORY file or GO menu) on your
browser to return to this page.

http://www.chemguide.co.uk/organicprops/amines/nucleophile.html (3 of 5)30/12/2004 11:38:12

amines as nucleophiles

The reactions of amines with acyl chlorides (acid chlorides)


We'll take the reaction between methylamine and ethanoyl chloride as
typical.
If you add concentrated methylamine solution to ethanoyl chloride, there
is a violent reaction in the cold. N-methylethanamide and
methylammonium chloride are formed - partly as a white solid mixture,
and partly in solution.
The overall equation is:

Note: This reaction (and the corresponding one with


ammonia) is discussed in detail on a page about the
reactions between acyl chlorides and nitrogen compounds.
You will also find a link to the introduction to the mechanisms
for these reactions on that page. If you want to go straight to
the mechanism for the amine reaction, you could follow this
link.
Use the BACK button on your browser to return to this page.

The reactions of amines with acid anhydrides


These reactions are chemically similar to those between amines and acyl
chlorides, but they are much slower, needing heat.
Taking the reaction between methylamine and ethanoic anhydride as
typical:
The product is N-methylethanamide (as with ethanoyl chloride), but this
time the other product is methylammonium ethanoate rather than
methylammonium chloride.

http://www.chemguide.co.uk/organicprops/amines/nucleophile.html (4 of 5)30/12/2004 11:38:12

amines as nucleophiles

Note: I have shown the ionic nature of the methylammonium


ethanoate this time because it isn't otherwise obvious how
everything joins up. Methylammonium chloride is, of course,
also ionic. I have also reversed the formula for the
methylammonium ion so that the positive charge is close to
the negative charge on the ethanoate ion.
This reaction (and the corresponding one with ammonia) is
discussed in detail on a page about the reactions between
acid anhydrides and nitrogen compounds. This is quite
difficult stuff - it would pay you to read the whole of that page
fairly carefully if you really want to understand what's going
on.
Use the BACK button on your browser to return to this page.

Where would you like to go now?


To the section on phenylamine
To the amines menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/amines/nucleophile.html (5 of 5)30/12/2004 11:38:12

amines and nitrous acid

AMINES AND NITROUS ACID

This page looks at the reactions of nitrous acid with aliphatic amines
(those where the amine group isn't attached directly to a benzene ring).
Nitrous acid is properly called nitric(III) acid, but that name isn't
commonly used.
Reactions between nitrous acid and aromatic amines like phenylamine
(where the -NH2 group is attached directly to a benzene ring) are dealt
with elsewhere.

Note: Unless you want to compare the reactions of aliphatic


and aromatic amines with nitrous acid, if you are interested
mainly in the reactions of phenylamine with nitrous acid, it
would pay you to go straight to this page.

Testing for the various types of amines


Background
The reaction between amines and nitrous acid was used in the past as a
very neat way of distinguishing between primary, secondary and tertiary
amines. However, the product with a secondary amine is a powerful
carcinogen, and so this reaction is no longer carried out at this level.
Nitrous acid, HNO2, (sometimes written as HONO to show its structure)
is unstable and is always prepared in situ.
It is usually made by reacting a solution containing sodium or potassium
nitrite (sodium or potassium nitrate(III)) with hydrochloric acid.
Nitrous acid is a weak acid and so you get the reaction:

http://www.chemguide.co.uk/organicprops/amines/nitrousacid.html (1 of 4)30/12/2004 11:38:18

amines and nitrous acid

Because nitrous acid is a weak acid, the position of equilibrium lies well
the right.

In each of the following reactions, the amine would be acidified with


hydrochloric acid and a solution of sodium or potassium nitrite added.
The acid and the nitrite form nitrous acid which then reacts with the
amine.

Primary amines and nitrous acid


The main observation is a burst of colourless, odourless gas. Nitrogen is
given off.
Unfortunately, there is no single clear-cut equation that you can quote for
this. You get lots of different organic products. For example, amongst the
products you get an alcohol where the -NH2 group has been replaced by
OH. If you want a single equation, you could quote (taking 1aminopropane as an example):

. . . but the propan-1-ol will be only one product among many - including
propan-2-ol, propene, 1-chloropropane, 2-chloropropane and others.
The nitrogen, however, is given off in quantities exactly as suggested by
the equation. By measuring the amount of nitrogen produced, you could
use this reaction to work out the amount of amine present in the solution.

http://www.chemguide.co.uk/organicprops/amines/nitrousacid.html (2 of 4)30/12/2004 11:38:18

amines and nitrous acid

Note: The reason for the complexity lies in how the reaction
happens. In the first instance, you get a diazonium ion
formed - for example, CH3CH2CH2N2+. Unless the -N2+
group is attached directly to a benzene ring, these ions are
very unstable and fall apart immediately to give nitrogen gas
and a carbocation (carbonium ion) - in this example,
CH3CH2CH2+.
It is rearrangements and reactions of this ion which lead to
the mess of products.
With aromatic amines like phenylamine (aniline) the
diazonium ion formed is much more stable. If you are
interested in reactions of diazonium ions (but only in the
context of amines like phenylamine), follow this link.
Use the BACK button on your browser if you want to return to
this page later.

Secondary amines and nitrous acid


This time there isn't any gas produced. Instead, you get a yellow oil
called a nitrosamine. These compounds are powerful carcinogens - avoid
them!
For example:

Tertiary amines and nitrous acid


Again, a quite different result. This time, nothing visually interesting
happens - you are left with a colourless solution.
http://www.chemguide.co.uk/organicprops/amines/nitrousacid.html (3 of 4)30/12/2004 11:38:18

amines and nitrous acid

All that has happened is that the amine has formed an ion by reacting
with the acid present. With trimethylamine, for example, you would get a
trimethylammonium ion, (CH3)3NH+.

Note: Textbooks often suggest the formation of a salt such


as trimethylammonium nitrite. This is actually a bit
misleading. The solution will contain trimethylammonium ions
and nitrite ions and also chloride ions from the hydrochloric
acid. There isn't any reason why the trimethylammonium ions
should be thought of as combining with the nitrite ions in
some way rather than with the chloride ions. They are all just
free-swimming ions, milling around in the solution.

Where would you like to go now?


To the amines menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/amines/nitrousacid.html (4 of 4)30/12/2004 11:38:18

making diazonium salts from phenylamine (aniline)

MAKING DIAZONIUM SALTS FROM


PHENYLAMINE

This page looks at the reaction between phenylamine (also known as


aniline and aminobenzene) and nitrous acid - particularly its reaction at
temperatures of less than 5C to produce diazonium salts. If you want to
know about the reactions of the diazonium ions formed, you will find a
link at the bottom of the page.

The reactions of phenylamine with nitrous acid


Nitrous acid (also known as nitric(III) acid) has the formula HNO2. It is
sometimes written as HONO to show the way it is joined up.
Nitrous acid decomposes very readily and is always made in situ. In the
case of its reaction with phenylamine, the phenylamine is first dissolved
in hydrochloric acid, and then a solution of sodium or potassium nitrite is
added. The reaction between the hydrochloric acid and the nitrite ions
produces the nitrous acid.
You get the reaction:

Because nitrous acid is a weak acid, the position of equilibrium lies well
the right.
Phenylamine reacts with nitrous acid differently depending on the
temperature.

The reaction on warming


If the mixture is warmed, you get a black oily product which contains
phenol (amongst other things), and nitrogen gas is given off.
http://www.chemguide.co.uk/organicprops/aniline/makediazo.html (1 of 4)30/12/2004 11:38:24

making diazonium salts from phenylamine (aniline)

Note: This is similar to the reaction between aliphatic


primary amines and nitrous acid.
Use the BACK button on your browser to return to this page
later if you choose to follow this link.
The explanation for the reaction (including the reason why
there is a different reaction in the hot or cold) can be found
on the page about reactions of diazonium ions (link at the
bottom of this page).

The reaction at low temperatures


The solution of phenylamine in hydrochloric acid (phenylammonium
chloride solution) is stood in a beaker of ice. The sodium or potassium
nitrite solution is also cooled in the ice.
The solution of the nitrite is then added very slowly to the
phenylammonium chloride solution - so that the temperature never goes
above 5C.
You end up with a solution containing benzenediazonium chloride:

The positive ion, containing the -N2+ group, is known as a diazonium


ion. The "azo" bit of the name refers to nitrogen.

http://www.chemguide.co.uk/organicprops/aniline/makediazo.html (2 of 4)30/12/2004 11:38:24

making diazonium salts from phenylamine (aniline)

Note: I've drawn the diazonium ion with the -N2+ group at
the side for two reasons. First, because it is easier to show
that the benzene ring and the two nitrogen atoms lie in a
straight line. Secondly, because that's the way I shall need to
draw it in some of its reactions to make certain structures
clear.

The ionic equation for the reaction is:

Notice that the chloride ions from the acid aren't involved in this in any
way. If you use hydrochloric acid, the solution will contain
benzenediazonium chloride. If you used a different acid, you would just
get a different salt - a sulphate or hydrogensulphate, for example, if you
used sulphuric acid.
The reactions of a diazonium salt are always done with a freshly
prepared solution made in this way. The solutions don't keep. Diazonium
salts are very unstable and tend to be explosive as solids.

Where would you like to go now?


To look at the reactions of diazonium ions . . .
To the phenylamine menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

http://www.chemguide.co.uk/organicprops/aniline/makediazo.html (3 of 4)30/12/2004 11:38:24

making diazonium salts from phenylamine (aniline)

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/aniline/makediazo.html (4 of 4)30/12/2004 11:38:24

some reactions of diazonium ions

REACTIONS OF DIAZONIUM SALTS

This page looks at some typical reactions of diazonium ions, including


examples of both substitution reactions and coupling reactions. If you
have come straight to this page from a search engine and want to know
about the preparation of the diazonium ions, you will find a link at the
bottom of the page.

Substitution reactions of diazonium ions


Diazonium ions are present in solutions such as benzenediazonium
chloride solution. They contain an -N2+ group. In the case of
benzenediazonium chloride, this is attached to a benzene ring.
Benzenediazonium chloride looks like this:

In this set of reactions of the diazonium ion, the -N2+ group is replaced by
something else. The nitrogen is released as nitrogen gas.
Substitution by an -OH group
To get this reaction, all you need to do is warm the benzenediazonium
chloride solution. The diazonium ion reacts with the water in the solution
and phenol is formed - either in solution or as a black oily liquid
(depending on how much is formed). Nitrogen gas is evolved.

http://www.chemguide.co.uk/organicprops/aniline/propsdiazo.html (1 of 7)30/12/2004 11:38:29

some reactions of diazonium ions

This is the same reaction that you get if you react phenylamine with
nitrous acid in the warm. The diazonium ion is formed first and then
immediately reacts with the water in the solution to give phenol.

Note: In this, and all the other reactions of diazonium ions on


this page, I am going to write the attached groups at the side
of the benzene ring rather than at the top. There's no very
sophisticated reason for this! When I write the equations for
the coupling reactions further down this page, that's the way I
want to draw the structures, because I think it makes them
clearer. I am just trying to keep everything else consistent with
that.

Substitution by an iodine atom


This is a good example of the use of diazonium salts to substitute things
into a benzene ring which are otherwise quite difficult to attach. (That's
equally true of the previous reaction, by the way.)
If you add potassium iodide solution to the benzenediazonium chloride
solution in the cold, nitrogen gas is given off, and you get oily droplets of
iodobenzene formed.
There is a simple reaction between the diazonium ions and the iodide ions
from the potassium iodide solution.

Coupling reactions of diazonium ions


In the substitution reactions above, the nitrogen in the diazonium ion is
lost. In the rest of the reactions on this page, the nitrogen is retained and
http://www.chemguide.co.uk/organicprops/aniline/propsdiazo.html (2 of 7)30/12/2004 11:38:29

some reactions of diazonium ions

used to make a bridge between two benzene rings.


The reaction with phenol
Phenol is dissolved in sodium hydroxide solution to give a solution of
sodium phenoxide.

The solution is cooled in ice, and cold benzenediazonium chloride solution


is added. There is a reaction between the diazonium ion and the
phenoxide ion and a yellow-orange solution or precipitate is formed.
The product is one of the simplest of what are known as azo compounds,
in which two benzene rings are linked by a nitrogen bridge.

Note: You might wonder where the hydrogen in the -OH


group comes from. It was pushed off that same ring when the
nitrogen became attached.
I have made a positive decision not to give names for the
products of these coupling reactions. The problem is that there
are lots of different variations on the names in common use.
Almost every book you look in calls them something different.
One web source quoted 45 versions for the compound
sometimes called "aniline yellow" (see below). Admittedly, 29
of these were terms like "fast spirit yellow" and "Sudan yellow
R", but 16 were variations on chemical names.
Any name which I give is quite likely to conflict with one which
you have got from another source - teacher, lecturer or
textbook. I'm not going to add to your confusion! Learn the
structures, and don't worry too much about the names.

http://www.chemguide.co.uk/organicprops/aniline/propsdiazo.html (3 of 7)30/12/2004 11:38:29

some reactions of diazonium ions

The reaction with naphthalen-2-ol


Naphthalen-2-ol is also known as 2-naphthol or beta-naphthol. It contains
an -OH group attached to a naphthalene molecule rather than to a simple
benzene ring. Naphthalene has two benzene rings fused together.
The reaction is done under exactly the same conditions as with phenol.
The naphthalen-2-ol is dissolved in sodium hydroxide solution to produce
an ion just like the phenol one. This solution is cooled and mixed with the
benzenediazonium chloride solution.
An intense orange-red precipitate is formed - another azo compound.

The reaction with phenylamine (aniline)


Some liquid phenylamine is added to a cold solution of benzenediazonium
chloride, and the mixture is shaken vigorously. A yellow solid is produced.

These strongly coloured azo compounds are frequently used as dyes


known as azo dyes. The one made from phenylamine (aniline) is known
as "aniline yellow" (amongst many other things - see note above).
Azo compounds account for more than half of modern dyes.

The use of an azo dye as an indicator - methyl orange

http://www.chemguide.co.uk/organicprops/aniline/propsdiazo.html (4 of 7)30/12/2004 11:38:29

some reactions of diazonium ions

Azo compounds contain a highly delocalised system of electrons which


takes in both benzene rings and the two nitrogen atoms bridging the rings.
The delocalisation can also extend to things attached to the benzene rings
as well.
If white light falls on one of these molecules, some wavelengths are
absorbed by these delocalised electrons. The colour you see is the result
of the non-absorbed wavelengths. The groups which contribute to the
delocalisation (and so to the absorption of light) are known as a
chromophore.

Note: A Canadian university site describes "chromophore" as


"one of those useful but sloppy words whose meaning
depends somewhat on the context."
Some sources take the chromophore as being the whole of the
two benzene rings and the nitrogen bridge. Others seem to
take it as just the -N=N- group.
I'm taking the line that it is the two benzene rings plus the N=N- group. If your examiners take the more restricted
meaning, obviously you should go with what they want. The
only way you will find that out is to look at recent exam papers
and mark schemes. If you are a UK A level student follow this
link to the syllabuses page to find out how to get hold of these
if you haven't already got them.

Modifying the groups present in the molecule can have an effect on the
light absorbed, and so on the colour you see. You can take advantage of
this in indicators.
Methyl orange is an azo dye which exists in two forms in equilibrium with
each other:

http://www.chemguide.co.uk/organicprops/aniline/propsdiazo.html (5 of 7)30/12/2004 11:38:29

some reactions of diazonium ions

As the hydrogen ion is lost or gained there is a shift in the exact nature of
the delocalisation in the molecule, and that causes a shift in the
wavelength of light absorbed. Obviously that means that you see a
different colour.
When you add acid to methyl orange, the equilibrium is shifted to give
more of the red form. Methyl orange is red in acidic solutions (in fact
solutions of pH less than 3.1).
If you add an alkali, hydrogen ions are removed and the equilibrium shifts
to give more of the yellow form. Methyl orange is yellow at pH's greater
than 4.4.
In between, at some point there will be equal amounts of the red and
yellow forms and so methyl orange looks orange.

Note: You can find out much more about indicators in the
physical chemistry part of this site.

http://www.chemguide.co.uk/organicprops/aniline/propsdiazo.html (6 of 7)30/12/2004 11:38:29

some reactions of diazonium ions

Where would you like to go now?


To look at the preparation of diazonium ions . . .
To the phenylamine menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/aniline/propsdiazo.html (7 of 7)30/12/2004 11:38:29

hydrogenation of alkenes

THE HYDROGENATION OF ALKENES

This page looks at the reaction of the carbon-carbon double bond in


alkenes with hydrogen in the presence of a metal catalyst. This is called
hydrogenation. It includes the manufacture of margarine from animal or
vegetable fats and oils.

Hydrogenation in the lab


The hydrogenation of ethene
Ethene reacts with hydrogen in the presence of a finely divided nickel
catalyst at a temperature of about 150C. Ethane is produced.

This is a fairly pointless reaction because ethene is a far more useful


compound than ethane! However, what is true of the reaction of the
carbon-carbon double bond in ethene is equally true of it in much more
complicated cases.

Note: You will find the mechanism for this hydrogenation


reaction described in some detail in the catalysis section of
this site.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/alkenes/hydrogenation.html (1 of 7)30/12/2004 11:38:41

hydrogenation of alkenes

Margarine manufacture
Some margarine is made by hydrogenating carbon-carbon double bonds
in animal or vegetable fats and oils. You can recognise the presence of
this in foods because the ingredients list will include words showing that
it contains "hydrogenated vegetable oils" or "hydrogenated fats".
The impression is sometimes given that all margarine is made by
hydrogenation - that's simply not true.
Animal and vegetable fats and oils
These are similar molecules, differing in their melting points. If the
compound is a solid at room temperature, you usually call it a fat. If it is a
liquid, it is often described as an oil.
Their melting points are largely determined by the presence of carboncarbon double bonds in the molecule. The higher the number of carboncarbon double bonds, the lower the melting point.
If there aren't any carbon-carbon double bonds, the substance is said to
be saturated. A typical saturated fat might have the structure:

Molecules of this sort are usually solid at room temperature.


If there is only one carbon-carbon double bond in each of the
hydrocarbon chains, it is called a mono-unsaturated fat (or monounsaturated oil, because it is likely to be a liquid at room temperature.)

http://www.chemguide.co.uk/organicprops/alkenes/hydrogenation.html (2 of 7)30/12/2004 11:38:41

hydrogenation of alkenes

A typical mono-unsaturated oil might be:

If there are two or more carbon-carbon double bonds in each chain, then
it is said to be polyunsaturated.
For example:

For simplicity, in all these diagrams, all three hydrocarbon chains in each
molecule are the same. That doesn't have to be the case - you can have
a mixture of types of chain in the same molecule.

Making margarine
Vegetable oils often contain high proportions of polyunsaturated and
mono-unsaturated fats (oils), and as a result are liquids at room
temperature. That makes them messy to spread on your bread or toast,
http://www.chemguide.co.uk/organicprops/alkenes/hydrogenation.html (3 of 7)30/12/2004 11:38:41

hydrogenation of alkenes

and inconvenient for some baking purposes.


You can "harden" (raise the melting point of) the oil by hydrogenating it in
the presence of a nickel catalyst. Conditions (like the precise
temperature, or the length of time the hydrogen is passed through the oil)
are carefully controlled so that some, but not necessarily all, of the
carbon-carbon double bonds are hydrogenated.
This produces a "partially hydrogenated oil" or "partially hydrogenated
fat".
You need to hydrogenate enough of the bonds to give the final texture
you want. However, there are possible health benefits in eating monounsaturated or polyunsaturated fats or oils rather than saturated ones so you wouldn't want to remove all the carbon-carbon double bonds.

Note: There is constant (and constantly changing)


controversy about the health risks and benefits of eating the
various types of fat. If you do a Google search on this topic,
you will find all sorts of conflicting information. Look carefully
at the reliability of the source of that information before you
try to make up your mind. Just because it is on the internet
doesn't mean that it is right!
You will find a Google search box on the Main Menu page
(link below). Don't forget that you need to search the whole
web and not just chemguide - otherwise you will just end up
back here!

http://www.chemguide.co.uk/organicprops/alkenes/hydrogenation.html (4 of 7)30/12/2004 11:38:41

hydrogenation of alkenes

The flow diagram below shows the complete hydrogenation of a typical


mono-unsaturated oil.

The downside of hydrogenation as a means of hardening fats and


oils
There are some probable health risks from eating hydrogenated fats or
oils. Consumers are becoming more aware of this, and manufacturers
are increasingly finding alternative ways of converting oils into
spreadable solids.
One of the problems arises from the hydrogenation process.
The double bonds in unsaturated fats and oils tend to have the groups
around them arranged in the "cis" form.

http://www.chemguide.co.uk/organicprops/alkenes/hydrogenation.html (5 of 7)30/12/2004 11:38:41

hydrogenation of alkenes

Note: If you don't know what this means, you will find more
about it towards the bottom of the introductory page about
esters. If you don't understand what cis and trans isomers
are, you won't make any sense of what comes next!
Use the BACK button on your browser to return to this page.

The relatively high temperatures used in the hydrogenation process tend


to flip some of the carbon-carbon double bonds into the "trans" form. If
these particular bonds aren't hydrogenated during the process, they will
still be present in the final margarine in molecules of trans fats.
The consumption of trans fats has been shown to increase cholesterol
levels (particularly of the more harmful LDL form) - leading to an
increased risk of heart disease.
Any process which tends to increase the amount of trans fat in the diet is
best avoided. Read food labels, and avoid any food which contains (or is
cooked in) hydrogenated oil or hydrogenated fat.

Note: As I said earlier, ideas about this change all the time.
It would be worth doing a Google search on trans fatty
acids. The US Food and Drugs Administration (FDA) has a
useful up-to-date fact sheet about them, and the Google
search will find it for you amongst a lot of other information.

http://www.chemguide.co.uk/organicprops/alkenes/hydrogenation.html (6 of 7)30/12/2004 11:38:41

hydrogenation of alkenes

Where would you like to go now?


To the alkenes menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2003 (modified 2004)

http://www.chemguide.co.uk/organicprops/alkenes/hydrogenation.html (7 of 7)30/12/2004 11:38:41

alkenes and hydrogen halides

ALKENES and HYDROGEN HALIDES

This page looks at the reaction of the carbon-carbon double bond in


alkenes such as ethene with hydrogen halides such as hydrogen chloride
and hydrogen bromide.
Symmetrical alkenes (like ethene or but-2-ene) are dealt with first. These
are alkenes where identical groups are attached to each end of the
carbon-carbon double bond. The extra problems associated with
unsymmetrical ones like propene are covered in a separate section
afterwards.

Addition to symmetrical alkenes


What happens?
All alkenes undergo addition reactions with the hydrogen halides. A
hydrogen atom joins to one of the carbon atoms originally in the double
bond, and a halogen atom to the other.
For example, with ethene and hydrogen chloride, you get chloroethane:

With but-2-ene you get 2-chlorobutane:

http://www.chemguide.co.uk/organicprops/alkenes/hhal.html (1 of 9)30/12/2004 11:38:48

alkenes and hydrogen halides

Note: Follow this link if you aren't happy about naming


organic compounds
Use the BACK button on your browser to return to this page.

What happens if you add the hydrogen to the carbon atom at the righthand end of the double bond, and the chlorine to the left-hand end? You
would still have the same product.
The chlorine would be on a carbon atom next to the end of the chain you would simply have drawn the molecule flipped over in space.
That would be different of the alkene was unsymmetrical - that's why we
have to look at them separately.
Conditions
The alkenes react with gaseous hydrogen halides at room temperature. If
the alkene is also a gas, you can simply mix the gases. If the alkene is a
liquid, you can bubble the hydrogen halide through the liquid.
Alkenes will also react with concentrated solutions of the gases in water.
A solution of hydrogen chloride in water is, of course, hydrochloric acid.
A solution of hydrogen bromide in water is hydrobromic acid - and so on.
There are, however, problems with this. The water will also get involved
in the reaction and you end up with a mixture of products.

http://www.chemguide.co.uk/organicprops/alkenes/hhal.html (2 of 9)30/12/2004 11:38:48

alkenes and hydrogen halides

Warning! The mechanism for this reaction is almost


invariably given for the reaction involving the alkene and the
simple molecules H-Cl or H-Br or whatever. In the presence
of water, these molecules will already have reacted with the
water to produce hydroxonium ions, H3O+, and halide ions.
The mechanism will therefore be different - involving an initial
attack by a hydroxonium ion. Avoid this problem by using the
pure gaseous hydrogen halide.
If you choose to follow this link to the mechanism, use the
BACK button on your browser to return to this page.

Reaction rates
Variation of rates when you change the halogen
Reaction rates increase in the order HF - HCl - HBr - HI. Hydrogen
fluoride reacts much more slowly than the other three, and is normally
ignored in talking about these reactions.
When the hydrogen halides react with alkenes, the hydrogen-halogen
bond has to be broken. The bond strength falls as you go from HF to HI,
and the hydrogen-fluorine bond is particularly strong. Because it is
difficult to break the bond between the hydrogen and the fluorine, the
addition of HF is bound to be slow.
Variation of rates when you change the alkene
This applies to unsymmetrical alkenes as well as to symmetrical ones.
For simplicity the examples given below are all symmetrical ones- but
they don't have to be.
Reaction rates increase as the alkene gets more complicated - in the
sense of the number of alkyl groups (such as methyl groups) attached to
the carbon atoms at either end of the double bond.
For example:

http://www.chemguide.co.uk/organicprops/alkenes/hhal.html (3 of 9)30/12/2004 11:38:48

alkenes and hydrogen halides

There are two ways of looking at the reasons for this - both of which
need you to know about the mechanism for the reactions.

Note: If you should know about the mechanism, but are a bit
uncertain about it, then you should spend some time
exploring the electrophilic addition mechanisms menu before
you go on, and then come back to this page later. You should
look at the addition of hydrogen halides to unsymmetrical
alkenes as well as symmetrical ones.
This will take you some time. Use the BACK button (or, more
efficiently, the HISTORY file or GO menu) on your browser to
return to this page.
If you don't need to know about the mechanisms, skip over
the next bit!

Alkenes react because the electrons in the pi bond attract things with any
degree of positive charge. Anything which increases the electron density
around the double bond will help this.
Alkyl groups have a tendency to "push" electrons away from themselves
towards the double bond. The more alkyl groups you have, the more
negative the area around the double bonds becomes.
The more negatively charged that region becomes, the more it will attract
molecules like hydrogen chloride.

http://www.chemguide.co.uk/organicprops/alkenes/hhal.html (4 of 9)30/12/2004 11:38:48

alkenes and hydrogen halides

Note: If you aren't sure about pi bonds, you will find a simple
mention of them in the introductory page on alkenes
You will find more about the electron pushing effect of alkyl
groups on a page about carbocations in the mechanism
section of this site. That is also important reading if you are to
understand the next bit.
Use the BACK button on your browser to return to this page
later.

The more important reason, though, lies in the stability of the


intermediate ion formed during the reaction. The three examples given
above produce these carbocations (carbonium ions) at the half-way
stage of the reaction:

The stability of the intermediate ions governs the activation energy for
the reaction. As you go towards the more complicated alkenes, the
activation energy for the reaction falls. That means that the reactions
become faster.

http://www.chemguide.co.uk/organicprops/alkenes/hhal.html (5 of 9)30/12/2004 11:38:48

alkenes and hydrogen halides

Didn't understand this? You should have followed the link


to the page about carbocations mentioned above!
Use the BACK button on your browser to return to this page
afterwards.

Addition to unsymmetrical alkenes


What happens?
In terms of reaction conditions and the factors affecting the rates of the
reaction, there is no difference whatsoever between these alkenes and
the symmetrical ones described above. The problem comes with the
orientation of the addition - in other words, which way around the
hydrogen and the halogen add across the double bond.
Orientation of addition
If HCl adds to an unsymmetrical alkene like propene, there are two
possible ways it could add. However, in practice, there is only one major
product.

http://www.chemguide.co.uk/organicprops/alkenes/hhal.html (6 of 9)30/12/2004 11:38:48

alkenes and hydrogen halides

This is in line with Markovnikov's Rule which says:


When a compound HX is added to an unsymmetrical alkene, the
hydrogen becomes attached to the carbon with the most
hydrogens attached to it already.
In this case, the hydrogen becomes attached to the CH2 group, because
the CH2 group has more hydrogens than the CH group.
Notice that only the hydrogens directly attached to the carbon atoms at
either end of the double bond count. The ones in the CH3 group are
totally irrelevant.

Warning! Markovnikov's Rule is a useful guide for you to


work out which way round to add something across a double
bond, but it isn't the reason why things add that way. As a
general principle, don't quote Markovnikov's Rule in an exam
unless you are specifically asked for it.
You will find the proper reason for this in a page about the
addition of hydrogen halides to unsymmetrical alkenes in the
mechanism section of this site.
Use the BACK button on your browser if you want to return to
this page.

A special problem with hydrogen bromide


Unlike the other hydrogen halides, hydrogen bromide can add to a
carbon-carbon double bond either way around - depending on the
conditions of the reaction.

If the hydrogen bromide and alkene are entirely pure


In this case, the hydrogen bromide adds on according to Markovnikov's

http://www.chemguide.co.uk/organicprops/alkenes/hhal.html (7 of 9)30/12/2004 11:38:48

alkenes and hydrogen halides

Rule. For example, with propene you would get 2-bromopropane.

That is exactly the same as the way the other hydrogen halides add.

If the hydrogen bromide and alkene contain traces of organic


peroxides
Oxygen from the air tends to react slowly with alkenes to produce some
organic peroxides, and so you don't necessarily have to add them
separately. This is therefore the reaction that you will tend to get unless
you take care to exclude all air from the system.
In this case, the addition is the other way around, and you get 1bromopropane:

This is sometimes described as an anti-Markovnikov addition or as the


peroxide effect.
Organic peroxides are excellent sources of free radicals. In the presence
of these, the hydrogen bromide reacts with alkenes using a different
(faster) mechanism. For various reasons, this doesn't happen with the
other hydrogen halides.
This reaction can also happen in this way in the presence of ultra-violet
light of the right wavelength to break the hydrogen-bromine bond into
hydrogen and bromine free radicals.

http://www.chemguide.co.uk/organicprops/alkenes/hhal.html (8 of 9)30/12/2004 11:38:48

alkenes and hydrogen halides

Note: All this is explored in detail on the page about free


radical addition of HBr to alkenes in the mechanism section
of this site.

Where would you like to go now?


To the alkenes menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/alkenes/hhal.html (9 of 9)30/12/2004 11:38:48

alkenes and sulphuric (sulfuric) acid

ALKENES and SULPHURIC ACID

This page looks at the reaction of the carbon-carbon double bond in


alkenes such as ethene with concentrated sulphuric acid. It includes the
conversion of the product into an alcohol.

The addition of sulphuric acid to alkenes


The reaction with ethene
Alkenes react with concentrated sulphuric acid in the cold to produce
alkyl hydrogensulphates. Ethene reacts to give ethyl hydrogensulphate.

The structure of the product molecule is sometimes written as


CH3CH2HSO4, but the version in the equation is better because it shows
how all the atoms are linked up. You may also find it written as
CH3CH2OSO3H.
Confused by all this? Don't be!
All you need to do is to learn the structure of sulphuric
acid. A hydrogen from the sulphuric acid joins on to one
of the carbon atoms, and the rest joins on to the other
one. Make sure that you can see how the structure of the
sulphuric acid relates to the various ways of writing the
formula for the product.

http://www.chemguide.co.uk/organicprops/alkenes/h2so4.html (1 of 4)30/12/2004 11:38:52

alkenes and sulphuric (sulfuric) acid

Important! Learn this structure for sulphuric acid. Sketch it


over and over again until you can't possibly get it wrong.
Follow this link if you want the mechanism for this reaction.
Use the BACK button on your browser to return to this page.

The reaction with propene


This is typical of the reaction with unsymmetrical alkenes. An
unsymmetrical alkene has different groups at either end of the carboncarbon double bond.
If sulphuric acid adds to an unsymmetrical alkene like propene, there are
two possible ways it could add. You could end up with one of two
products depending on which carbon atom the hydrogen attaches itself
to.
However, in practice, there is only one major product.

This is in line with Markovnikov's Rule which says:


When a compound HX is added to an unsymmetrical alkene, the
hydrogen becomes attached to the carbon with the most
hydrogens attached to it already.
In this case, the hydrogen becomes attached to the CH2 group, because
the CH2 group has more hydrogens than the CH group.
Notice that only the hydrogens directly attached to the carbon atoms at
either end of the double bond count. The ones in the CH3 group are
totally irrelevant.

http://www.chemguide.co.uk/organicprops/alkenes/h2so4.html (2 of 4)30/12/2004 11:38:52

alkenes and sulphuric (sulfuric) acid

Warning! Markovnikov's Rule is a useful guide for you to


work out which way round to add something across a double
bond, but it isn't the reason why things add that way. As a
general principle, don't quote Markovnikov's Rule in an exam
unless you are specifically asked for it.
You can find more about this in the mechanism section of this
site. You will find the mechanism for this reaction discussed
in detail if you follow this link.
Use the BACK button (or HISTORY file or GO menu, if you
have to explore several pages) on your browser if you want
to return to this page.

Using these reactions to make alcohols


Making ethanol
Ethene is passed into concentrated sulphuric acid to make ethyl
hydrogensulphate (as above). The product is diluted with water and then
distilled.
The water reacts with the ethyl hydrogensulphate to produce ethanol
which distils off.

Making propan-2-ol
More complicated alkyl hydrogensulphates react with water in exactly the
same way. For example:

http://www.chemguide.co.uk/organicprops/alkenes/h2so4.html (3 of 4)30/12/2004 11:38:52

alkenes and sulphuric (sulfuric) acid

Notice that the position of the -OH group is determined by where the
HSO4 group was attached. You get propan-2-ol rather than propan-1-ol
because of the way the sulphuric acid originally added across the double
bond in propene.
Using these reactions
These reactions were originally used as a way of manufacturing alcohols
from alkenes in the petrochemical industry. These days, alcohols like
ethanol or propan-2-ol tend to be manufactured by direct hydration of the
alkene because it is cheaper and easier.

Note: You can find out about the manufacture of alcohols by


direct hydration by following this link.

Where would you like to go now?


To the alkenes menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/alkenes/h2so4.html (4 of 4)30/12/2004 11:38:52

electrophilic addition - symmetrical alkenes and sulphuric acid

THE REACTION BETWEEN SYMMETRICAL


ALKENES AND SULPHURIC ACID

This page gives you the facts and a simple, uncluttered mechanism for
the electrophilic addition reactions between sulphuric acid and alkenes
like ethene and cyclohexene. If you want the mechanisms explained to
you in detail, there is a link at the bottom of the page.

The electrophilic addition reaction between ethene and


sulphuric acid
The facts
Alkenes react with concentrated sulphuric acid in the cold to produce
alkyl hydrogensulphates. Ethene reacts to give ethyl hydrogensulphate.

The structure of the product molecule is sometimes written as


CH3CH2HSO4, but the version in the equation is better because it shows
how all the atoms are linked up. You may also find it written as
CH3CH2OSO3H.
Confused by all this? Don't be!
All you need to do is to learn the structure of sulphuric
acid, and after that the mechanism is exactly the same as
the one with hydrogen bromide. As you will find out, the
formula of the product follows from the mechanism in an
inevitable way.

http://www.chemguide.co.uk/mechanisms/eladd/symh2so4.html (1 of 4)30/12/2004 11:39:01

electrophilic addition - symmetrical alkenes and sulphuric acid

Important! Learn this structure for sulphuric acid. Sketch it


over and over again until you can't possibly get it wrong.

The mechanism for the reaction between ethene and sulphuric acid
Sulphuric acid as an electrophile
The hydrogen atoms are attached to very electronegative oxygen atoms
which means that the hydrogens will have a slight positive charge while
the oxygens will be slightly negative. In the mechanism, we just focus on
one of the hydrogen to oxygen bonds, because the other one is too far
from the carbon-carbon double bond to be involved in any way.
The mechanism

Look carefully at the structure of the product so that you can see how it
relates to the various formulae given earlier (CH3CH2OSO2OH etc).

The electrophilic addition reaction between cyclohexene


and sulphuric acid
This time we are going straight for the mechanism without producing an
initial equation. This is to show that you can work out the structure of
obscure products provided you can write the mechanism.
The mechanism for the reaction between cyclohexene and
sulphuric acid
http://www.chemguide.co.uk/mechanisms/eladd/symh2so4.html (2 of 4)30/12/2004 11:39:01

electrophilic addition - symmetrical alkenes and sulphuric acid

Having worked out the structure of the product, you could then write a
simple equation for the reaction if you wanted to.

Where would you like to go now?


Help! Talk me through these mechanisms . . .
Look at the same reaction involving unsymmetrical
alkenes . . .
To menu of electrophilic addition reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .
http://www.chemguide.co.uk/mechanisms/eladd/symh2so4.html (3 of 4)30/12/2004 11:39:01

electrophilic addition - symmetrical alkenes and sulphuric acid

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/eladd/symh2so4.html (4 of 4)30/12/2004 11:39:01

Explaining electrophilic addition involving sulphuric acid

EXPLAINING THE REACTION BETWEEN


SYMMETRICAL ALKENES AND SULPHURIC
ACID

This page guides you through the mechanism for the electrophilic
addition of sulphuric acid to symmetrical alkenes like ethene or
cyclohexene. Unsymmetrical alkenes are covered separately, and you
will find a link at the bottom of the page.

The electrophilic addition reaction between ethene and


sulphuric acid
This reaction looks more complicated than the reaction between ethene
and hydrogen bromide, but it isn't! The only problem is that H2SO4 is a
more complicated structure than HBr. The mechanisms are exactly the
same.

Important! If you aren't sure about the reaction of ethene


with HBr follow this link before you read on.

The structure of sulphuric acid


Compare the structure of sulphuric acid with that of hydrogen bromide:

We are focussing on only one of the hydrogens in the sulphuric acid


because the other one will be pointing away from the double bond in the
alkene as the molecules approach each other.
http://www.chemguide.co.uk/mechanisms/eladd/symh2so4tt.html (1 of 7)30/12/2004 11:39:05

Explaining electrophilic addition involving sulphuric acid

In each case, the hydrogen is attached to a more electronegative


element, and so carries a slight positive charge. That means that the
hydrogen atoms will serve as electrophiles.

Electrophile: A substance with a strong attraction to a


negative region in another substance. Electrophiles are either
fully positive ions, or the slightly positive end of a polar
molecule.
If you aren't sure about electronegativity and polar bonds
follow this link before you read on.

When the sulphuric acid reacts, the whole of the shaded part of the
molecule remains as a complete unit. What happens to that unit is
exactly the same as happens to the bromine in the reaction involving HBr.
When you write the mechanisms involving sulphuric acid, keep that
shaded part unchanged throughout - apart from where you would change
the bromine. For example, you will need to put a lone pair and a negative
charge on the oxygen atom in the middle of the mechanism. That's
exactly what you had to do with the bromine in the HBr case.
The mechanism
The structure of ethene is shown in the
diagram on the right. The pi bond is an
orbital above and below the plane of the rest
of the molecule, and relatively exposed to
things around it. The two electrons in this
orbital are highly attractive to anything which
is positively charged.

http://www.chemguide.co.uk/mechanisms/eladd/symh2so4tt.html (2 of 7)30/12/2004 11:39:05

Explaining electrophilic addition involving sulphuric acid

Note: If you aren't sure about this, it would be a good idea to


read the introductory page on electrophilic addition before
you go on.
Use the BACK button on your browser to return to this page.

The slightly positive hydrogen atom in the sulphuric acid acts as an


electrophile, and is strongly attracted to the electrons in the pi bond.
The electrons from the pi bond move down towards the slightly positive
hydrogen atom.
In the process, the electrons in the hydrogen-oxygen bond are repelled
down until they are entirely on the oxygen atom, producing a negative ion.
So the first stage of the reaction is:

Help! If you aren't sure about the use of curly arrows in


mechanisms, you must follow this link before you go on.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/mechanisms/eladd/symh2so4tt.html (3 of 7)30/12/2004 11:39:05

Explaining electrophilic addition involving sulphuric acid

The ion with a positive charge on the carbon atom is called a


carbocation or carbonium ion (an older term).
Why is there a positive charge on the carbon atom? The pi bond was
originally made up of an electron from each of the carbon atoms. Both of
those electrons have been used to make a new bond to the hydrogen.
That leaves the right-hand carbon an electron short - hence positively
charged.
In the second stage of the mechanism, the lone pair of electrons on the
oxygen atom is strongly attracted to the positive carbon and moves
towards it until a bond is formed.

Note: There are other lone pairs around the oxygen atom as
well, but we are only showing one of them for clarity.

The overall mechanism is therefore

http://www.chemguide.co.uk/mechanisms/eladd/symh2so4tt.html (4 of 7)30/12/2004 11:39:05

Explaining electrophilic addition involving sulphuric acid

The electrophilic addition reaction between cyclohexene


and sulphuric acid
Once again

the pi bond breaks and the pair of electrons is used to form a bond
with the hydrogen atom;
the electrons in the hydrogen-oxygen bond are pushed on to the
oxygen atom giving it a full negative charge;
the lower carbon atom in the original C=C bond becomes
positively charged because the electron it originally supplied to the
pi bond has been moved away to form the new bond.

Note: Be prepared to draw the sulphuric acid various ways


around (on its side, upside-down, etc) so that it fits more tidily
into the mechanism you are writing.
Also: Be careful to attach the hydrogen to the correct carbon
atom. As the curly arrow has been drawn, you can think of
the electron pair pivotting around the top carbon atom. The
electrons stay attached to that carbon, and so that's the one
the hydrogen must join on to.

http://www.chemguide.co.uk/mechanisms/eladd/symh2so4tt.html (5 of 7)30/12/2004 11:39:05

Explaining electrophilic addition involving sulphuric acid

In the second stage, the lone pair on the negatively charged oxygen is
attracted towards the positively charge carbon and forms a bond with it.

The overall mechanism is therefore

Where would you like to go now?


Look at the same reactions involving unsymmetrical
alkenes . . .
To menu of electrophilic addition reactions. . .
To menu of other types of mechanism. . .

http://www.chemguide.co.uk/mechanisms/eladd/symh2so4tt.html (6 of 7)30/12/2004 11:39:05

Explaining electrophilic addition involving sulphuric acid

To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/eladd/symh2so4tt.html (7 of 7)30/12/2004 11:39:05

electrophilic addition - unsymmetrical alkenes and sulphuric acid

THE REACTION BETWEEN UNSYMMETRICAL


ALKENES AND SULPHURIC ACID

This page gives you the facts and a simple, uncluttered mechanism for
the electrophilic addition reactions between sulphuric acid and
unsymmetrical alkenes like propene. If you want the mechanism
explained to you in detail, there is a link at the bottom of the page.

The electrophilic addition reaction between propene and


sulphuric acid
The facts
Alkenes react with concentrated sulphuric acid in the cold to produce
alkyl hydrogensulphates. In the case of propene, the equation is:

Note: If you are confused about the structure of the product,


a similar structure is described in more detail in the page
about the reaction between sulphuric acid and symmetrical
alkenes.

http://www.chemguide.co.uk/mechanisms/eladd/unsymh2so4.html (1 of 3)30/12/2004 11:39:09

electrophilic addition - unsymmetrical alkenes and sulphuric acid

This is in line with Markovnikov's Rule which says:


When a compound HX is added to an unsymmetrical alkene,
the hydrogen becomes attached to the carbon with the most
hydrogens attached to it already.
In this case, the hydrogen becomes attached to the CH2 group, because
the CH2 group has more hydrogens than the CH group.
Notice that only the hydrogens directly attached to the carbon atoms at
either end of the double bond count. The ones in the CH3 group are
totally irrelevant.
The mechanism
This is an example of electrophilic addition.

The addition is this way around because it is easier to produce the


secondary carbocation (carbonium ion) than the primary one which would
be formed if the hydrogen became attached to the centre carbon atom
and the rest of the sulphuric acid to the end one.

Where would you like to go now?


Help! Talk me through this mechanism . . .
To menu of electrophilic addition reactions. . .
To menu of other types of mechanism. . .
http://www.chemguide.co.uk/mechanisms/eladd/unsymh2so4.html (2 of 3)30/12/2004 11:39:09

electrophilic addition - unsymmetrical alkenes and sulphuric acid

To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/eladd/unsymh2so4.html (3 of 3)30/12/2004 11:39:09

Explaining electrophilic addition involving sulphuric acid and unsymmetrical alkenes

EXPLAINING THE REACTION BETWEEN


UNSYMMETRICAL ALKENES AND SULPHURIC
ACID

This page guides you through the mechanism for the electrophilic
addition of sulphuric acid to unsymmetrical alkenes like propene.

The electrophilic addition reaction between propene and


sulphuric acid
Important! The reaction between propene and sulphuric acid
has exactly the same mechanism as the one involving
hydrogen bromide. It just looks more complicated because of
the structure of sulphuric acid.
Make sure that you understand the mechanism for the
reaction between propene and HBr before you read on.
It is also essential that you read about the reaction between
ethene and sulphuric acid before you worry about the
additional problem caused by an unsymmetrical alkene like
propene.

The structure of sulphuric acid


Compare the structure of sulphuric acid with that of hydrogen bromide:

http://www.chemguide.co.uk/mechanisms/eladd/unsymh2so4tt.html (1 of 5)30/12/2004 11:39:12

Explaining electrophilic addition involving sulphuric acid and unsymmetrical alkenes

We are focussing on only one of the hydrogens in the sulphuric acid


because the other one will be pointing away from the double bond in the
alkene as the molecules approach each other.
In each case, the hydrogen is attached to a more electronegative
element, and so carries a slight positive charge. That means that the
hydrogen atoms will serve as electrophiles.
When the sulphuric acid reacts, the whole of the shaded part of the
molecule remains as a complete unit. What happens to that unit is exactly
the same as happens to the bromine in the reaction involving HBr.
The mechanism
Remember that the active part of the double bond is the pi bond which
lies in an orbital above and below the plane of the rest of the molecule.
The pi bond is very vulnerable to attack.

Note: If this isn't fairly obvious to you, you really ought to


read the page introducing electrophilic addition before you go
on.
Use the BACK button on your browser to return to this page.

The slightly positive hydrogen atom in the sulphuric acid acts as an


electrophile, and is strongly attracted to the electrons in the pi bond.
As the sulphuric acid approaches the pi bond, the electrons in that bond
are drawn down towards the slightly positive hydrogen atom. That repels
the electrons in the hydrogen-oxygen bond down towards the oxygen.

http://www.chemguide.co.uk/mechanisms/eladd/unsymh2so4tt.html (2 of 5)30/12/2004 11:39:12

Explaining electrophilic addition involving sulphuric acid and unsymmetrical alkenes

The electron movements continue until a new bond is made between one
of the carbon atoms and the hydrogen. The oxygen now has both
electrons from the H-O bond, and so becomes negatively charged.
The problem is that there are two possible ways that the pi bond
electrons could move.
They could form a bond between the hydrogen and the left-hand carbon:

or they could form a bond with the right-hand one:

http://www.chemguide.co.uk/mechanisms/eladd/unsymh2so4tt.html (3 of 5)30/12/2004 11:39:12

Explaining electrophilic addition involving sulphuric acid and unsymmetrical alkenes

It's the second of these changes that happens more readily. In that case,
a secondary carbocation is formed - and that's more energetically stable
than the primary one formed in the first possibility.
Because the secondary ion is more energetically stable, it will form more
easily and so the reaction needs less activation energy - and so happens
faster.

Important! If you don't understand about the structure and


stability of carbocations (previously known as carbonium ions)
follow this link.

Once the ions have been formed, the lone pair on the negative oxygen is
strongly attracted towards the positive carbon atom. It moves towards it
and forms a bond.

That leaves you with the over-all mechanism:

Where would you like to go now?


http://www.chemguide.co.uk/mechanisms/eladd/unsymh2so4tt.html (4 of 5)30/12/2004 11:39:12

Explaining electrophilic addition involving sulphuric acid and unsymmetrical alkenes

To menu of electrophilic addition reactions. . .


To menu of other types of mechanism. . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/eladd/unsymh2so4tt.html (5 of 5)30/12/2004 11:39:12

hydration of alkenes to make alcohols

THE DIRECT HYDRATION OF ALKENES

This page looks at the production of alcohols by the direct hydration of


alkenes - adding water directly to the carbon-carbon double bond.

Manufacturing ethanol
Ethanol is manufactured by reacting ethene with steam. The reaction is
reversible.

Only 5% of the ethene is converted into ethanol at each pass through the
reactor. By removing the ethanol from the equilibrium mixture and
recycling the ethene, it is possible to achieve an overall 95% conversion.
A flow scheme for the reaction looks like this:

http://www.chemguide.co.uk/organicprops/alkenes/hydration.html (1 of 4)30/12/2004 11:39:16

hydration of alkenes to make alcohols

Note: This is a bit of a simplification! When the gases from


the reactor are cooled, then excess steam will condense as
well as the ethanol. The ethanol will have to be separated
from the water by fractional distillation.
All the sources I have looked at gloss over this, so I don't
have any details. I assume it is a normal fractional distillation
of an ethanol-water mixture.
You can find a full explanation of the reasons for the reaction
conditions by following this link to the physical chemistry part
of this site.
If you are interested in the mechanism for this reaction you
will find it by following this link.
Use the BACK button (or HISTORY file or GO menu) on your
browser if you want to return to this page later.

Manufacturing other alcohols


If you start from an unsymmetrical alkene like propene, you have to be
careful to think about which way around the water adds across the
carbon-carbon double bond.
Markovnikov's Rule says that when you add a molecule HX across a
carbon-carbon double bond, the hydrogen joins to the carbon atom which
already has the more hydrogen atoms attached to it.
Thinking of water as H-OH, the hydrogen will add to the carbon with the
more hydrogens already attached. That means that in the propene case,
you will get propan-2-ol rather than propan-1-ol.

http://www.chemguide.co.uk/organicprops/alkenes/hydration.html (2 of 4)30/12/2004 11:39:16

hydration of alkenes to make alcohols

The conditions used during manufacture vary from alcohol to alcohol.


The only conditions you will need for UK A level purposes are those for
making ethanol.

Note: I haven't included the mechanism for the hydration of


these more complicated alkenes anywhere on the site, but it
isn't too difficult to work out for yourself if you know the
mechanism for the hydration of ethene, and know about the
stability of carbocations (carbonium ions).
The water adds to the propene in the way shown above
because the secondary carbocation formed during the
process is more stable than the primary one formed if the
addition was the other way around.
These two pages are in two different parts of this site, so it
would be best to come back to this page using the BACK
button on your browser to get from one to the other if you are
interested.

Where would you like to go now?


To the alkenes menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

http://www.chemguide.co.uk/organicprops/alkenes/hydration.html (3 of 4)30/12/2004 11:39:16

hydration of alkenes to make alcohols

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/alkenes/hydration.html (4 of 4)30/12/2004 11:39:16

epoxyethane (ethylene oxide)

EPOXYETHANE ( ETHYLENE OXIDE )

This page looks at the manufacture of epoxyethane from ethene, and then
at some of the products that are made from epoxyethane.

The manufacture of epoxyethane


Conditions
Temperature:

about 250 - 300C

Pressure:

about 15 atmospheres

Catalyst:

silver

Problems and hazards during manufacture


The main problem comes in controlling the temperature. The reaction is
exothermic and so the temperature will tend to rise unless it is carefully
controlled.
At higher temperatures the ethene burns in the oxygen to produce carbon
dioxide and water which means that the temperature would increase even
more - and the whole thing get completely out of hand!
Two hazards during manufacture come from the nature of epoxyethane. It
is

poisonous and carcinogenic (cancer producing);

http://www.chemguide.co.uk/organicprops/alkenes/epoxyethane.html (1 of 5)30/12/2004 11:39:22

epoxyethane (ethylene oxide)

highly inflammable or explosive in contact with air.

The reactivity of epoxyethane


Ring strain
The reason that epoxyethane is so reactive is that bonding pairs in the ring
of atoms in the molecule are forced very close together. The bond angles
are about 60 rather than about 109.5 when carbon atoms normally form
single bonds.

The overlap between the atomic orbitals in forming the carbon-carbon and
carbon-oxygen bonds is less good than it is normally, and there is
considerable repulsion between the bonding pairs. The system becomes
more stable if the ring is broken.
When epoxyethane reacts a carbon-oxygen bond is always broken and the
ring opens up.

Uses of epoxyethane
Manufacture of ethane-1,2-diol (ethylene glycol)
Acid catalysed hydrolysis of epoxyethane
Epoxyethane reacts with water in the presence of an acid catalyst (very
dilute sulphuric acid) at a temperature of about 60C. Ethane-1,2-diol is
produced.

http://www.chemguide.co.uk/organicprops/alkenes/epoxyethane.html (2 of 5)30/12/2004 11:39:22

epoxyethane (ethylene oxide)

A large excess of water is used to try to prevent the product from reacting
with the original epoxyethane. Ethane-1,2-diol is an alcohol (because it
contains simple -OH groups), and alcohols react with epoxyethane (see
below).
Even in the presence of a large excess of water, this reaction happens as
well:

The product is still an alcohol, and similar reactions can also lead to quite
long chains.
Uses of ethane-1,2-diol
Ethane-1,2-diol is used as an antifreeze in car engines. It is added to the
cooling water to prevent it from freezing under very cold conditions.
Ethane-1,2-diol is also used in the manufacture of polyesters such as poly
(ethylene terephthalate). You may have come across this as a fibre used to
make clothes (perhaps under the brand name Terylene), or as a clear
material used to make plastic drinks bottles (PET).

The reaction of epoxyethane with alcohols


This is a reaction which students at this level often find difficulty
remembering. It is actually probably easier to work out than remember.
Think of it as an extension of the reaction with water.
Alcohols have the formula R-OH, where R is an alkyl group. Water can be
thought of as H-OH.
The reaction of epoxyethane with water can be colour-coded like this:

http://www.chemguide.co.uk/organicprops/alkenes/epoxyethane.html (3 of 5)30/12/2004 11:39:22

epoxyethane (ethylene oxide)

Now do the same thing with the alcohol:

Product molecules of this type are used as solvents.


Notice that the product is still an alcohol. It has an -OH group at the righthand end of the molecule. If the epoxyethane is in excess, the reaction can
continue. (In fact, it continues to some extent even if the epoxyethane isn't
in excess.)

The product from this reaction is again an alcohol, and can go on to react
with even more epoxyethane! What you get eventually is a chain with a
structure:

Compounds of this type are used as plasticisers (added, for example, to


PVC to make it more flexible) or as non-ionic surfactants (detergents). To
make the surfactant, you would start with a fairly long chain alcohol to
produce a molecule such as:

Where would you like to go now?


To the alkenes menu . . .
To the menu of other organic compounds . . .

http://www.chemguide.co.uk/organicprops/alkenes/epoxyethane.html (4 of 5)30/12/2004 11:39:22

epoxyethane (ethylene oxide)

To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/alkenes/epoxyethane.html (5 of 5)30/12/2004 11:39:22

Phenol Menu

Understanding Chemistry

PHENOL MENU

Background . . .
An introduction to phenol and its physical properties.
The acidity of phenol . . .
Including its reaction with sodium and sodium hydroxide, and its
lack of reaction with carbonates and hydrogencarbonates.
Ring reactions of phenol . . .
The reactions of phenol with bromine and with nitric acid.
Making an azo dye from phenol . . .
The coupling reaction of phenol with benzenediazonium chloride
solution to make an azo dye. The reaction you want is one of
several on the page that this will take you to. (That page is in a
different part of the site - use the BACK button on your browser if
you need to come back here again.)
Other reactions of phenol . . .
The combustion of phenol, its esterification, and its reaction with
iron(III) chloride solution (ferric chloride solution).

Go to menu of other organic compounds . . .


http://www.chemguide.co.uk/organicprops/phenolmenu.html (1 of 2)30/12/2004 11:39:25

Phenol Menu

Go to Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/phenolmenu.html (2 of 2)30/12/2004 11:39:25

an introduction to phenol

INTRODUCING PHENOL

This page looks at the structure and physical properties of phenol (very
old name: carbolic acid). Phenol is the simplest member of a family of
compounds in which an -OH group is attached directly to a benzene ring.
Phenol itself is the only one of the family that you are likely to need to
know about for UK A level purposes.

The structure of phenol


The simplest way to draw the structure of phenol is:

. . . but to understand phenol properly, you need to dig a bit deeper than
this.

Warning! You need to understand about the bonding in


benzene in order to make sense of this next bit.
If you follow this link, you may have to explore several other
pages before you are ready to come back here again. Use
the BACK button (or HISTORY file or GO menu) on your
browser to return to this page.

http://www.chemguide.co.uk/organicprops/phenol/background.html (1 of 4)30/12/2004 11:39:28

an introduction to phenol

There is an interaction between the delocalised electrons in the benzene


ring and one of the lone pairs on the oxygen atom. This has an important
effect on both the properties of the ring and of the -OH group.
One of the lone pairs on the oxygen overlaps with the delocalised ring
electron system . . .

. . . giving a structure rather like this:

The donation of the oxygen's lone pair into the ring system increases the
electron density around the ring. That makes the ring much more
reactive than it is in benzene itself. That is explored in another page in
this phenol section.
It also helps to make the -OH group's hydrogen a lot more acidic than it
is in alcohols. That will also be explored elsewhere in this section.

Physical properties
http://www.chemguide.co.uk/organicprops/phenol/background.html (2 of 4)30/12/2004 11:39:28

an introduction to phenol

Pure phenol is a white crystalline solid, smelling of disinfectant. It has to


be handled with great care because it causes immediate white blistering
to the skin. The crystals are often rather wet and discoloured.
Melting and boiling points
It is useful to compare phenol's melting and boiling points with those of
methylbenzene (toluene). Both molecules contain the same number of
electrons and are a very similar shape. That means that the
intermolecular attractions due to van der Waals dispersion forces are
going to be very similar.

Note: If you aren't happy about intermolecular forces


(including van der Waals dispersion forces and hydrogen
bonds) then you really ought to follow this link before you go
on. This section won't make much sense to you if you aren't
familiar with the various sorts of intermolecular forces.
Use the BACK button on your browser to return to this page.

melting point (C) boiling point (C)


C6H5OH

40 - 43

182

C6H5CH3

-95.0

111

Note: The melting point of phenol is quoted variously by


different sources within this 40 - 43C range. I suspect (but
don't know for sure) that the problem lies in the difficulty of
getting the phenol absolutely dry.

http://www.chemguide.co.uk/organicprops/phenol/background.html (3 of 4)30/12/2004 11:39:28

an introduction to phenol

The reason for the higher values for phenol is in part due to permanent
dipole-dipole attractions due to the electronegativity of the oxygen - but is
mainly due to hydrogen bonding.
Hydrogen bonds can form between a lone pair on an oxygen on one
molecule and the hydrogen on the -OH group of one of its neighbours.

Solubility in water
Phenol is moderately soluble in water - about 8 g of phenol will dissolve
in 100 g of water.
If you try to dissolve more than this, you get two layers of liquid. The top
layer is a solution of phenol in water, and the bottom one a solution of
water in phenol. The solubility behaviour of phenol and water is
complicated, and beyond UK A level.
Phenol is somewhat soluble in water because of its ability to form
hydrogen bonds with the water.

Where would you like to go now?


To the phenol menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/phenol/background.html (4 of 4)30/12/2004 11:39:28

the acidity of phenol

THE ACIDITY OF PHENOL

This page explains why phenol is a weak acid and looks at its reactions
(or in some cases, lack of reaction) with bases and with sodium metal.

Why is phenol acidic?


Unlike alcohols (which also contain an -OH group) phenol is a weak acid.
A hydrogen ion can break away from the -OH group and transfer to a
base.
For example, in solution in water:

Phenol is a very weak acid and the position of equilibrium lies well to the
left.
Phenol can lose a hydrogen ion because the phenoxide ion formed is
stabilised to some extent. The negative charge on the oxygen atom is
delocalised around the ring. The more stable the ion is, the more likely it
is to form.
One of the lone pairs on the oxygen atom overlaps with the delocalised
electrons on the benzene ring.

http://www.chemguide.co.uk/organicprops/phenol/acidity.html (1 of 5)30/12/2004 11:39:32

the acidity of phenol

This overlap leads to a delocalisation which extends from the ring out
over the oxygen atom. As a result, the negative charge is no longer
entirely localised on the oxygen, but is spread out around the whole ion.

Spreading the charge around makes the ion more stable than it would be
if all the charge remained on the oxygen.
However . . . oxygen is the most electronegative element in the ion and
the delocalised electrons will be drawn towards it. That means that there
will still be a lot of charge around the oxygen which will tend to attract the
hydrogen ion back again.
That's why phenol is only a very weak acid.

http://www.chemguide.co.uk/organicprops/phenol/acidity.html (2 of 5)30/12/2004 11:39:32

the acidity of phenol

Note: You will find more about the acidity of phenol,


including a comparison between it and other things like
carboxylic acids and alcohols on a page about organic acids
in a different part of the site. You will also find links from that
page to pages about the structure of benzene if that is
worrying you.
If you follow this link, you may have to explore several other
pages before you are ready to come back here again. Use
the BACK button (or HISTORY file or GO menu) on your
browser to return to this page.

Properties of phenol as an acid


With indicators
The pH of a typical dilute solution of phenol in water is likely to be around
5 - 6 (depending on its concentration). That means that a very dilute
solution isn't really acidic enough to turn litmus paper fully red. Litmus
paper is blue at pH 8 and red at pH 5. Anything in between is going to
show as some shade of "neutral".

With sodium hydroxide solution


Phenol reacts with sodium hydroxide solution to give a colourless
solution containing sodium phenoxide.

In this reaction, the hydrogen ion has been removed by the strongly
http://www.chemguide.co.uk/organicprops/phenol/acidity.html (3 of 5)30/12/2004 11:39:32

the acidity of phenol

basic hydroxide ion in the sodium hydroxide solution.

With sodium carbonate or sodium hydrogencarbonate


Phenol isn't acidic enough to react with either of these. Or, looked at
another way, the carbonate and hydrogencarbonate ions aren't strong
enough bases to take a hydrogen ion from the phenol.
Unlike the majority of acids, phenol doesn't give carbon dioxide when you
mix it with one of these.
This lack of reaction is actually useful. You can recognise phenol
because:

It is fairly insoluble in water.


It reacts with sodium hydroxide solution to give a colourless
solution (and therefore must be acidic).
It doesn't react with sodium carbonate or hydrogencarbonate
solutions (and so must be only very weakly acidic).

With metallic sodium


Acids react with the more reactive metals to give hydrogen gas. Phenol
is no exception - the only difference is the slow reaction because phenol
is such a weak acid.
Phenol is warmed in a dry tube until it is molten, and a small piece of
sodium added. There is some fizzing as hydrogen gas is given off. The
mixture left in the tube will contain sodium phenoxide.

http://www.chemguide.co.uk/organicprops/phenol/acidity.html (4 of 5)30/12/2004 11:39:32

the acidity of phenol

Warning! Under no circumstances should you try this


without professional supervision, and good access to medical
help. The risks involved in careless handling of the hot
phenol and sodium are too great.

Where would you like to go now?


To the phenol menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/phenol/acidity.html (5 of 5)30/12/2004 11:39:32

ring reactions of phenol

RING REACTIONS OF PHENOL

This page looks at the way the reactions of the benzene ring in phenol
are modified by the presence of the attached -OH group. It covers the
reactions of phenol with bromine water and with nitric acid.

How does the -OH group modify the ring reactions?


Activation of the ring
The -OH group attached to the benzene ring in phenol has the effect of
making the ring much more reactive than it would otherwise be.
For example, as you will find below, phenol will react with a solution of
bromine in water (bromine water) in the cold and in the absence of any
catalyst. It also reacts with dilute nitric acid, whereas benzene itself
needs a nitrating mixture of concentrated nitric acid and concentrated
sulphuric acid.

Warning! You need to understand about the bonding in


benzene in order to make sense of this next bit.
If you follow this link, you may have to explore several other
pages before you are ready to come back here again. Use
the BACK button (or HISTORY file or GO menu) on your
browser to return to this page.

http://www.chemguide.co.uk/organicprops/phenol/ring.html (1 of 5)30/12/2004 11:39:37

ring reactions of phenol

One of the lone pairs on the oxygen atom in the -OH group overlaps with
the delocalised ring electron system . . .

. . . giving a structure rather like this:

The donation of the oxygen's lone pair into the ring system increases the
electron density around the ring.
A benzene ring undergoes substitution reactions in which the ring
electrons are attacked by positive ions or the slightly positive parts of
molecules. In other words, it undergoes electrophilic substitution. If you
increase the electron density around the ring, it becomes even more
attractive to incoming electrophiles. That's what happens in phenol.

http://www.chemguide.co.uk/organicprops/phenol/ring.html (2 of 5)30/12/2004 11:39:37

ring reactions of phenol

Note: If you aren't sure what any of these terms mean, you
could find out by following this link to a page introducing
electrophilic substitution.
If you choose to follow this link, use the BACK button (or the
HISTORY file or GO menu) on your browser to return to this
page later.

The directing effect of the -OH group


The -OH group has more activating effect on some positions around the
ring than others (for reasons which go beyond UK A level). That means
that incoming groups will go into some positions much faster than they
will into others.
The net effect of this is that the -OH group has a 2,4-directing effect.
That means that incoming groups will tend to go into the 2- position (next
door to the -OH group) or the 4- position (opposite the -OH group). You
will get hardly any of the 3- isomer formed - it is produced too slowly.

Specific examples
Reaction with bromine water
If bromine water is added to a solution of phenol in water, the bromine
water is decolourised and a white precipitate is formed which smells of
antiseptic.
The precipitate is 2,4,6-tribromophenol. (The smell is similar to the
antiseptic TCP: 2,4,6-trichlorophenol.)

http://www.chemguide.co.uk/organicprops/phenol/ring.html (3 of 5)30/12/2004 11:39:37

ring reactions of phenol

Notice the multiple substitution around the ring - into all the activated
positions. (The 6- position is, of course, just the same as the 2- position.
Both are next door to the -OH group.)

Note: Bromine water is normally used as a test for a C=C


double bond. The important difference with phenol is the
formation of a white precipitate as well as the bromine water
being decolourised.
If you choose to follow this link, use the BACK button on your
browser to return to this page.

Reactions with nitric acid


The reactions with nitric acid are complicated because nitric acid is an
oxidising agent, and phenol is very easily oxidised to give complex tarry
products. What follows misses all that complication out, and just
concentrates on the ring substitution which happens as well.
With dilute nitric acid
Phenol reacts with dilute nitric acid at room temperature to give a mixture
of 2-nitrophenol and 4-nitrophenol.

http://www.chemguide.co.uk/organicprops/phenol/ring.html (4 of 5)30/12/2004 11:39:37

ring reactions of phenol

With concentrated nitric acid


With concentrated nitric acid, more nitro groups substitute around the
ring to give 2,4,6-trinitrophenol (old name: picric acid).

Where would you like to go now?


To the phenol menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/phenol/ring.html (5 of 5)30/12/2004 11:39:37

some more reactions of phenol

ASSORTED REACTIONS OF PHENOL

This page gives details of some reactions of phenol not covered


elsewhere in this section. It deals with the combustion and esterification
of phenol, and the use of iron(III) chloride solution (ferric chloride solution)
as a test for phenol.

Combustion of phenol
Phenol burns in a plentiful supply of oxygen to give carbon dioxide and
water.

However, for compounds containing benzene rings, combustion is hardly


ever complete, especially if they are burnt in air. The high proportion of
carbon in phenol means that you need a very high proportion of oxygen to
phenol to get complete combustion. Look at the equation.
As a general rule, the hydrogen in a molecule tends to get what oxygen is
available first, leaving the carbon to form carbon itself, or carbon
monoxide, if there isn't enough oxygen to go round.
Phenol tends to burn in air with an extremely smoky flame - full of carbon
particles.

Esterification of phenol
You will probably remember that you can make esters from alcohols by
reacting them with carboxylic acids. You might expect phenol to be similar.

http://www.chemguide.co.uk/organicprops/phenol/other.html (1 of 5)30/12/2004 11:39:43

some more reactions of phenol

Note: If you aren't sure about esters, you would probably do


better to skip the next bit and instead read the page about
making esters (which includes esters made from phenol).
There is much more detail on that page.
If you choose to follow this link, use the BACK button on your
browser if you want to return to this page later.

However, unlike alcohols, phenol reacts so slowly with carboxylic acids


that you normally react it with acyl chlorides (acid chlorides) or acid
anhydrides instead.

Making esters from phenol using an acyl chloride


A typical acyl chloride is ethanoyl chloride, CH3COCl.
Phenol reacts with ethanoyl chloride at room temperature, although the
reaction isn't as fast as the one between ethanoyl chloride and an alcohol.
Phenyl ethanoate is formed together with hydrogen chloride gas.

Sometimes it is necessary to modify the phenol first to make the reaction


faster.
For example, benzoyl chloride has the formula C6H5COCl. The -COCl
group is attached directly to a benzene ring. It is much less reactive than
simple acyl chlorides like ethanoyl chloride.
In order to get a reasonably quick reaction with benzoyl chloride, the
phenol is first converted into sodium phenoxide by dissolving it in sodium
http://www.chemguide.co.uk/organicprops/phenol/other.html (2 of 5)30/12/2004 11:39:43

some more reactions of phenol

hydroxide solution.

The phenoxide ion reacts more rapidly with benzoyl chloride than the
original phenol does, but even so you have to shake it with benzoyl
chloride for about 15 minutes. Solid phenyl benzoate is formed.

Making esters from phenol using an acid anhydride


A typical acid anhydride is ethanoic anhydride, (CH3CO)2O.
The reactions of acid anhydrides are slower than the corresponding
reactions with acyl chlorides, and you usually need to warm the mixture.
Again, you can react the phenol with sodium hydroxide solution first,
producing the more reactive phenoxide ion.
If you simply use phenol and ethanoic anhydride, phenyl ethanoate is
formed together with ethanoic acid.

This reaction isn't important itself, but a very similar reaction is involved in
the manufacture of aspirin (covered in detail on another page - link below).

http://www.chemguide.co.uk/organicprops/phenol/other.html (3 of 5)30/12/2004 11:39:43

some more reactions of phenol

If the phenol is first converted into sodium phenoxide by adding sodium


hydroxide solution, the reaction is faster. Phenyl ethanoate is again
formed, but this time the other product is sodium ethanoate rather than
ethanoic acid.

Note: These reactions (including the formation of aspirin) are


discussed in more detail on a page about reactions of acid
anhydrides.
Use the BACK button on your browser to return to this page
later if you want to.

The reaction with iron(III) chloride solution


Iron(III) chloride is sometimes known as ferric chloride.
Iron(III) ions form strongly coloured complexes with several organic
compounds including phenol. The colour of the complexes vary from
compound to compound. The reaction with iron(III) chloride solution can
be used as a test for phenol.
If you add a crystal of phenol to iron(III) chloride solution, you get an
intense violet-purple solution formed.

http://www.chemguide.co.uk/organicprops/phenol/other.html (4 of 5)30/12/2004 11:39:43

some more reactions of phenol

Note: The test is often done using "neutral" iron(III) chloride


solution. Dilute ammonia solution is added dropwise to iron(III)
chloride solution to give a faint precipitate of iron(III)
hydroxide. Then that precipitate is removed by adding a small
amount of the original iron(III) chloride solution.

Where would you like to go now?


To the phenol menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/phenol/other.html (5 of 5)30/12/2004 11:39:43

combustion of alkanes and cycloalkanes

THE COMBUSTION OF ALKANES AND


CYCLOALKANES

This page deals briefly with the combustion of alkanes and cycloalkanes.
In fact, there is very little difference between the two.

Complete combustion
Complete combustion (given sufficient oxygen) of any hydrocarbon
produces carbon dioxide and water.
Equations
It is quite important that you can write properly balanced equations for
these reactions, because they often come up as a part of
thermochemistry calculations. Don't try to learn the equations - there are
far too many possibilities. Work them out as you need them.
Some are easier than others. For example, with alkanes, the ones with
an even number of carbon atoms are marginally harder than those with
an odd number!
For example, with propane (C3H8), you can balance the carbons and
hydrogens as you write the equation down. Your first draft would be:

Counting the oxygens leads directly to the final version:

With butane (C4H10), you can again balance the carbons and hydrogens
as you write the equation down.

http://www.chemguide.co.uk/organicprops/alkanes/oxygen.html (1 of 5)30/12/2004 11:39:47

combustion of alkanes and cycloalkanes

Counting the oxygens leads to a slight problem - with 13 on the righthand side. The simple trick is to allow yourself to have "six-and-a-half"
O2 molecules on the left.

If that offends you, double everything:

Note: You might well come across either version of these


equations. The ones with the halves left in are often used in
calculation work.
Forgive me if you find this last bit on equations unbearably
trivial - not everybody does! Just be grateful that you have
been well taught.

Trends
The hydrocarbons become harder to ignite as the molecules get bigger.
This is because the bigger molecules don't vaporise so easily - the
reaction is much better if the oxygen and the hydrocarbon are well mixed
as gases. If the liquid isn't very volatile, only those molecules on the
surface can react with the oxygen.
Bigger molecules have greater Van der Waals attractions which makes it
more difficult for them to break away from their neighbours and turn to a
gas.

http://www.chemguide.co.uk/organicprops/alkanes/oxygen.html (2 of 5)30/12/2004 11:39:47

combustion of alkanes and cycloalkanes

Note: If you aren't sure about Van der Waals forces, then
you should follow this link before you go on.
Use the BACK button on your browser to return to this page.

Provided the combustion is complete, all the hydrocarbons will burn with
a blue flame. However, combustion tends to be less complete as the
number of carbon atoms in the molecules rises. That means that the
bigger the hydrocarbon, the more likely you are to get a yellow, smoky
flame.

Incomplete combustion
Incomplete combustion (where there isn't enough oxygen present) can
lead to the formation of carbon or carbon monoxide.
As a simple way of thinking about it, the hydrogen in the hydrocarbon
gets the first chance at the oxygen, and the carbon gets whatever is left
over!
The presence of glowing carbon particles in a flame turns it yellow, and
black carbon is often visible in the smoke. Carbon monoxide is produced
as a colourless poisonous gas.
Why carbon monoxide is poisonous
Oxygen is carried around the blood by haemoglobin (US: hemoglobin).
Unfortunately carbon monoxide binds to exactly the same site on the
haemoglobin that oxygen does.
The difference is that carbon monoxide binds irreversibly - making that
particular molecule of haemoglobin useless for carrying oxygen. If you
breath in enough carbon monoxide you will die from a sort of internal
suffocation.

http://www.chemguide.co.uk/organicprops/alkanes/oxygen.html (3 of 5)30/12/2004 11:39:47

combustion of alkanes and cycloalkanes

Note: There is more about haemoglobin towards the bottom


of the page about complex ions.
If you want some description of catalytic converters which
help to remove carbon monoxide and some other pollutants,
see the introductory page on catalysis.
If you want full details about any of the environmental
problems associated with burning hydrocarbons, you can't do
better than explore the excellent US Environmental
Protection Agency site.
If you want details about the role of hydrocarbons in the
formation of photochemical smog, a Google search on
photochemical smog will quickly lead you to some detailed
chemistry. There is a Google search box on the Main Menu
(link below). Don't forget that you want to search the whole
web - not chemguide.
Use the BACK button (or the HISTORY or GO menus) on
your browser if you want to return to this page later.

Where would you like to go now?


To the alkanes menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/alkanes/oxygen.html (4 of 5)30/12/2004 11:39:47

combustion of alkanes and cycloalkanes

http://www.chemguide.co.uk/organicprops/alkanes/oxygen.html (5 of 5)30/12/2004 11:39:47

alkanes and cycloalkanes with chlorine or bromine

THE HALOGENATION OF ALKANES AND


CYCLOALKANES

This page describes the reactions between alkanes and cycloalkanes


with the halogens fluorine, chlorine, bromine and iodine - mainly
concentrating on chlorine and bromine.

Alkanes
The reaction between alkanes and fluorine
This reaction is explosive even in the cold and dark, and you tend to get
carbon and hydrogen fluoride produced. It is of no particular interest. For
example:

The reaction between alkanes and iodine


Iodine doesn't react with the alkanes - at least, under normal lab
conditions.

Note: I can't discover the truth about this! Some sources say
that it doesn't react; others say that it reacts very slowly. I
have found it impossible to find all the data I need to estimate
whether there might be a temperature at which the reaction
becomes feasible. The information needed depends too
much on what assumptions you make about the physical
states of the reactants and products - for example, whether
the iodine is present as a solid, a gas or a solution.
If you have any hard information, could you contact me via
the address on the about this site page.

http://www.chemguide.co.uk/organicprops/alkanes/halogenation.html (1 of 6)30/12/2004 11:39:51

alkanes and cycloalkanes with chlorine or bromine

The reactions between alkanes and chlorine or bromine


There is no reaction in the dark.
In the presence of a flame, the reactions are rather like the fluorine one producing a mixture of carbon and the hydrogen halide. The violence of
the reaction drops considerably as you go from fluorine to chlorine to
bromine.
The interesting reactions happen in the presence of ultra-violet light
(sunlight will do). These are photochemical reactions, and happen at
room temperature.
We'll look at the reactions with chlorine. The reactions with bromine are
similar, but rather slower.
Methane and chlorine
Substitution reactions happen in which hydrogen atoms in the methane
are replaced one at a time by chlorine atoms. You end up with a mixture
of chloromethane, dichloromethane, trichloromethane and
tetrachloromethane.

http://www.chemguide.co.uk/organicprops/alkanes/halogenation.html (2 of 6)30/12/2004 11:39:51

alkanes and cycloalkanes with chlorine or bromine

Note: Follow this link if you aren't happy about naming


organic compounds.
Use the BACK button on your browser to return to this page.

The original mixture of a colourless and a green gas would produce


steamy fumes of hydrogen chloride and a mist of organic liquids. All of
the organic products are liquid at room temperature with the exception of
the chloromethane which is a gas.
If you were using bromine, you could either mix methane with bromine
vapour, or bubble the methane through liquid bromine - in either case,
exposed to UV light. The original mixture of gases would, of course, be
red-brown rather than green.
You wouldn't choose to use these reactions as a means of preparing
these organic compounds in the lab because the mixture of products
would be too tedious to separate.
The mechanisms for the reactions are explained on separate pages.

Note: If you want the methane-chlorine mechanism, follow


this link.
If you want the methane-bromine mechanism, follow this one.
Use the BACK button on your browser to return to this page.

http://www.chemguide.co.uk/organicprops/alkanes/halogenation.html (3 of 6)30/12/2004 11:39:51

alkanes and cycloalkanes with chlorine or bromine

Larger alkanes and chlorine


You would again get a mixture of substitution products, but it is worth just
looking briefly at what happens if only one of the hydrogen atoms gets
substituted (monosubstitution) - just to show that things aren't always as
straightforward as they seem!
For example, with propane, you could get one of two isomers:

Note: If you aren't sure about isomerism, you might like to


follow this link.
Use the BACK button on your browser to return to this page.

If chance was the only factor, you would expect to get 3 times as much of
the isomer with the chlorine on the end. There are 6 hydrogens that
could get replaced on the end carbon atoms compared with only 2 in the
middle.
In fact, you get about the same amount of each of the two isomers.
If you use bromine instead of chlorine, the great majority of the product is
where the bromine is attached to the centre carbon atom.
The reasons for this are beyond UK A level chemistry.

http://www.chemguide.co.uk/organicprops/alkanes/halogenation.html (4 of 6)30/12/2004 11:39:51

alkanes and cycloalkanes with chlorine or bromine

Cycloalkanes
The reactions of the cycloalkanes are generally just the same as the
alkanes, with the exception of the very small ones - particularly
cyclopropane.
The extra reactivity of cyclopropane
In the presence of UV light, cyclopropane will undergo substitution
reactions with chlorine or bromine just like a non-cyclic alkane. However,
it also has the ability to react in the dark.
In the absence of UV light, cyclopropane can undergo addition
reactions in which the ring is broken. For example, with bromine,
cyclopropane gives 1,3-dibromopropane.

This can still happen in the presence of light - but you will get substitution
reactions as well.
The ring is broken because cyclopropane suffers badly from ring strain.
The bond angles in the ring are 60 rather than the normal value of about
109.5 when the carbon makes four single bonds.
The overlap between the atomic orbitals in forming the carbon-carbon
bonds is less good than it is normally, and there is considerable repulsion
between the bonding pairs. The system becomes more stable if the ring
is broken.

Where would you like to go now?


To the alkanes menu . . .
To the menu of other organic compounds . . .

http://www.chemguide.co.uk/organicprops/alkanes/halogenation.html (5 of 6)30/12/2004 11:39:51

alkanes and cycloalkanes with chlorine or bromine

To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/alkanes/halogenation.html (6 of 6)30/12/2004 11:39:51

free radical substitution in the methane and bromine reaction

THE REACTION BETWEEN METHANE AND


BROMINE

A Free Radical Substitution Reaction


This page gives you the facts and a simple, uncluttered mechanism for
the free radical substitution reaction between methane and bromine. If
you want the mechanism explained to you in detail, there is a link at the
bottom of the page.
The facts
This reaction between methane and bromine happens in the presence of
ultraviolet light - typically sunlight. This is a good example of a
photochemical reaction - a reaction brought about by light.

Note: These reactions are sometimes described as


examples of photocatalysis - reactions catalysed by light. It
is better to use the term "photochemical" and keep the keep
the word "catalysis" for reactions speeded up by actual
substances rather than light.

CH4 + Br2

CH3Br + HBr

The organic product is bromomethane.


One of the hydrogen atoms in the methane has been replaced by a
bromine atom, so this is a substitution reaction. However, the reaction
doesn't stop there, and all the hydrogens in the methane can in turn be
replaced by bromine atoms. Multiple substitution is dealt with on a
separate page, and you will find a link to that at the bottom of this page.

http://www.chemguide.co.uk/mechanisms/freerad/ch4andbr2.html (1 of 4)30/12/2004 11:39:53

free radical substitution in the methane and bromine reaction

Warning! Check your syllabus at this point. If your syllabus


wants you to know about the free radical substitution reaction
between methane and chlorine as well as this one, don't
waste time trying to learn both mechanisms.
The two mechanisms are identical. You just need to learn
one of them. If you are asked for the other one, all you need
to do is to write bromine, say, instead of chlorine.
In writing the bromine mechanisms on these pages, that's
exactly what I've done! If you read both chlorine and bromine
versions, you'll find them boringly repetitive!

The mechanism
The mechanism involves a chain reaction. During a chain reaction, for
every reactive species you start off with, a new one is generated at the
end - and this keeps the process going.

Species: a useful word which is used in chemistry to mean


any sort of particle you want it to mean. It covers molecules,
ions, atoms, or (in this case) free radicals.

The over-all process is known as free radical substitution, or as a free


radical chain reaction.

Note: If you aren't sure about the words free radical or


substitution, read the page What is free radical substitution?
Use the BACK button on your browser to return quickly to this
page.

http://www.chemguide.co.uk/mechanisms/freerad/ch4andbr2.html (2 of 4)30/12/2004 11:39:53

free radical substitution in the methane and bromine reaction

Chain initiation
The chain is initiated (started) by UV light breaking a bromine molecule
into free radicals.
Br2

2Br

Chain propagation reactions


These are the reactions which keep the chain going.
CH4 + Br
CH3

CH3

+ Br2

+ HBr

CH3Br + Br

Chain termination reactions


These are reactions which remove free radicals from the system without
replacing them by new ones.
2Br
CH3
Br

Br2
+

CH3 +
CH3

CH3Br
CH3CH3

Where would you like to go now?


Help! Talk me through this mechanism . . .
Look at multiple substitution in this reaction . . .
Look at why side reactions happen in this reaction . . .
To menu of free radical reactions. . .
To menu of other types of mechanism. . .

http://www.chemguide.co.uk/mechanisms/freerad/ch4andbr2.html (3 of 4)30/12/2004 11:39:53

free radical substitution in the methane and bromine reaction

To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/freerad/ch4andbr2.html (4 of 4)30/12/2004 11:39:53

Explaining the methane - bromine free radical substitution mechanism

EXPLAINING THE REACTION BETWEEN


METHANE AND BROMINE

A Free Radical Substitution Reaction


This page guides you through the mechanism for the substitution of one
of the hydrogen atoms in methane by one bromine atom. Multiple
substitution is covered separately, and you will find a link at the bottom of
the page.
We are going to talk through this mechanism in a very detailed way so
that you get a feel for what is going on. You couldn't possibly do the
same thing in an exam. At the bottom of the page, you will find the
condensed down version corresponding to the sort of answer you would
produce in an exam.
The role of the UV light
The ultraviolet light is simply a source of energy, and is being used to
break bonds. In fact, the energies in UV are exactly right to break the
bonds in bromine molecules to produce bromine atoms.

Note: Only the outer electrons of the bromine are shown.


Notice also that it is quite acceptable to use a simple view of
atomic structure. There is no point in using a complicated
model of the atom if a simple one will do the job.

http://www.chemguide.co.uk/mechanisms/freerad/ch4andbr2tt.html (1 of 6)30/12/2004 11:39:57

Explaining the methane - bromine free radical substitution mechanism

Because we want to stress the fact that the bromine atoms have single
unpaired electrons, then we call them bromine free radicals - or more
usually just bromine radicals.
To show that a species (either an atom or a group of atoms) is a free
radical, the symbol is written with a dot attached to show the unpaired
electron. The splitting of the bromine molecule would be shown as:
Br2

2Br

Free radicals are formed if a bond splits evenly - each atom getting one
of the two electrons. The name given to this is homolytic fission.
What happens to the bromine radicals?
There's nothing magic about reaction mechanisms. Reactions happen
because things hit each other. If the conditions are right something useful
might happen. In this case, you need to think about what the bromine
radicals are likely to hit, and what could happen as a result of that
collision.
At the moment the mixture contains

lots of methane molecules


lots of bromine molecules (only a few will have been fractured by
the UV light)
a few bromine radicals

Let's start with the unproductive collisions.


The least likely collision is between two bromine radicals. There aren't
very many of them in the mixture and so the chances of them hitting
each other are relatively small. If they do collide, they will combine to
form a bromine molecule. That's worse than useless because it removes
the active free radicals from the system.
2Br

Br2

http://www.chemguide.co.uk/mechanisms/freerad/ch4andbr2tt.html (2 of 6)30/12/2004 11:39:57

Explaining the methane - bromine free radical substitution mechanism

A bromine radical could also hit a bromine molecule. If this happens


there could possibly be an exchange of bromine atoms, but nothing new
would be formed. It is just a wasted collision.
Br + Br-Br
Br

Br-Br +

Note: There is no difference between the bromine atoms


shown in bold type or ordinary type. They are shown
differently so that the exchange is made clear.

The productive collision happens if a bromine radical hits a methane


molecule.

The bromine radical removes a hydrogen atom from the methane. That
hydrogen atom only needs to bring one electron with it to form a new
bond to the bromine, and so one electron is left behind on the carbon
atom. A new free radical is formed - this time a methyl radical, CH3 .
CH4 + Br

CH3

+ HBr

What happens to the methyl radicals?


It depends what they collide with. There are three interesting collisions
which need to be explored. Two of these involve a set-back to the
reaction, and only one is useful.
Luckily, the two unhelpful collisions don't happen very often, because
they involve collisions between two free radicals - and there won't be
many of these present in the mixture at any one time.

http://www.chemguide.co.uk/mechanisms/freerad/ch4andbr2tt.html (3 of 6)30/12/2004 11:39:57

Explaining the methane - bromine free radical substitution mechanism

CH3
Br

+
CH3Br

CH3 +
CH3

CH3CH3

Even though the first reaction seems to produce what you want, the
problem with both of these reactions is that they use up the free radicals
in the system - we'll come back to that problem shortly. The second
reaction, of course, also introduces an impurity into the mixture.
So what is the useful collision? If a methyl radical hits a bromine
molecule (something that's quite likely to occur), the following change
can happen:
CH3

+ Br2

CH3Br + Br

The methyl radical takes one of the bromine atoms to form


bromomethane (which is what we want to make), but in the process
generates another bromine radical. This new bromine radical can now go
through the whole sequence again, and at the end will produce yet
another bromine radical - and so on and so on.
The process is described as a free radical chain reaction. The chain
continues because for every bromine radical that goes in at the
beginning, a new one is generated at the end.
Chain termination
Does this mean that one tiny burst of UV light, splitting one bromine
molecule into two free radicals, is enough to convert a whole reactionsworth of methane and bromine into bromomethane and HBr?
Sadly, no! As we've seen, there are collisions which result in the removal
of free radicals without producing any new ones. These radicals can only
be replaced by starting the process all over again with a new burst of
light energy. In practice, then, the chains propagate many thousands of
times, but eventually any chain will be brought to an end by one of these
chain termination processes.

http://www.chemguide.co.uk/mechanisms/freerad/ch4andbr2tt.html (4 of 6)30/12/2004 11:39:57

Explaining the methane - bromine free radical substitution mechanism

Simplifying all this for exam purposes:


The over-all process is known as free radical substitution, or as a free
radical chain reaction.
Chain initiation
The chain is initiated (started) by UV light breaking a bromine molecule
into free radicals.
Br2

2Br

Chain propagation reactions


These are the reactions which keep the chain going.
CH4 + Br
CH3

CH3

+ Br2

+ HBr

CH3Br + Br

Chain termination reactions


These are reactions which remove free radicals from the system without
replacing them by new ones.
2Br
CH3
Br

Br2
+

CH3 +
CH3

CH3Br
CH3CH3

Where would you like to go now?


Look at multiple substitution in this reaction . . .
Look at why side reactions happen in this reaction . . .
To menu of free radical reactions. . .

http://www.chemguide.co.uk/mechanisms/freerad/ch4andbr2tt.html (5 of 6)30/12/2004 11:39:57

Explaining the methane - bromine free radical substitution mechanism

To menu of other types of mechanism. . .


To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/freerad/ch4andbr2tt.html (6 of 6)30/12/2004 11:39:57

Multiple substitution in the methane and bromine reaction

MULTIPLE SUBSTITUTION IN THE METHANE


AND BROMINE REACTION

Warning! We are just about to muddy the water quite


considerably! Don't go on until you are sure that you
understand the mechanism for the production of
bromomethane - and are confident that you could write it in
an exam. If you aren't sure about it, go back to that reaction
and look at it again.
It would be worth checking your syllabus and past exam
papers to see if you need to know about these further
substitution reactions.

The facts
When a mixture of methane and bromine is exposed to ultraviolet light typically sunlight - a substitution reaction occurs and the organic product
is bromomethane.
CH4 + Br2

CH3Br + HBr

However, the reaction doesn't stop there, and all the hydrogens in the
methane can in turn be replaced by bromine atoms. That means that you
could get any of bromomethane, dibromomethane, tribromomethane or
tetrabromomethane.
CH4 + Br2

CH3Br + HBr

CH3Br + Br2

CH2Br2 + HBr

CH2Br2 + Br2

CHBr3 + HBr

CHBr3 + Br2

CBr4 + HBr

You might think that you could control which product you got by the
proportions of methane and bromine you used, but it isn't as simple as
http://www.chemguide.co.uk/mechanisms/freerad/multisubbr.html (1 of 5)30/12/2004 11:40:00

Multiple substitution in the methane and bromine reaction

that. If you use enough bromine you will eventually get CBr4, but any
other proportions will always lead to a mixture of products.

The mechanisms
The formation of multiple substitution products like di-, tri- and
tetrabromomethane can be explained in just the same sort of way as the
formation of the original bromomethane. You just have to look at the
likely collisions as the reaction progresses.
Making dibromomethane
You will remember that the over-all equation for the first stage of the
reaction is
CH4 + Br2

CH3Br + HBr

As the reaction proceeds, the methane is getting used up and


bromomethane is taking its place. That means that the argument about
what a bromine radical is likely to hit changes during the course of the
reaction. As time goes by there is an increasing chance of it hitting a
bromomethane molecule rather than a methane molecule.
When that happens, the bromine radical can take a hydrogen from the
bromomethane just as well as it could from a methane. In this new case:
CH3Br + Br

CH2Br + HBr

Notice: The dot representing the electron has been moved


against the carbon which is the atom with the unpaired
electron. It would be potentially confusing to leave it next to
the bromine.

http://www.chemguide.co.uk/mechanisms/freerad/multisubbr.html (2 of 5)30/12/2004 11:40:00

Multiple substitution in the methane and bromine reaction

The bromomethyl radical formed can then interact with a bromine


molecule in a new propagation step . . .
CH2Br + Br2

CH2Br2 + Br

. . . and so dibromomethane is formed and a bromine radical regenerated.


These propagation steps continue until the chain is terminated by any
two radicals colliding and combining together.
Making tri- and tetrabromomethane
Obviously, as time goes on, there is an increasing chance of the
dibromomethane being hit by a bromine radical - producing these
propagation steps giving tribromomethane:
CH2Br2 + Br
CHBr2 + Br2

CHBr2 + HBr
CHBr3 + Br

Care! Don't just skip lightly over these equations. Look


carefully at each one so that you understand what is
happening, and can relate it to what has gone before. Talk
through the equations with yourself.
For example: "A bromine radical hits the dibromomethane
molecule and steals a hydrogen. That leaves a new radical (I
don't know what it's called, but that doesn't really matter, as
long as I can work out its formula if I have to!), which then
bumps into a bromine molecule - etc, etc."
Doing this helps you to focus properly on the equations. If
you just read them quickly, you'll have forgotten all about
them again in 15 seconds!

http://www.chemguide.co.uk/mechanisms/freerad/multisubbr.html (3 of 5)30/12/2004 11:40:00

Multiple substitution in the methane and bromine reaction

As the amount of tribromomethane builds up, then you will get these
steps giving tetrabromomethane:
CHBr3 + Br
CBr3 + Br2

CBr3 + HBr
CBr4 + Br

This is why you will always get a mixture of products whatever the
reaction proportions of methane and bromine you use. The whole
process is simply governed by chance. Having produced some
bromomethane there is no way that you can prevent it from being hit by
bromine radicals, and similarly for dibromomethane and
tribromomethane.
Trying to produce mainly one product
If you wanted tetrabromomethane, you could of course get it by using a
large excess of bromine, so that eventually all the hydrogens would be
replaced.
If you wanted mainly bromomethane, you could favour this by using a
huge excess of methane so that the chances were always greater of a
bromine radical hitting a methane rather than anything else - but even so,
you would still get some mixture of products.
There is no obvious way of getting mainly dibromomethane or
tribromomethane.

Where would you like to go now?


Look at single substitution again . . .
Look at why side reactions happen in this reaction . . .
To menu of free radical reactions. . .
To menu of other types of mechanism. . .

http://www.chemguide.co.uk/mechanisms/freerad/multisubbr.html (4 of 5)30/12/2004 11:40:00

Multiple substitution in the methane and bromine reaction

To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/freerad/multisubbr.html (5 of 5)30/12/2004 11:40:00

Side reactions in the methane and bromine reaction

SIDE REACTIONS IN THE METHANE AND


BROMINE REACTION

You may remember that one of the chain termination steps produces
ethane, CH3CH3.
CH3 +
CH3

CH3CH3

If bromine radicals hit that, you are going to get bromoethane and
dibromoethane and so on - and in the course of those reactions you will
get ethyl radicals which could themselves become involved in chain
termination steps leading to propane (from methyl radical hitting ethyl
radical) or butane (from two ethyl radicals combining), which could then
start to undergo substitution - and on and on!
To be honest, all of these side products are going to be present in very
small amounts because the reaction producing ethane won't, by chance,
happen very often, but it nicely illustrates a typical organic chemistry
problem - when you do a reaction in the lab to produce an organic
chemical, a high proportion of your time is spent in purifying the product
from all the side reactions that have gone on!

Where would you like to go now?


To basic facts and mechanism for this reaction . . .
Look at multiple substitution in this reaction . . .
To menu of free radical reactions. . .
To menu of other types of mechanism. . .
To Main Menu . . .

http://www.chemguide.co.uk/mechanisms/freerad/sidereactbr.html (1 of 2)30/12/2004 11:40:02

Side reactions in the methane and bromine reaction

Jim Clark 2000

http://www.chemguide.co.uk/mechanisms/freerad/sidereactbr.html (2 of 2)30/12/2004 11:40:02

cracking alkanes - thermal and catalytic

CRACKING ALKANES

This page describes what cracking is, and the differences between
catalytic cracking and thermal cracking used in the petrochemical
industry.

Cracking
What is cracking?
Cracking is the name given to breaking up large hydrocarbon molecules
into smaller and more useful bits. This is achieved by using high
pressures and temperatures without a catalyst, or lower temperatures
and pressures in the presence of a catalyst.
The source of the large hydrocarbon molecules is often the naphtha
fraction or the gas oil fraction from the fractional distillation of crude oil
(petroleum). These fractions are obtained from the distillation process as
liquids, but are re-vaporised before cracking.
There isn't any single unique reaction happening in the cracker. The
hydrocarbon molecules are broken up in a fairly random way to produce
mixtures of smaller hydrocarbons, some of which have carbon-carbon
double bonds. One possible reaction involving the hydrocarbon C15H32
might be:

Or, showing more clearly what happens to the various atoms and bonds:

http://www.chemguide.co.uk/organicprops/alkanes/cracking.html (1 of 5)30/12/2004 11:40:08

cracking alkanes - thermal and catalytic

This is only one way in which this particular molecule might break up.
The ethene and propene are important materials for making plastics or
producing other organic chemicals. The octane is one of the molecules
found in petrol (gasoline).

Catalytic cracking
Modern cracking uses zeolites as the catalyst. These are complex
aluminosilicates, and are large lattices of aluminium, silicon and oxygen
atoms carrying a negative charge. They are, of course, associated with
positive ions such as sodium ions. You may have come across a zeolite
if you know about ion exchange resins used in water softeners.
The alkane is brought into contact with the catalyst at a temperature of
about 500C and moderately low pressures.
The zeolites used in catalytic cracking are chosen to give high
percentages of hydrocarbons with between 5 and 10 carbon atoms particularly useful for petrol (gasoline). It also produces high proportions
of branched alkanes and aromatic hydrocarbons like benzene.
For UK A level purposes, you aren't expected to know how the catalyst
works, but you may be expected to know that it involves an ionic
intermediate.

http://www.chemguide.co.uk/organicprops/alkanes/cracking.html (2 of 5)30/12/2004 11:40:08

cracking alkanes - thermal and catalytic

Note: You should check your syllabus to find out exactly


what you need to know. If you are a UK A level student and
haven't got one, follow this link.
Use the BACK button on your browser to return quickly to this
page.

The zeolite catalyst has sites which can remove a hydrogen from an
alkane together with the two electrons which bound it to the carbon. That
leaves the carbon atom with a positive charge. Ions like this are called
carbonium ions (or carbocations). Reorganisation of these leads to the
various products of the reaction.

Note: If you are interested in other examples of catalysis in


the petrochemical industry, you should follow this link. It will
lead you to information on reforming and isomerisation (as
well as a repeat of what you have just read about catalytic
cracking).
Use the BACK button on your browser if you want to return
quickly to this page.

http://www.chemguide.co.uk/organicprops/alkanes/cracking.html (3 of 5)30/12/2004 11:40:08

cracking alkanes - thermal and catalytic

Thermal cracking
In thermal cracking, high temperatures (typically in the range of 450C to
750C) and pressures (up to about 70 atmospheres) are used to break
the large hydrocarbons into smaller ones. Thermal cracking gives
mixtures of products containing high proportions of hydrocarbons with
double bonds - alkenes.

Warning! This is a gross oversimplification, and is written to


satisfy the needs of one of the UK A level Exam Boards
(AQA). In fact, there are several versions of thermal cracking
designed to produce different mixtures of products. These
use completely different sets of conditions.
If you need to know about thermal cracking in detail, a
Google search on thermal cracking will throw up lots of
useful leads. Be careful to go to industry (or similarly reliable)
sources. You will find a Google search box at the bottom of
the Main Menu (link below). Remember to search the whole
web rather than Chemguide otherwise you will just end up
back here again!

Thermal cracking doesn't go via ionic intermediates like catalytic


cracking. Instead, carbon-carbon bonds are broken so that each carbon
atom ends up with a single electron. In other words, free radicals are
formed.

http://www.chemguide.co.uk/organicprops/alkanes/cracking.html (4 of 5)30/12/2004 11:40:08

cracking alkanes - thermal and catalytic

Reactions of the free radicals lead to the various products.

Where would you like to go now?


To the alkanes menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/alkanes/cracking.html (5 of 5)30/12/2004 11:40:08

making aldehydes and ketones

MAKING ALDEHYDES AND KETONES

This page explains how aldehydes and ketones are made in the lab by
the oxidation of primary and secondary alcohols.

Oxidising alcohols to make aldehydes and ketones


General
The oxidising agent used in these reactions is normally a solution of
sodium or potassium dichromate(VI) acidified with dilute sulphuric acid. If
oxidation occurs, the orange solution containing the dichromate(VI) ions
is reduced to a green solution containing chromium(III) ions.
The net effect is that an oxygen atom from the oxidising agent removes a
hydrogen from the -OH group of the alcohol and one from the carbon to
which it is attached.

[O] is often used to represent oxygen coming from an oxidising agent.


R and R' are alkyl groups or hydrogen. They could also be groups
containing a benzene ring, but I'm ignoring these to keep things simple.

http://www.chemguide.co.uk/organicprops/carbonyls/preparation.html (1 of 5)30/12/2004 11:40:16

making aldehydes and ketones

If at least one of these groups is a hydrogen atom, then you will get an
aldehyde. If they are both alkyl groups then you get a ketone.
If you now think about where they are coming from, you will get an
aldehyde if your starting molecule looks like this:

In other words, if you start from a primary alcohol, you will get an
aldehyde.
You will get a ketone if your starting molecule looks like this:

. . . where R and R' are both alkyl groups.


Secondary alcohols oxidise to give ketones.

Making aldehydes
Aldehydes are made by oxidising primary alcohols. There is, however, a
problem.
The aldehyde produced can be oxidised further to a carboxylic acid by
the acidified potassium dichromate(VI) solution used as the oxidising
agent. In order to stop at the aldehyde, you have to prevent this from
happening.

http://www.chemguide.co.uk/organicprops/carbonyls/preparation.html (2 of 5)30/12/2004 11:40:16

making aldehydes and ketones

Note: This further oxidation is explained in more detail on


the page about oxidation of alcohols.
If you choose to follow this link (not important for the
purposes of the present page), use the BACK button on your
browser to return to this page.

To stop the oxidation at the aldehyde, you . . .

use an excess of the alcohol. That means that there isn't enough
oxidising agent present to carry out the second stage and oxidise
the aldehyde formed to a carboxylic acid.
distil off the aldehyde as soon as it forms. Removing the aldehyde
as soon as it is formed means that it doesn't stay in the mixture to
be oxidised further.

If you used ethanol as a typical primary alcohol, you would produce the
aldehyde ethanal, CH3CHO.
The full equation for this reaction is fairly complicated, and you need to
understand about electron-half-equations in order to work it out.

http://www.chemguide.co.uk/organicprops/carbonyls/preparation.html (3 of 5)30/12/2004 11:40:16

making aldehydes and ketones

Note: If you should know about electron-half-equations but


aren't happy about them, you could follow this link. This
particular equation isn't worked out on that page, but there
should be enough information there to enable you to do it if
you wanted to.
If you choose to follow this link, use the BACK button on your
browser to return to this page.

In organic chemistry, simplified versions are often used which


concentrate on what is happening to the organic substances. To do that,
oxygen from an oxidising agent is represented as [O]. That would
produce the much simpler equation:

Secondary alcohols
Secondary alcohols are oxidised to ketones. There is no further reaction
which might complicate things. For example, if you heat the secondary
alcohol propan-2-ol with sodium or potassium dichromate(VI) solution
acidified with dilute sulphuric acid, you get propanone formed.
Playing around with the reaction conditions makes no difference
whatsoever to the product.
Using the simple version of the equation:

http://www.chemguide.co.uk/organicprops/carbonyls/preparation.html (4 of 5)30/12/2004 11:40:16

making aldehydes and ketones

Where would you like to go now?


To the aldehydes and ketones menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/organicprops/carbonyls/preparation.html (5 of 5)30/12/2004 11:40:16

aldehydes and ketones with grignard reagents

REACTION OF ALDEHYDES AND KETONES


WITH GRIGNARD REAGENTS

This page looks at the reaction of aldehydes and ketones with Grignard
reagents to produce potentially quite complicated alcohols. It is mainly a
duplication of the information on these same reactions from a page on
Grignard reagents in the section on properties of halogenoalkanes.

Note: If you want to read more about these and other


reactions of Grignard reagents you might like to follow this link.
Use the BACK button on your browser if you want to return to
this page.

What are Grignard reagents?


A Grignard reagent has a formula RMgX where X is a halogen, and R is
an alkyl or aryl (based on a benzene ring) group. For the purposes of this
page, we shall take R to be an alkyl group.
A typical Grignard reagent might be CH3CH2MgBr.
The preparation of a Grignard reagent
Grignard reagents are made by adding the halogenoalkane to small bits of
magnesium in a flask containing ethoxyethane (commonly called diethyl
ether or just "ether"). The flask is fitted with a reflux condenser, and the
mixture is warmed over a water bath for 20 - 30 minutes.

http://www.chemguide.co.uk/organicprops/carbonyls/grignard.html (1 of 6)30/12/2004 11:40:24

aldehydes and ketones with grignard reagents

Everything must be perfectly dry because Grignard reagents react with


water.

Warning! Ethoxyethane (ether) is very dangerous to work


with. It is an anaesthetic, and is extremely inflammable. Under
no circumstances should you try to carry out this reaction
without properly qualified guidance.

Any reactions using the Grignard reagent are carried out with the mixture
produced from this reaction. You can't separate it out in any way.

Reactions of Grignard reagents with aldehydes and ketones


These are reactions of the carbon-oxygen double bond, and so aldehydes
and ketones react in exactly the same way - all that changes are the
groups that happen to be attached to the carbon-oxygen double bond.
It is much easier to understand what is going on by looking closely at the
general case (using "R" groups rather than specific groups) - and then
slotting in the various real groups as and when you need to. The "R"
groups can be either hydrogen or alkyl in any combination.
In the first stage, the Grignard reagent adds across the carbon-oxygen
double bond:

Dilute acid is then added to this to hydrolyse it.

http://www.chemguide.co.uk/organicprops/carbonyls/grignard.html (2 of 6)30/12/2004 11:40:24

aldehydes and ketones with grignard reagents

Note: Almost all sources quote the formation of a basic halide


such as Mg(OH)Br as the other product of the reaction. That's
actually misleading because these compounds react with
dilute acids. What you end up with would be a mixture of
ordinary hydrated magnesium ions, halide ions and sulphate
or chloride ions - depending on which dilute acid you added.
What you need to learn about this depends on what your
examiners want. The only way to find that out is to look at old
exam papers and mark schemes. If you are a UK A level
student and haven't got copies of these, find out how to get
hold of them by going to the syllabuses page to find your
Exam Board's web address.

An alcohol is formed. One of the key uses of Grignard reagents is the


ability to make complicated alcohols easily.
What sort of alcohol you get depends on the carbonyl compound you
started with - in other words, what R and R' are.

The reaction between Grignard reagents and methanal


In methanal, both R groups are hydrogen. Methanal is the simplest
possible aldehyde.

Assuming that you are starting with CH3CH2MgBr and using the general
equation above, the alcohol you get always has the form:

http://www.chemguide.co.uk/organicprops/carbonyls/grignard.html (3 of 6)30/12/2004 11:40:24

aldehydes and ketones with grignard reagents

Since both R groups are hydrogen atoms, the final product will be:

A primary alcohol is formed. A primary alcohol has only one alkyl group
attached to the carbon atom with the -OH group on it.
You could obviously get a different primary alcohol if you started from a
different Grignard reagent.

The reaction between Grignard reagents and other aldehydes


The next biggest aldehyde is ethanal. One of the R groups is hydrogen
and the other CH3.

Again, think about how that relates to the general case. The alcohol
formed is:

So this time the final product has one CH3 group and one hydrogen
attached:
http://www.chemguide.co.uk/organicprops/carbonyls/grignard.html (4 of 6)30/12/2004 11:40:24

aldehydes and ketones with grignard reagents

A secondary alcohol has two alkyl groups (the same or different) attached
to the carbon with the -OH group on it.
You could change the nature of the final secondary alcohol by either:

changing the nature of the Grignard reagent - which would change


the CH3CH2 group into some other alkyl group;
changing the nature of the aldehyde - which would change the CH3
group into some other alkyl group.

The reaction between Grignard reagents and ketones


Ketones have two alkyl groups attached to the carbon-oxygen double
bond. The simplest one is propanone.

This time when you replace the R groups in the general formula for the
alcohol produced you get a tertiary alcohol.

A tertiary alcohol has three alkyl groups attached to the carbon with the http://www.chemguide.co.uk/organicprops/carbonyls/grignard.html (5 of 6)30/12/2004 11:40:24

aldehydes and ketones with grignard reagents

OH attached. The alkyl groups can be any combination of same or


different.
You could ring the changes on the product by

changing the nature of the Grignard reagent - which would change


the CH3CH2 group into some other alkyl group;
changing the nature of the ketone - which would change the CH3
groups into whatever other alkyl groups you choose to have in the
original ketone.

Where would you like to go now?


To the aldehydes and ketones menu . . .
To the menu of other organic compounds . . .
To Main Menu . . .

Jim Clark 2004

http://www.chemguide.co.uk/organicprops/carbonyls/grignard.html (6 of 6)30/12/2004 11:40:24

successive ionisation energies (second, third, etc)

SUCCESSIVE IONISATION ENERGIES

This page explains what second, third, (etc) ionisation energy means,
and then looks at patterns in successive ionisation energies for selected
elements. It assumes that you understand about first ionisation energy.

Important! If you have come straight to this page via a


search engine, you should read the page on first ionisation
energy before you go any further.

Defining second ionisation energy


Second ionisation energy is defined by the equation:

It is the energy needed to remove a second electron from each ion in 1


mole of gaseous 1+ ions to give gaseous 2+ ions.
More ionisation energies
You can then have as many successive ionisation energies as there are
electrons in the original atom.
The first four ionisation energies of aluminium, for example, are given by
1st I.E. = 577 kJ mol-1
2nd I.E. = 1820 kJ mol-1
3rd I.E. = 2740 kJ mol-1
4th I.E. = 11600 kJ mol-1
In order to form an Al3+(g) ion from Al(g) you would have to supply:
http://www.chemguide.co.uk/atoms/properties/moreies.html (1 of 7)30/12/2004 11:41:24

successive ionisation energies (second, third, etc)

577 + 1820 + 2740 = 5137 kJ mol-1


That's a lot of energy. Why, then, does aluminium form Al3+ ions?
It can only form them if it can get that energy back from somewhere, and
whether that's feasible depends on what it is reacting with.
For example, if aluminium reacts with fluorine or oxygen, it can recover
that energy in various changes involving the fluorine or oxygen - and so
aluminium fluoride or aluminium oxide contain Al3+ ions.
If it reacts with chlorine, it can't recover sufficient energy, and so solid
anhydrous aluminium chloride isn't actually ionic - instead, it forms
covalent bonds.
Why doesn't aluminium form an Al4+ ion? The fourth ionisation energy is
huge compared with the first three, and there is nothing that aluminium
can react with which would enable it to recover that amount of extra
energy.
Why do successive ionisation energies get larger?
Once you have removed the first electron you are left with a positive ion.
Trying to remove a negative electron from a positive ion is going to be
more difficult than removing it from an atom. Removing an electron from
a 2+ or 3+ (etc) ion is going to be progressively more difficult.
Why is the fourth ionisation energy of aluminium so large?
The electronic structure of aluminium is 1s22s22p63s23px1. The first three
electrons to be removed are the three electrons in the 3p and 3s orbitals.
Once they've gone, the fourth electron is removed from the 2p level much closer to the nucleus, and only screened by the 1s2 (and to some
extent the 2s2) electrons.

Using ionisation energies to work out which group an element is in


This big jump between two successive ionisation energies is typical of
http://www.chemguide.co.uk/atoms/properties/moreies.html (2 of 7)30/12/2004 11:41:24

successive ionisation energies (second, third, etc)

suddenly breaking in to an inner level. You can use this to work out
which group of the Periodic Table an element is in from its successive
ionisation energies.
Magnesium (1s22s22p63s2) is in group 2 of the Periodic Table and has
successive ionisation energies:

Here the big jump occurs after the second ionisation energy. It means
that there are 2 electrons which are relatively easy to remove (the 3s2
electrons), while the third one is much more difficult (because it comes
from an inner level - closer to the nucleus and with less screening).
Silicon (1s22s22p63s23px13py1) is in group 4 of the Periodic Table and
has successive ionisation energies:

Here the big jump comes after the fourth electron has been removed.
The first 4 electrons are coming from the 3-level orbitals; the fifth from
the 2-level.
The lesson from all this:
Count the easy electrons - those up to (but not including) the big jump.
That is the same as the group number.

http://www.chemguide.co.uk/atoms/properties/moreies.html (3 of 7)30/12/2004 11:41:24

successive ionisation energies (second, third, etc)

Another example:
Decide which group an atom is in if it has successive ionisation energies:

The ionisation energies are going up one or two thousand at a time for
the first five. Then there is a huge jump of about 15000. There are 5
relatively easy electrons - so the element is in group 5.

Exploring the patterns in more detail


If you plot graphs of successive ionisation energies for a particular
element, you can see the fluctuations in it caused by the different
electrons being removed.
Not only can you see the big jumps in ionisation energy when an electron
comes from an inner level, but you can also see the minor fluctuations
within a level depending on whether the electron is coming from an s or a
p orbital, and even whether it is paired or unpaired in that orbital.
Chlorine has the electronic structure 1s22s22p63s23px23py23pz1.
This graph plots the first eight ionisation energies of chlorine. The green
labels show which electron is being removed for each of the ionisation
energies.

http://www.chemguide.co.uk/atoms/properties/moreies.html (4 of 7)30/12/2004 11:41:24

successive ionisation energies (second, third, etc)

If you put a ruler on the first and second points to establish the trend,
you'll find that the third, fourth and fifth points lie above the value you
would expect. That is because the first two electrons are coming from
pairs in the 3p levels and are therefore rather easier to remove than if
they were unpaired.
Again, if you put a ruler on the 3rd, 4th and 5th points to establish their
trend, you'll find that the 6th and 7th points lie well above the values you
would expect from a continuation of the trend. That is because the 6th
and 7th electrons are coming from the 3s level - slightly closer to the
nucleus and slightly less well screened.
The massive jump as you break into the inner level at the 8th electron is
fairly obvious!

Warning! People sometimes get confused with these graphs


because they forget that they are removing electrons from
the atom. For example, the first point refers to the first
electron being lost - from a 3p orbital. Basically, you start
from the outside of the atom and work in towards the middle.
If you start from the 1s orbital and work outwards, you are
doomed to failure!

http://www.chemguide.co.uk/atoms/properties/moreies.html (5 of 7)30/12/2004 11:41:24

successive ionisation energies (second, third, etc)

To plot any more ionisation energies for chlorine needs a change of


vertical scale. The seventeenth ionisation energy of chlorine is nearly
400,000 kJ mol-1, and the vertical scale has to be squashed to
accommodate this.

This is now a "log graph" - plotted by finding the logarithm of each


ionisation energy (press the "log" button on your calculator). This doesn't
simply squash the vertical scale. It distorts it as well, to such an extent
that the only useful thing the graph now shows is the major jumps where
the next electron to be removed comes from an inner level. The distortion
is so great in the first 8 ionisation energies, for example, that the patterns
shown by the previous graph are completely (and misleadingly)
destroyed.

Where would you like to go now?


To the atomic properties menu . . .
To the atomic structure and bonding menu . . .
To Main Menu . . .

http://www.chemguide.co.uk/atoms/properties/moreies.html (6 of 7)30/12/2004 11:41:24

successive ionisation energies (second, third, etc)

Jim Clark 2000

http://www.chemguide.co.uk/atoms/properties/moreies.html (7 of 7)30/12/2004 11:41:24

Understanding Chemistry - Instrumental Analysis Menu

Understanding Chemistry

INSTRUMENTAL ANALYSIS MENU

The mass spectrometer . . .


Includes an explanation of how the mass spectrometer works and
how it can be used both to find relative atomic masses of elements
and to help to identify organic compounds.
Infra-red spectra . . .
Explains how infra-red spectra can be used to identify particular
groups in an organic compound, or even to identify the whole
compound.
Nuclear magnetic resonance (NMR) spectra . . .
Explains how both low and high resolution NMR spectra can be
used to help to work out the structures of organic compounds.

Go to Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/analysismenu.html30/12/2004 11:41:30

Mass spectrometry menu

Understanding Chemistry

MASS SPECTROMETRY MENU

How the mass spectrometer works . . .


An explanation of how a mass spectrum is produced
The mass spectra of elements . . .
How the mass spectrum of an element can be used to find its
relative atomic mass.

The mass spectra of organic compounds


The formation of fragmentation patterns . . .
What the mass spectra of organic compounds look like, how the
patterns are formed, and the sort of information you can get from
them.
The molecular ion (M+) peak . . .
How the molecular ion peak can be used to find the relative
formula mass of the compound and its molecular formula.
The M+1 peak . . .
How the M+1 peak arises, and its use in finding the number of
carbon atoms a compound contains.
Organic compounds containing halogen atoms . . .

http://www.chemguide.co.uk/analysis/masspecmenu.html (1 of 2)30/12/2004 11:41:31

Mass spectrometry menu

How the M+2 (and possibly M+4) peak arises, and its use in
showing the presence of chlorine or bromine in a compound.

Go to instrumental analysis menu . . .


Go to Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/analysis/masspecmenu.html (2 of 2)30/12/2004 11:41:31

the mass spectrometer - how it works

THE MASS SPECTROMETER

This page describes how a mass spectrum is produced using a mass


spectrometer.

How a mass spectrometer works


The basic principle
If something is moving and you subject it to a sideways force, instead of
moving in a straight line, it will move in a curve - deflected out of its
original path by the sideways force.
Suppose you had a cannonball travelling past you and you wanted to
deflect it as it went by you. All you've got is a jet of water from a hosepipe that you can squirt at it. Frankly, its not going to make a lot of
difference! Because the cannonball is so heavy, it will hardly be deflected
at all from its original course.
But suppose instead, you tried to deflect a table tennis ball travelling at
the same speed as the cannonball using the same jet of water. Because
this ball is so light, you will get a huge deflection.
The amount of deflection you will get for a given sideways force depends
on the mass of the ball. If you knew the speed of the ball and the size of
the force, you could calculate the mass of the ball if you knew what sort
of curved path it was deflected through. The less the deflection, the
heavier the ball.

Note: I'm not suggesting that you personally would have to


do the calculation, although the maths isn't actually very
difficult - certainly no more than A'level standard!

http://www.chemguide.co.uk/analysis/masspec/howitworks.html (1 of 9)30/12/2004 11:41:37

the mass spectrometer - how it works

You can apply exactly the same principle to atomic sized particles.

An outline of what happens in a mass spectrometer


Atoms can be deflected by magnetic fields - provided the atom is first
turned into an ion. Electrically charged particles are affected by a
magnetic field although electrically neutral ones aren't.
The sequence is :
Stage 1: Ionisation
The atom is ionised by knocking one or more electrons off to give a
positive ion. This is true even for things which you would normally expect
to form negative ions (chlorine, for example) or never form ions at all
(argon, for example). Mass spectrometers always work with positive ions.
Stage 2: Acceleration
The ions are accelerated so that they all have the same kinetic energy.
Stage 3: Deflection
The ions are then deflected by a magnetic field according to their
masses. The lighter they are, the more they are deflected.
The amount of deflection also depends on the number of positive
charges on the ion - in other words, on how many electrons were
knocked off in the first stage. The more the ion is charged, the more it
gets deflected.
Stage 4: Detection
The beam of ions passing through the machine is detected electrically.

A full diagram of a mass spectrometer

http://www.chemguide.co.uk/analysis/masspec/howitworks.html (2 of 9)30/12/2004 11:41:37

the mass spectrometer - how it works

Understanding what's going on


The need for a vacuum
It's important that the ions produced in the ionisation chamber have a
free run through the machine without hitting air molecules.
Ionisation

http://www.chemguide.co.uk/analysis/masspec/howitworks.html (3 of 9)30/12/2004 11:41:37

the mass spectrometer - how it works

The vaporised sample passes into the ionisation chamber. The


electrically heated metal coil gives off electrons which are attracted to the
electron trap which is a positively charged plate.
The particles in the sample (atoms or molecules) are therefore
bombarded with a stream of electrons, and some of the collisions are
energetic enough to knock one or more electrons out of the sample
particles to make positive ions.
Most of the positive ions formed will carry a charge of +1 because it is
much more difficult to remove further electrons from an already positive
ion.
These positive ions are persuaded out into the rest of the machine by the
ion repeller which is another metal plate carrying a slight positive charge.

Note: As you will see in a moment, the whole ionisation


chamber is held at a positive voltage of about 10,000 volts.
Where we are talking about the two plates having positive
charges, these charges are in addition to that 10,000 volts.

http://www.chemguide.co.uk/analysis/masspec/howitworks.html (4 of 9)30/12/2004 11:41:37

the mass spectrometer - how it works

Acceleration

The positive ions are repelled away from the very positive ionisation
chamber and pass through three slits, the final one of which is at 0 volts.
The middle slit carries some intermediate voltage. All the ions are
accelerated into a finely focused beam.
Deflection

Different ions are deflected by the magnetic field by different amounts.


The amount of deflection depends on:

the mass of the ion. Lighter ions are deflected more than heavier
ones.
the charge on the ion. Ions with 2 (or more) positive charges are
deflected more than ones with only 1 positive charge.

http://www.chemguide.co.uk/analysis/masspec/howitworks.html (5 of 9)30/12/2004 11:41:37

the mass spectrometer - how it works

These two factors are combined into the mass/charge ratio. Mass/
charge ratio is given the symbol m/z (or sometimes m/e).
For example, if an ion had a mass of 28 and a charge of 1+, its mass/
charge ratio would be 28. An ion with a mass of 56 and a charge of 2+
would also have a mass/charge ratio of 28.
In the last diagram, ion stream A is most deflected - it will contain ions
with the smallest mass/charge ratio. Ion stream C is the least deflected it contains ions with the greatest mass/charge ratio.
It makes it simpler to talk about this if we assume that the charge on all
the ions is 1+. Most of the ions passing through the mass spectrometer
will have a charge of 1+, so that the mass/charge ratio will be the same
as the mass of the ion.

Note: You must be aware of the possibility of 2+ (etc) ions,


but the vast majority of A'level questions will give you mass
spectra which only involve 1+ ions. Unless there is some hint
in the question, you can reasonably assume that the ions you
are talking about will have a charge of 1+.

Assuming 1+ ions, stream A has the lightest ions, stream B the next
lightest and stream C the heaviest. Lighter ions are going to be more
deflected than heavy ones.
Detection
Only ion stream B makes it right through the machine to the ion detector.
The other ions collide with the walls where they will pick up electrons and
be neutralised. Eventually, they get removed from the mass
spectrometer by the vacuum pump.

http://www.chemguide.co.uk/analysis/masspec/howitworks.html (6 of 9)30/12/2004 11:41:37

the mass spectrometer - how it works

When an ion hits the metal box, its charge is neutralised by an electron
jumping from the metal on to the ion (right hand diagram). That leaves a
space amongst the electrons in the metal, and the electrons in the wire
shuffle along to fill it.
A flow of electrons in the wire is detected as an electric current which can
be amplified and recorded. The more ions arriving, the greater the
current.
Detecting the other ions
How might the other ions be detected - those in streams A and C which
have been lost in the machine?
Remember that stream A was most deflected - it has the smallest value
of m/z (the lightest ions if the charge is 1+). To bring them on to the
detector, you would need to deflect them less - by using a smaller
magnetic field (a smaller sideways force).
To bring those with a larger m/z value (the heavier ions if the charge is
+1) on to the detector you would have to deflect them more by using a
larger magnetic field.
If you vary the magnetic field, you can bring each ion stream in turn on to
the detector to produce a current which is proportional to the number of
ions arriving. The mass of each ion being detected is related to the size
of the magnetic field used to bring it on to the detector. The machine can
be calibrated to record current (which is a measure of the number of

http://www.chemguide.co.uk/analysis/masspec/howitworks.html (7 of 9)30/12/2004 11:41:37

the mass spectrometer - how it works

ions) against m/z directly. The mass is measured on the 12C scale.

Note: The 12C scale is a scale on which the 12C isotope


weighs exactly 12 units.

What the mass spectrometer output looks like


The output from the chart recorder is usually simplified into a "stick
diagram". This shows the relative current produced by ions of varying
mass/charge ratio.
The stick diagram for molybdenum looks lilke this:

You may find diagrams in which the vertical axis is labelled as either
"relative abundance" or "relative intensity". Whichever is used, it means
the same thing. The vertical scale is related to the current received by
the chart recorder - and so to the number of ions arriving at the detector:
the greater the current, the more abundant the ion.
As you will see from the diagram, the commonest ion has a mass/charge
ratio of 98. Other ions have mass/charge ratios of 92, 94, 95, 96, 97 and
100.
That means that molybdenum consists of 7 different isotopes. Assuming
that the ions all have a charge of 1+, that means that the masses of the 7
http://www.chemguide.co.uk/analysis/masspec/howitworks.html (8 of 9)30/12/2004 11:41:37

the mass spectrometer - how it works

isotopes on the carbon-12 scale are 92, 94, 95, 96, 97, 98 and 100.

Note: If there were also 2+ ions present, you would know


because every one of the lines in the stick diagram would
have another line at exactly half its m/z value (because, for
example, 98/2 = 49). Those lines would be much less tall
than the 1+ ion lines because the chances of forming 2+ ions
are much less than forming 1+ ions.
If you want to go straight on to how you use these mass
spectra to calculate relative atomic masses you can jump
straight to that page by following this link rather than going
via the menus below.

Where would you like to go now?


To the mass spectrometry menu . . .
To the instrumental analysis menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/analysis/masspec/howitworks.html (9 of 9)30/12/2004 11:41:37

the mass spectra of elements

THE MASS SPECTRA OF ELEMENTS

This page looks at the information you can get from the mass spectrum
of an element. It shows how you can find out the masses and relative
abundances of the various isotopes of the element and use that
information to calculate the relative atomic mass of the element.
It also looks at the problems thrown up by elements with diatomic
molecules - like chlorine, Cl2.

The mass spectrum of monatomic elements


Monatomic elements include all those except for things like chlorine, Cl2,
with molecules containing more than one atom.
The mass spectrum for boron

Note: If you need to know how this diagram is obtained, you


should read the page describing how a mass spectrometer
works.

http://www.chemguide.co.uk/analysis/masspec/elements.html (1 of 8)30/12/2004 11:41:41

the mass spectra of elements

The number of isotopes


The two peaks in the mass spectrum shows that there are 2 isotopes of
boron - with relative isotopic masses of 10 and 11 on the 12C scale.

Notes: Isotopes are atoms of the same element (and so


with the same number of protons), but with different masses
due to having different numbers of neutrons.
We are assuming (and shall do all through this page) that all
the ions recorded have a charge of 1+. That means that the
mass/charge ratio (m/z) gives you the mass of the isotope
directly.
The carbon-12 scale is a scale on which the mass of the 12C
isotope weighs exactly 12 units.

The abundance of the isotopes


The relative sizes of the peaks gives you a direct measure of the relative
abundances of the isotopes. The tallest peak is often given an arbitrary
height of 100 - but you may find all sorts of other scales used. It doesn't
matter in the least.
You can find the relative abundances by measuring the lines on the stick
diagram.
In this case, the two isotopes (with their relative abundances) are:
boron-10
boron-11

23
100

Working out the relative atomic mass


The relative atomic mass (RAM) of an element is given the symbol Ar
and is defined as:

http://www.chemguide.co.uk/analysis/masspec/elements.html (2 of 8)30/12/2004 11:41:41

the mass spectra of elements

The relative atomic mass of an element is the weighted average


of the masses of the isotopes relative to 1/12 of the mass of a
carbon-12 atom.
A "weighted average" allows for the fact that there won't be equal
amounts of the various isotopes. The example coming up should make
that clear.
Suppose you had 123 typical atoms of boron. 23 of these would be 10B
and 100 would be 11B.
The total mass of these would be (23 x 10) + (100 x 11) = 1330
The average mass of these 123 atoms would be 1330 / 123 = 10.8 (to 3
significant figures).
10.8 is the relative atomic mass of boron.
Notice the effect of the "weighted" average. A simple average of 10 and
11 is, of course, 10.5. Our answer of 10.8 allows for the fact that there
are a lot more of the heavier isotope of boron - and so the "weighted"
average ought to be closer to that.

The mass spectrum for zirconium

The number of isotopes

http://www.chemguide.co.uk/analysis/masspec/elements.html (3 of 8)30/12/2004 11:41:41

the mass spectra of elements

The 5 peaks in the mass spectrum shows that there are 5 isotopes of
zirconium - with relative isotopic masses of 90, 91, 92, 94 and 96 on the
12C scale.
The abundance of the isotopes
This time, the relative abundances are given as percentages. Again you
can find these relative abundances by measuring the lines on the stick
diagram.
In this case, the 5 isotopes (with their relative percentage abundances)
are:
zirconium-90
zirconium-91
zirconium-92
zirconium-94
zirconium-96

51.5
11.2
17.1
17.4
2.8

Note: You almost certainly wouldn't be able to measure


these peaks to this degree of accuracy, but your examiners
may well give you the data in number form anyway. We'll do
the sum with the more accurate figures.

Working out the relative atomic mass


Suppose you had 100 typical atoms of zirconium. 51.5 of these would be
90Zr, 11.2 would be 91Zr and so on.

Note: If you object to the idea of having 51.5 atoms or 11.2


atoms and so on, just assume you've got 1000 atoms instead
of 100. That way you will have 515 atoms, 112 atoms, etc.
Most people don't get in a sweat over this, and just use the
numbers as they are!

http://www.chemguide.co.uk/analysis/masspec/elements.html (4 of 8)30/12/2004 11:41:41

the mass spectra of elements

The total mass of these 100 typical atoms would be


(51.5 x 90) + (11.2 x 91) + (17.1 x 92) + (17.4 x 94) + (2.8 x 96) = 9131.8
The average mass of these 100 atoms would be 9131.8 / 100 = 91.3 (to
3 significant figures).
91.3 is the relative atomic mass of zirconium.

Note: If you want further examples of calculating relative


atomic masses from mass spectra, you might like to refer to
my book, Calculations in A level Chemistry.

The mass spectrum of chlorine


Chlorine is taken as typical of elements with more than one atom per
molecule. We'll look at its mass spectrum to show the sort of problems
involved.
Chlorine has two isotopes, 35Cl and 37Cl, in the approximate ratio of 3
atoms of 35Cl to 1 atom of 37Cl. You might suppose that the mass
spectrum would look like this:

You would be wrong!


http://www.chemguide.co.uk/analysis/masspec/elements.html (5 of 8)30/12/2004 11:41:41

the mass spectra of elements

The problem is that chlorine consists of molecules, not individual atoms.


When chlorine is passed into the ionisation chamber, an electron is
knocked off the molecule to give a molecular ion, Cl2+. These ions won't
be particularly stable, and some will fall apart to give a chlorine atom and
a Cl+ ion. The term for this is fragmentation.

If the Cl atom formed isn't then ionised in the ionisation chamber, it


simply gets lost in the machine - neither accelerated nor deflected.
The Cl+ ions will pass through the machine and will give lines at 35 and
37, depending on the isotope and you would get exactly the pattern in
the last diagram. The problem is that you will also record lines for the
unfragmented Cl2+ ions.
Think about the possible combinations of chlorine-35 and chlorine-37
atoms in a Cl2+ ion.
Both atoms could be 35Cl, both atoms could be 37Cl, or you could have
one of each sort. That would give you total masses of the Cl2+ ion of:
35 + 35 = 70
35 + 37 = 72
37 + 37 = 74
That means that you would get a set of lines in the m/z = 70 region
looking like this:

http://www.chemguide.co.uk/analysis/masspec/elements.html (6 of 8)30/12/2004 11:41:41

the mass spectra of elements

These lines would be in addition to the lines at 35 and 37.


The relative heights of the 70, 72 and 74 lines are in the ratio 9:6:1. If you
know the right bit of maths, it's very easy to show this. If not, don't worry.
Just remember that the ratio is 9:6:1.
What you can't do is make any predictions about the relative heights of
the lines at 35/37 compared with those at 70/72/74. That depends on
what proportion of the molecular ions break up into fragments. That's
why you've got the chlorine mass spectrum in two separate bits so far.
You must realise that the vertical scale in the diagrams of the two parts
of the spectrum isn't the same.
The overall mass spectrum looks like this:

http://www.chemguide.co.uk/analysis/masspec/elements.html (7 of 8)30/12/2004 11:41:41

the mass spectra of elements

Note: This is based on information from the NIST Chemistry


WebBook. NIST is the US National Institute of Standards and
Technology.

Where would you like to go now?


To the mass spectrometry menu . . .
To the instrumental analysis menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/analysis/masspec/elements.html (8 of 8)30/12/2004 11:41:41

mass spectra - fragmentation patterns

FRAGMENTATION PATTERNS IN THE MASS


SPECTRA OF ORGANIC COMPOUNDS

This page looks at how fragmentation patterns are formed when organic
molecules are fed into a mass spectrometer, and how you can get
information from the mass spectrum.

The origin of fragmentation patterns


The formation of molecular ions
When the vaporised organic sample passes into the ionisation chamber
of a mass spectrometer, it is bombarded by a stream of electrons. These
electrons have a high enough energy to knock an electron off an organic
molecule to form a positive ion. This ion is called the molecular ion - or
sometimes the parent ion.

Note: If you aren't sure about how a mass spectrum is


produced, it might be worth taking a quick look at the page
describing how a mass spectrometer works.

The molecular ion is often given the symbol M+ or


- the dot in this
second version represents the fact that somewhere in the ion there will
be a single unpaired electron. That's one half of what was originally a
pair of electrons - the other half is the electron which was removed in the
ionisation process.
Fragmentation
The molecular ions are energetically unstable, and some of them will
break up into smaller pieces. The simplest case is that a molecular ion
breaks into two parts - one of which is another positive ion, and the other
is an uncharged free radical.
http://www.chemguide.co.uk/analysis/masspec/fragment.html (1 of 13)30/12/2004 11:41:50

mass spectra - fragmentation patterns

Note: A free radical is an atom or group of atoms which


contains a single unpaired electron.
More complicated break-ups are beyond the scope of A'level
syllabuses.

The uncharged free radical won't produce a line on the mass spectrum.
Only charged particles will be accelerated, deflected and detected by the
mass spectrometer. These uncharged particles will simply get lost in the
machine - eventually, they get removed by the vacuum pump.
The ion, X+, will travel through the mass spectrometer just like any other
positive ion - and will produce a line on the stick diagram.
All sorts of fragmentations of the original molecular ion are possible - and
that means that you will get a whole host of lines in the mass spectrum.
For example, the mass spectrum of pentane looks like this:

http://www.chemguide.co.uk/analysis/masspec/fragment.html (2 of 13)30/12/2004 11:41:50

mass spectra - fragmentation patterns

Note: All the mass spectra on this page have been drawn
using data from the Spectral Data Base System for Organic
Compounds (SDBS) at the National Institute of Materials and
Chemical Research in Japan.
They have been simplified by omitting all the minor lines with
peak heights of 2% or less of the base peak (the tallest peak).

It's important to realise that the pattern of lines in the mass spectrum of
an organic compound tells you something quite different from the pattern
of lines in the mass spectrum of an element. With an element, each line
represents a different isotope of that element. With a compound, each
line represents a different fragment produced when the molecular ion
breaks up.

Note: If you are interested in the mass spectra of elements,


you could follow this link.

The molecular ion peak and the base peak


In the stick diagram showing the mass spectrum of pentane, the line
produced by the heaviest ion passing through the machine (at m/z = 72)
is due to the molecular ion.
Note: You have to be a bit careful about this, because in
some cases, the molecular ion is so unstable that every
single one of them splits up, and none gets through the
machine to register in the mass spectrum. You are very
unlikely to come across such a case at A'level.

http://www.chemguide.co.uk/analysis/masspec/fragment.html (3 of 13)30/12/2004 11:41:50

mass spectra - fragmentation patterns

The tallest line in the stick diagram (in this case at m/z = 43) is called the
base peak. This is usually given an arbitrary height of 100, and the
height of everything else is measured relative to this. The base peak is
the tallest peak because it represents the commonest fragment ion to be
formed - either because there are several ways in which it could be
produced during fragmentation of the parent ion, or because it is a
particularly stable ion.

Using fragmentation patterns


This section will ignore the information you can get from the molecular
ion (or ions). That is covered in three other pages which you can get at
via the mass spectrometry menu. You will find a link at the bottom of the
page.
Working out which ion produces which line
This is generally the simplest thing you can be asked to do.
The mass spectrum of pentane
Let's have another look at the mass spectrum for pentane:

What causes the line at m/z = 57?

http://www.chemguide.co.uk/analysis/masspec/fragment.html (4 of 13)30/12/2004 11:41:50

mass spectra - fragmentation patterns

How many carbon atoms are there in this ion? There can't be 5 because
5 x 12 = 60. What about 4? 4 x 12 = 48. That leaves 9 to make up a total
of 57. How about C4H9+ then?
C4H9+ would be [CH3CH2CH2CH2]+, and this would be produced by the
following fragmentation:

The methyl radical produced will simply get lost in the machine.
The line at m/z = 43 can be worked out similarly. If you play around with
the numbers, you will find that this corresponds to a break producing a 3carbon ion:

The line at m/z = 29 is typical of an ethyl ion, [CH3CH2]+:

The other lines in the mass spectrum are more difficult to explain. For
example, lines with m/z values 1 or 2 less than one of the easy lines are
often due to loss of one or more hydrogen atoms during the
fragmentation process. You are very unlikely to have to explain any but
the most obvious cases in an A'level exam.

The mass spectrum of pentan-3-one

http://www.chemguide.co.uk/analysis/masspec/fragment.html (5 of 13)30/12/2004 11:41:50

mass spectra - fragmentation patterns

This time the base peak (the tallest peak - and so the commonest
fragment ion) is at m/z = 57. But this isn't produced by the same ion as
the same m/z value peak in pentane.
If you remember, the m/z = 57 peak in pentane was produced by
[CH3CH2CH2CH2]+. If you look at the structure of pentan-3-one, it's
impossible to get that particular fragment from it.
Work along the molecule mentally chopping bits off until you come up
with something that adds up to 57. With a small amount of patience,
you'll eventually find [CH3CH2CO]+ - which is produced by this
fragmentation:

You would get exactly the same products whichever side of the CO
group you split the molecular ion.
The m/z = 29 peak is produced by the ethyl ion - which once again could
be formed by splitting the molecular ion either side of the CO group.

http://www.chemguide.co.uk/analysis/masspec/fragment.html (6 of 13)30/12/2004 11:41:50

mass spectra - fragmentation patterns

Peak heights and the stability of ions


The more stable an ion is, the more likely it is to form. The more of a
particular sort of ion that's formed, the higher its peak height will be. We'll
look at two common examples of this.
Examples involving carbocations (carbonium ions)

Important! If you don't know what a carbocation (or


carbonium ion) is, or why the various sorts vary in stability,
it's essential that you follow this link before you go on.
Use the BACK button on your browser to return quickly to this
page.

Summarizing the most important conclusion from the page on


carbocations:
Order of stability of carbocations
primary < secondary < tertiary

Note: The symbol "<" means "is less than". So what this is
saying is that primary ions are less stable than secondary
ones which in turn are less stable than tertiary ones.

http://www.chemguide.co.uk/analysis/masspec/fragment.html (7 of 13)30/12/2004 11:41:50

mass spectra - fragmentation patterns

Applying the logic of this to fragmentation patterns, it means that a split


which produces a secondary carbocation is going to be more successful
than one producing a primary one. A split producing a tertiary
carbocation will be more successful still.
Let's look at the mass spectrum of 2-methylbutane. 2-methylbutane is an
isomer of pentane - isomers are molecules with the same molecular
formula, but a different spatial arrangement of the atoms.

Look first at the very strong peak at m/z = 43. This is caused by a
different ion than the corresponding peak in the pentane mass spectrum.
This peak in 2-methylbutane is caused by:

The ion formed is a secondary carbocation - it has two alkyl groups


attached to the carbon with the positive charge. As such, it is relatively
stable.
The peak at m/z = 57 is much taller than the corresponding line in
pentane. Again a secondary carbocation is formed - this time, by:

http://www.chemguide.co.uk/analysis/masspec/fragment.html (8 of 13)30/12/2004 11:41:50

mass spectra - fragmentation patterns

You would get the same ion, of course, if the left-hand CH3 group broke
off instead of the bottom one as we've drawn it.
In these two spectra, this is probably the most dramatic example of the
extra stability of a secondary carbocation.

Examples involving acylium ions, [RCO]+


Ions with the positive charge on the carbon of a carbonyl group, C=O,
are also relatively stable. This is fairly clearly seen in the mass spectra of
ketones like pentan-3-one.

The base peak, at m/z=57, is due to the [CH3CH2CO]+ ion. We've


already discussed the fragmentation that produces this.

http://www.chemguide.co.uk/analysis/masspec/fragment.html (9 of 13)30/12/2004 11:41:50

mass spectra - fragmentation patterns

Note: There are lots of other examples of positive ions with


extra stability and which are produced in large numbers in a
mass spectrometer as a result. Without making this article
even longer than it already is, it's impossible to cover every
possible case.
Check past exam papers to find out whether you are likely to
need to know about other possibilities. If you haven't got past
papers, follow the link on the syllabuses page to find out how
to get hold of them.

Using mass spectra to distinguish between compounds


Suppose you had to suggest a way of distinguishing between pentan-2one and pentan-3-one using their mass spectra.
pentan-2-one

CH3COCH2CH2CH3

pentan-3-one

CH3CH2COCH2CH3

Each of these is likely to split to produce ions with a positive charge on


the CO group.
In the pentan-2-one case, there are two different ions like this:

[CH3CO]+

[COCH2CH2CH3]+

That would give you strong lines at m/z = 43 and 71.


With pentan-3-one, you would only get one ion of this kind:

[CH3CH2CO]+

http://www.chemguide.co.uk/analysis/masspec/fragment.html (10 of 13)30/12/2004 11:41:50

mass spectra - fragmentation patterns

In that case, you would get a strong line at 57.


You don't need to worry about the other lines in the spectra - the 43, 57
and 71 lines give you plenty of difference between the two. The 43 and
71 lines are missing from the pentan-3-one spectrum, and the 57 line is
missing from the pentan-2-one one.

Note: Don't confuse the line at m/z = 58 in the pentan-2-one


spectrum. That's due to a complicated rearrangement which
you couldn't possibly predict at A'level.

The two spectra look like this:

http://www.chemguide.co.uk/analysis/masspec/fragment.html (11 of 13)30/12/2004 11:41:50

mass spectra - fragmentation patterns

Computer matching of mass spectra


As you've seen, the mass spectrum of even very similar organic
compounds will be quite different because of the different fragmentations
that can occur. Provided you have a computer data base of mass
spectra, any unkown spectrum can be computer analysed and simply
matched against the data base.

Where would you like to go now?


To the mass spectrometry menu . . .
To the instrumental analysis menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/analysis/masspec/fragment.html (12 of 13)30/12/2004 11:41:50

mass spectra - fragmentation patterns

http://www.chemguide.co.uk/analysis/masspec/fragment.html (13 of 13)30/12/2004 11:41:50

infra-red spectroscopy menu

Understanding Chemistry

INFRA-RED SPECTROSCOPY MENU

Background . . .
An explanation of how an infra-red spectrum arises.
The fingerprint region . . .
How an infra-red spectrum can be used to identify a compound.
Identifying the presence of particular groups . . .
How you can use an infra-red spectrum to identify a few easily
recognised groups in an organic compound.

Go to instrumental analysis menu . . .


Go to Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/analysis/irmenu.html30/12/2004 11:41:51

the background to infra-red spectroscopy

WHAT IS AN INFRA-RED SPECTRUM?

This page describes what an infra-red spectrum is and how it arises from
bond vibrations within organic molecules.

The background to infra-red spectroscopy


How an infra-red spectrum is produced
You probably know that visible light is made up of a continuous range of
different electromagnetic frequencies - each frequency can be seen as a
different colour. Infra-red radiation also consists of a continuous range of
frequencies - it so happens that our eyes can't detect them.
If you shine a range of infra-red frequencies one at a time through a
sample of an organic compound, you find that some frequencies get
absorbed by the compound. A detector on the other side of the
compound would show that some frequencies pass through the
compound with almost no loss, but other frequencies are strongly
absorbed.
How much of a particular frequency gets through the compound is
measured as percentage transmittance.
A percentage transmittance of 100 would mean that all of that frequency
passed straight through the compound without any being absorbed. In
practice, that never happens - there is always some small loss, giving a
transmittance of perhaps 95% as the best you can achieve.
A transmittance of only 5% would mean that nearly all of that particular
frequency is absorbed by the compound. A very high absorption of this
sort tells you important things about the bonds in the compound.

What an infra-red spectrum looks like

http://www.chemguide.co.uk/analysis/ir/background.html (1 of 6)30/12/2004 11:41:56

the background to infra-red spectroscopy

A graph is produced showing how the percentage transmittance varies


with the frequency of the infra-red radiation.

Note: The infra-red spectra on this page have been


produced from graphs taken from the Spectral Data Base
System for Organic Compounds (SDBS) at the National
Institute of Materials and Chemical Research in Japan.
It is possible that small errors may have been introduced
during the process of converting them for use on this site, but
these won't affect the argument in any way.

http://www.chemguide.co.uk/analysis/ir/background.html (2 of 6)30/12/2004 11:41:56

the background to infra-red spectroscopy

Notice that an unusual measure of frequency is used on the horizontal


axis. Wavenumber is defined like this:

Don't worry about this - just accept it!


Similarly, don't worry about the change of scale half-way across the
horizontal axis. You will find infra-red spectra where the scale is
consistent all the way across, infra-red spectra where the scale changes
at around 2000 cm-1, and very occasionally where the scale changes
again at around 1000 cm-1.
As you will see when we look at how to interpret infra-red spectra, this
doesn't cause any problems - you simply need to be careful reading the
horizontal scale.

What causes some frequencies to be absorbed?


Each frequency of light (including infra-red) has a certain energy. If a
particular frequency is being absorbed as it passes through the
compound being investigated, it must mean that its energy is being
transferred to the compound.
Energies in infra-red radiation correspond to the energies involved in
bond vibrations.
Bond stretching
In covalent bonds, atoms aren't joined by rigid links - the two atoms are
held together because both nuclei are attracted to the same pair of
electrons. The two nuclei can vibrate backwards and forwards - towards
and away from each other - around an average position.

http://www.chemguide.co.uk/analysis/ir/background.html (3 of 6)30/12/2004 11:41:56

the background to infra-red spectroscopy

The energy involved in this vibration depends on things like the length of
the bond and the mass of the atoms at either end. That means that each
different bond will vibrate in a different way, involving different amounts
of energy.
Bonds are vibrating all the time, but if you shine exactly the right amount
of energy on a bond, you can kick it into a higher state of vibration. The
amount of energy it needs to do this will vary from bond to bond, and so
each different bond will absorb a different frequency (and hence energy)
of infra-red radiation.
Bond bending
As well as stretching, bonds can also bend.

Again, bonds will be vibrating like this all the time and, again, if you shine
exactly the right amount of energy on the bond, you can kick it into a
higher state of vibration. Since the energies involved with the bending will
be different for each kind of bond, each different bond will absorb a
different frequency of infra-red radiation in order to make this jump from
one state to a higher one.
http://www.chemguide.co.uk/analysis/ir/background.html (4 of 6)30/12/2004 11:41:56

the background to infra-red spectroscopy

Tying all this together


Look again at the infra-red spectrum of propan-1-ol, CH3CH2CH2OH:

In the diagram, three sample absorptions are picked out to show you the
bond vibrations which produced them. Notice that bond stretching and
bending produce different troughs in the spectrum.

Note: If you want to go straight on to explore how you


interpret infra-red spectra, you could follow this link rather
than going via the menu.

http://www.chemguide.co.uk/analysis/ir/background.html (5 of 6)30/12/2004 11:41:56

the background to infra-red spectroscopy

Where would you like to go now?


To the infra-red spectroscopy menu . . .
To the instrumental analysis menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/analysis/ir/background.html (6 of 6)30/12/2004 11:41:56

interpreting infra-red spectra

INTERPRETING AN INFRA-RED SPECTRUM

This page explains how to use an infra-red spectrum to identify the


presence of a few simple bonds in organic compounds.

Note: This page follows directly on from the introductory


page on infra-red spectra. If you haven't already done so, you
should read that page before you go on.

The infra-red spectrum for a simple carboxylic acid


Ethanoic acid
Ethanoic acid has the structure:

You will see that it contains the following bonds:


carbon-oxygen double, C=O
carbon-oxygen single, C-O
oxygen-hydrogen, O-H
carbon-hydrogen, C-H
carbon-carbon single, C-C
The carbon-carbon bond has absorptions which occur over a wide range
of wavenumbers in the fingerprint region - that makes it very difficult to
pick out on an infra-red spectrum.
http://www.chemguide.co.uk/analysis/ir/interpret.html (1 of 8)30/12/2004 11:42:03

interpreting infra-red spectra

The carbon-oxygen single bond also has an absorbtion in the fingerprint


region, varying between 1000 and 1300 cm-1 depending on the molecule
it is in. You have to be very wary about picking out a particular trough as
being due to a C-O bond.

Note: If you aren't sure what the fingerprint region is, follow
this link before you go on.

The other bonds in ethanoic acid have easily recognised absorptions


outside the fingerprint region.
The C-H bond (where the hydrogen is attached to a carbon which is
singly-bonded to everything else) absorbs somewhere in the range from
2853 - 2962 cm-1. Because that bond is present in most organic
compounds, that's not terribly useful! What it means is that you can
ignore a trough just under 3000 cm-1, because that is probably just due
to C-H bonds.
The carbon-oxygen double bond, C=O, is one of the really useful
absorptions, found in the range 1680 - 1750 cm-1. Its position varies
slightly depending on what sort of compound it is in.
The other really useful bond is the O-H bond. This absorbs differently
depending on its environment. It is easily recognised in an acid because
it produces a very broad trough in the range 2500 - 3300 cm-1.

Note: You will never have to remember where these


absorptions occur. You will always be given this data in an
A'level exam.

http://www.chemguide.co.uk/analysis/ir/interpret.html (2 of 8)30/12/2004 11:42:03

interpreting infra-red spectra

The infra-red spectrum for ethanoic acid looks like this:

The possible absorption due to the C-O single bond is queried because it
lies in the fingerprint region. You couldn't be sure that this trough wasn't
caused by something else.

Note: The infra-red spectra on this page have been


produced from graphs taken from the Spectral Data Base
System for Organic Compounds (SDBS) at the National
Institute of Materials and Chemical Research in Japan.
It is possible that small errors may have been introduced
during the process of converting them for use on this site, but
these won't affect the argument in any way.

http://www.chemguide.co.uk/analysis/ir/interpret.html (3 of 8)30/12/2004 11:42:03

interpreting infra-red spectra

The infra-red spectrum for an alcohol


Ethanol

The O-H bond in an alcohol absorbs at a higher wavenumber than it


does in an acid - somewhere between 3230 - 3550 cm-1. In fact this
absorption would be at a higher number still if the alcohol isn't hydrogen
bonded - for example, in the gas state. All the infra-red spectra on this
page are from liquids - so that possibility will never apply.
Notice the absorption due to the C-H bonds just under 3000 cm-1, and
also the troughs between 1000 and 1100 cm-1 - one of which will be due
to the C-O bond.

The infra-red spectrum for an ester


Ethyl ethanoate

http://www.chemguide.co.uk/analysis/ir/interpret.html (4 of 8)30/12/2004 11:42:03

interpreting infra-red spectra

This time the O-H absorption is missing completely. Don't confuse it with
the C-H trough fractionally less than 3000 cm-1. The presence of the
C=O double bond is seen at about 1740 cm-1.
The C-O single bond is the absorption at about 1240 cm-1. Whether or
not you could pick that out would depend on the detail given by the table
of data which you get in your exam, because C-O single bonds vary
anywhere between 1000 and 1300 cm-1 depending on what sort of
compound they are in. Some tables of data fine it down, so that they will
tell you that an absorption from 1230 - 1250 is the C-O bond in an
ethanoate.

The infra-red spectrum for a ketone


Propanone

http://www.chemguide.co.uk/analysis/ir/interpret.html (5 of 8)30/12/2004 11:42:03

interpreting infra-red spectra

You will find that this is very similar to the infra-red spectrum for ethyl
ethanoate, an ester. Again, there is no trough due to the O-H bond, and
again there is a marked absorption at about 1700 cm-1 due to the C=O.
Confusingly, there are also absorptions which look as if they might be
due to C-O single bonds - which, of course, aren't present in propanone.
This reinforces the care you have to take in trying to identify any
absorptions in the fingerprint region.
Aldehydes will have similar infra-red spectra to ketones.

The infra-red spectrum for a hydroxy-acid


2-hydroxypropanoic acid (lactic acid)

http://www.chemguide.co.uk/analysis/ir/interpret.html (6 of 8)30/12/2004 11:42:03

interpreting infra-red spectra

This is interesting because it contains two different sorts of O-H bond the one in the acid and the simple "alcohol" type in the chain attached to
the -COOH group.
The O-H bond in the acid group absorbs between 2500 and 3300, the
one in the chain between 3230 and 3550 cm-1. Taken together, that
gives this immense trough covering the whole range from 2500 to 3550
cm-1. Lost in that trough as well will be absorptions due to the C-H bonds.
Notice also the presence of the strong C=O absorption at about 1730 cm1.

The infra-red spectrum for a primary amine


1-aminobutane

http://www.chemguide.co.uk/analysis/ir/interpret.html (7 of 8)30/12/2004 11:42:03

interpreting infra-red spectra

Primary amines contain the -NH2 group, and so have N-H bonds. These
absorb somewhere between 3100 and 3500 cm-1. That double trough
(typical of primary amines) can be seen clearly on the spectrum to the
left of the C-H absorptions.

Where would you like to go now?


To the infra-red spectroscopy menu . . .
To the instrumental analysis menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/analysis/ir/interpret.html (8 of 8)30/12/2004 11:42:03

infra-red spectra - the fingerprint region

THE FINGERPRINT REGION OF AN INFRA-RED


SPECTRUM

This page explains what the fingerprint region of an infra-red spectrum is,
and how it can be used to identify an organic molecule.

Note: It would be helpful if you first read the introductory


page on infra-red spectra if you haven't already done so.

What is the fingerprint region


This is a typical infra-red spectrum:

http://www.chemguide.co.uk/analysis/ir/fingerprint.html (1 of 4)30/12/2004 11:42:06

infra-red spectra - the fingerprint region

Note: The infra-red spectra on this page have been


produced from graphs taken from the Spectral Data Base
System for Organic Compounds (SDBS) at the National
Institute of Materials and Chemical Research in Japan.
It is possible that small errors may have been introduced
during the process of converting them for use on this site, but
these won't affect the argument in any way.

Each trough is caused because energy is being absorbed from that


particular frequency of infra-red radiation to excite bonds in the molecule
to a higher state of vibration - either stretching or bending.
Some of the troughs are easily used to identify particular bonds in a
molecule. For example, the big trough at the left-hand side of the
spectrum is used to identify the presence of an oxygen-hydrogen bond in
an -OH group.

Note: Using troughs in this way to identify particular bonds is


covered on a separate page.

The region to the right-hand side of the diagram (from about 1500 to 500
cm-1) usually contains a very complicated series of absorptions. These
are mainly due to all manner of bending vibrations within the molecule.
This is called the fingerprint region.
It is much more difficult to pick out individual bonds in this region than it
is in the "cleaner" region at higher wavenumbers. The importance of the
fingerprint region is that each different compound produces a different
pattern of troughs in this part of the spectrum.

Using the fingerprint region

http://www.chemguide.co.uk/analysis/ir/fingerprint.html (2 of 4)30/12/2004 11:42:06

infra-red spectra - the fingerprint region

Compare the infra-red spectra of propan-1-ol and propan-2-ol. Both


compounds contain exactly the same bonds. Both compounds have very
similar troughs in the area around 3000 cm-1 - but compare them in the
fingerprint region between 1500 and 500 cm-1.

The pattern in the fingerprint region is completely different and could


therefore be used to identify the compound.
So . . . to positively identify an unknown compound, use its infra-red
http://www.chemguide.co.uk/analysis/ir/fingerprint.html (3 of 4)30/12/2004 11:42:06

infra-red spectra - the fingerprint region

spectrum to identify what sort of compound it is by looking for specific


bond absorptions. That might tell you, for example, that you had an
alcohol because it contained an -OH group.
You would then compare the fingerprint region of its infra-red spectrum
with known spectra measured under exactly the same conditions to find
out which alcohol (or whatever) you had.

Where would you like to go now?


To the infra-red spectroscopy menu . . .
To the instrumental analysis menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/analysis/ir/fingerprint.html (4 of 4)30/12/2004 11:42:06

nuclear magnetic resonance (nmr) menu

Understanding Chemistry

NUCLEAR MAGNETIC RESONANCE MENU

Background . . .
An explanation of how an NMR spectrum arises, and the meaning
of the term "chemical shift".
Low resolution NMR spectra . . .
How a low resolution NMR spectrum is used to identify where the
hydrogen atoms in a molecule are. Read this before you go on to
high resolution spectra.
High resolution NMR spectra . . .
Looks at the additional information which you can get from a high
resolution NMR spectrum.
Integrator traces . . .
How to find the ratio of the numbers of differently placed hydrogen
atoms from an integrator trace.

Go to instrumental analysis menu . . .


Go to Main Menu . . .

http://www.chemguide.co.uk/analysis/nmrmenu.html (1 of 2)30/12/2004 11:42:07

nuclear magnetic resonance (nmr) menu

Jim Clark 2000

http://www.chemguide.co.uk/analysis/nmrmenu.html (2 of 2)30/12/2004 11:42:07

the background to nuclear magnetic resonance (nmr) spectroscopy

WHAT IS NUCLEAR MAGNETIC RESONANCE


(NMR)?

This page describes what a proton NMR spectrum is and how it how it
tells you useful things about the hydrogen atoms in organic molecules.

The background to NMR spectroscopy


Nuclear magnetic resonance is concerned with the magnetic properties
of certain nuclei. At A'level we focus on the magnetic behaviour of
hydrogen nuclei - hence the term proton NMR or 1H-NMR.
Hydrogen atoms as little magnets
If you have a compass needle, it normally lines up with the Earth's
magnetic field with the north-seeking end pointing north. Provided it isn't
sealed in some sort of container, you could twist the needle around with
your fingers so that it pointed south - lining it up opposed to the Earth's
magnetic field.
It is very unstable opposed to the Earth's field, and as soon as you let it
go again, it will flip back to its more stable state.

Hydrogen nuclei also behave as little magnets and a hydrogen nucleus


can also be aligned with an external magnetic field or opposed to it.

http://www.chemguide.co.uk/analysis/nmr/background.html (1 of 8)30/12/2004 11:42:12

the background to nuclear magnetic resonance (nmr) spectroscopy

Again, the alignment where it is opposed to the field is less stable (at a
higher energy). It is possible to make it flip from the more stable
alignment to the less stable one by supplying exactly the right amount of
energy.

The energy needed to make this flip depends on the strength of the
external magnetic field used, but is usually in the range of energies found
in radio waves - at frequencies of about 60 - 100 MHz. (BBC Radio 4 is
found between 92 - 95 MHz!)
It's possible to detect this interaction between the radio waves of just the
right frequency and the proton as it flips from one orientation to the other
as a peak on a graph. This flipping of the proton from one magnetic
alignment to the other by the radio waves is known as the resonance
condition.

The importance of the hydrogen atom's environment


What we've said so far would apply to an isolated proton, but real protons
have other things around them - especially electrons. The effect of the
electrons is to cut down the size of the external magnetic field felt by the
hydrogen nucleus.

http://www.chemguide.co.uk/analysis/nmr/background.html (2 of 8)30/12/2004 11:42:12

the background to nuclear magnetic resonance (nmr) spectroscopy

Suppose you were using a radio frequency of 90 MHz, and you adjusted
the size of the magnetic field so that an isolated proton was in the
resonance condition.
If you replaced the isolated proton with one that was attached to
something, it wouldn't be feeling the full effect of the external field any
more and so would stop resonating (flipping from one magnetic
alignment to the other). The resonance condition depends on having
exactly the right combination of external magnetic field and radio
frequency.
How would you bring it back into the resonance condition again? You
would have to increase the external magnetic field slightly to compensate
for the effect of the electrons.
Now suppose that you attached the hydrogen to something more
electronegative. The electrons in the bond would be further away from
the hydrogen nucleus, and so would have less effect on the magnetic
field around the hydrogen.

http://www.chemguide.co.uk/analysis/nmr/background.html (3 of 8)30/12/2004 11:42:12

the background to nuclear magnetic resonance (nmr) spectroscopy

Note: Electronegativity is a measure of the ability of an atom


to attract a bonding pair of electrons. If you aren't happy
about electronegativity, you could follow this link at some
point in the future, but it probably isn't worth doing it now!

To bring the hydrogen back into resonance you would again have to
increase the external magnetic field slightly to compensate for the effect
of the electrons - but not as much as when the hydrogen was attached to
the X atom.
Summary
For a given radio frequency (say, 90 MHz) each hydrogen atom will need
a slightly different magnetic field applied to it to bring it into the
resonance condition depending on what exactly it is attached to - in other
words the magnetic field needed is a useful guide to the hydrogen atom's
environment in the molecule.

Features of an NMR spectrum


A simple NMR spectrum looks like this:

http://www.chemguide.co.uk/analysis/nmr/background.html (4 of 8)30/12/2004 11:42:12

the background to nuclear magnetic resonance (nmr) spectroscopy

Note: The nmr spectra on this page have been produced


from graphs taken from the Spectral Data Base System for
Organic Compounds (SDBS) at the National Institute of
Materials and Chemical Research in Japan.
It is possible that small errors may have been introduced
during the process of converting them for use on this site, but
these won't affect the argument in any way.

The peaks
There are two peaks because there are two different environments for
the hydrogens - in the CH3 group and attached to the oxygen in the
COOH group. They are in different places in the spectrum because they
need slightly different external magnetic fields to bring them in to
resonance at a particular radio frequency.
The sizes of the two peaks gives important information about the
numbers of hydrogen atoms in each environment. It isn't the height of the
peaks that matters, but the ratio of the areas under the peaks. If you
could measure the areas under the peaks in the diagram above, you
would find that they were in the ratio of 3 (for the larger peak) to 1 (for the
smaller one).
That shows a ratio of 3:1 in the number of hydrogen atoms in the two
environments - which is exactly what you would expect for CH3COOH.

The need for a standard for comparison - TMS


Before we can explain what the horizontal scale means, we need to
explain the fact that it has a zero point - at the right-hand end of the
scale. The zero is where you would find a peak due to the hydrogen
atoms in tetramethylsilane - usually called TMS. Everything else is
compared with this.

http://www.chemguide.co.uk/analysis/nmr/background.html (5 of 8)30/12/2004 11:42:12

the background to nuclear magnetic resonance (nmr) spectroscopy

You will find that some NMR spectra show the peak due to TMS (at
zero), and others leave it out. Essentially, if you have to analyse a
spectrum which has a peak at zero, you can ignore it because that's the
TMS peak.
TMS is chosen as the standard for several reasons. The most important
are:

It has 12 hydrogen atoms all of which are in exactly the same


environment. They are joined to exactly the same things in exactly
the same way. That produces a single peak, but it's also a strong
peak (because there are lots of hydrogen atoms).
The electrons in the C-H bonds are closer to the hydrogens in this
compound than in almost any other one. That means that these
hydrogen nuclei are the most shielded from the external magnetic
field, and so you would have to increase the magnetic field by the
greatest amount to bring the hydrogens back into resonance.
The net effect of this is that TMS produces a peak on the spectrum
at the extreme right-hand side. Almost everything else produces
peaks to the left of it.

The chemical shift


The horizontal scale is shown as (ppm). is called the chemical shift
and is measured in parts per million - ppm.
A peak at a chemical shift of, say, 2.0 means that the hydrogen atoms
which caused that peak need a magnetic field two millionths less than
the field needed by TMS to produce resonance.
A peak at a chemical shift of 2.0 is said to be downfield of TMS. The
further to the left a peak is, the more downfield it is.

http://www.chemguide.co.uk/analysis/nmr/background.html (6 of 8)30/12/2004 11:42:12

the background to nuclear magnetic resonance (nmr) spectroscopy

Solvents for NMR spectroscopy


NMR spectra are usually measured using solutions of the substance
being investigated. It is important that the solvent itself doesn't contain
any simple hydrogen atoms, because they would produce confusing
peaks in the spectrum.
There are two ways of avoiding this. You can use a solvent such as
tetrachloromethane, CCl4, which doesn't contain any hydrogen, or you
can use a solvent in which any ordinary hydrogen atoms are replaced by
its isotope, deuterium - for example, CDCl3 instead of CHCl3. All the
NMR spectra used on this site involve CDCl3 as the solvent.
Deuterium atoms have sufficiently different magnetic properties from
ordinary hydrogen that they don't produce peaks in the area of the
spectrum that we are looking at.

Note: Several text books say that deuterium atoms don't


have a magnetic field. It isn't true - they do have a field but it
is less than an ordinary hydrogen atom.

Where would you like to go now?


To the NMR menu . . .
To the instrumental analysis menu . . .
To Main Menu . . .

Jim Clark 2000


http://www.chemguide.co.uk/analysis/nmr/background.html (7 of 8)30/12/2004 11:42:12

the background to nuclear magnetic resonance (nmr) spectroscopy

http://www.chemguide.co.uk/analysis/nmr/background.html (8 of 8)30/12/2004 11:42:12

low resolution nuclear magnetic resonance (nmr) spectra

LOW RESOLUTION NMR SPECTRA

This page describes how you interpret simple low resolution nuclear
magnetic resonance (NMR) spectra. It assumes that you have already
read the background page on NMR so that you understand what an NMR
spectrum looks like and the use of the term "chemical shift".

Note: If you haven't read the background page on NMR, you


really ought to do that before you go on.

The difference between high and low resolution spectra

Note: This high resolution nmr spectrum has been produced


from a graph taken from the Spectral Data Base System for
Organic Compounds (SDBS) at the National Institute of
Materials and Chemical Research in Japan.

http://www.chemguide.co.uk/analysis/nmr/lowres.html (1 of 7)30/12/2004 11:42:18

low resolution nuclear magnetic resonance (nmr) spectra

A low resolution spectrum looks much simpler because it can't


distinguish between the individual peaks in the various groups of peaks.

Note: I haven't been able to find a reliable source of low


resolution NMR spectra. The low resolution spectra on this
page are my best guess at what the low resolution spectra
would look like, based on the high resolution ones.

The numbers against the peaks represent the relative areas under each
peak. That information is extremely important in interpreting the spectra.

Interpreting a low resolution spectrum


Using the total number of peaks
Each peak represents a different environment for hydrogen atoms in the
molecule. In the methyl propanoate spectrum above, there are three
peaks because there are three different environments for the hydrogens.
Remember that methyl propanoate is CH3CH2COOCH3. The hydrogens
in the CH2 group are obviously in a different environment from those in
the CH3 groups. The two CH3 groups aren't in the same environment
either. One is attached to a CH2 group, the other to an oxygen.

http://www.chemguide.co.uk/analysis/nmr/lowres.html (2 of 7)30/12/2004 11:42:18

low resolution nuclear magnetic resonance (nmr) spectra

Using the areas under the peaks


The ratio of the areas under the peaks tell you the ratio of the numbers of
hydrogens in the various environments. In the methyl propanoate case,
the areas were in the ratio of 3:2:3, which is exactly what you want for
the two differently placed CH3 groups and the CH2 group.
You will probably be told the relative areas under the peaks - especially if
you are only looking at low resolution spectra, but it is just possible that
you might have to work them out. NMR spectrometers have a device
which draws another line on the spectrum called an integrator trace (or
integration trace). You can measure the relative areas from this trace.

Note: You need to find out whether your examiners expect


you to know how to interpret an integrator trace. Check your
syllabus and, particularly, past papers to see whether they
ask it. If you haven't got copies of your syllabus and past
papers, follow this link to find out how to get them.
If you do need to be able to interpret integrator traces, you
can find out how by following this link. You can also find it
from the NMR menu.

Using chemical shifts


The position of the peaks tells you useful things about what groups the
various hydrogen atoms are in. In any exam, you will be given a table of
chemical shifts if you need them. The important shifts for the groups
present in methyl propanoate are:

http://www.chemguide.co.uk/analysis/nmr/lowres.html (3 of 7)30/12/2004 11:42:18

low resolution nuclear magnetic resonance (nmr) spectra

Notes: "R" represents an alkyl group (like methyl, ethyl, etc)


which in this case may have other things substituted in it.
The shifts are shown as ranges of values. The exact position
varies depending on what else is near that particular group in
the molecule.

Showing these groups on the low resolution spectrum gives:

Some sample questions


Example 1
An organic compound was known to be one of the following. Use its low

http://www.chemguide.co.uk/analysis/nmr/lowres.html (4 of 7)30/12/2004 11:42:18

low resolution nuclear magnetic resonance (nmr) spectra

resolution NMR spectrum to decide which it is.

Notice that there are three peaks showing three different environments
for the hydrogens. That eliminates methyl ethanoate as a possibility
because that would only give two peaks - due to the two differently
situated CH3 group hydrogens.
Does the ratio of the areas under the peaks help? Not in this case - both
the other compounds would have three peaks in the ratio of 1:2:3.
Now you need to look at the chemical shifts:

http://www.chemguide.co.uk/analysis/nmr/lowres.html (5 of 7)30/12/2004 11:42:18

low resolution nuclear magnetic resonance (nmr) spectra

Checking the positions of the various hydrogens in the two possible


compounds against the chemical shift table gives you this pattern of
shifts:

Comparing these with the actual spectrum means that the substance
was propanoic acid, CH3CH2COOH.

Example 2
How would you use low resolution NMR to distinguish between the
isomers propanone and propanal?

The propanone would only give one peak in its NMR spectrum because
both CH3 groups are in an identical environment - both are attached to COCH3.
The propanal would give three peaks with the areas underneath in the
ratio 3:2:1.

http://www.chemguide.co.uk/analysis/nmr/lowres.html (6 of 7)30/12/2004 11:42:18

low resolution nuclear magnetic resonance (nmr) spectra

You could refer to the chemical shift table above to decide where the
peaks are likely to be found, but it isn't really necessary.

Example 3
How many peaks would there be in the low resolution NMR spectrum of
the following compound, and what would be the ratio of the areas under
the peaks?

All the CH3 groups are exactly equivalent so would only produce 1 peak.
There would also be peaks for the hydrogens in the CH2 group and the
COOH group.
There would be three peaks in total with areas in the ratio 9:2:1.

Where would you like to go now?


To the NMR menu . . .
To the instrumental analysis menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/analysis/nmr/lowres.html (7 of 7)30/12/2004 11:42:18

nmr spectra - the integrator trace

NMR SPECTRA - THE INTEGRATOR TRACE

This page describes how you use an integrator trace (or integration
trace) to find the ratio of the numbers of hydrogen atoms in different
environments in an organic compound.

Note: If you have come straight to this page via a search


engine, it might pay you to explore the NMR menu before you
go on.

What an integrator trace looks like


An integrator trace is a computer generated line which is superimposed
on an NMR spectrum. In the diagram, the integrator trace is shown in red.

http://www.chemguide.co.uk/analysis/nmr/integration.html (1 of 3)30/12/2004 11:42:21

nmr spectra - the integrator trace

Note: This high resolution nmr spectrum has been produced


from a graph taken from the Spectral Data Base System for
Organic Compounds (SDBS) at the National Institute of
Materials and Chemical Research in Japan.
The integrator trace isn't a real one! I haven't been able to
find a source of NMR spectra which include the trace, and so
this is a simulation. That doesn't affect the argument in any
way.

What an integrator trace shows


An integrator trace measures the relative areas under the various peaks
in the spectrum. When the integrator trace crosses a peak or group of
peaks, it gains height. The height gained is proportional to the area under
the peak or group of peaks.
You measure the height gained at each peak or group of peaks by
measuring the distances shown in green in the diagram above - and then
find their ratio.
For example, if the heights were 0.7 cm, 1.4 cm and 2.1 cm, the ratio of
the peak areas would be 1:2:3.
That in turn shows that the ratio of the hydrogen atoms in the three
different environments is 1:2:3.

Note: Where the integrator trace crosses a group of peaks, it


gains height in a series of steps - one step for each of the
sub-peaks. When you measure the height gained, you
measure the total for that group of peaks, ignoring the
individual steps.

http://www.chemguide.co.uk/analysis/nmr/integration.html (2 of 3)30/12/2004 11:42:21

nmr spectra - the integrator trace

Where would you like to go now?


To the NMR menu . . .
To the instrumental analysis menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/analysis/nmr/integration.html (3 of 3)30/12/2004 11:42:21

high resolution nuclear magnetic resonance (nmr) spectra

HIGH RESOLUTION NMR SPECTRA

This page describes how you interpret simple high resolution nuclear
magnetic resonance (NMR) spectra. It assumes that you have already
read the background page on NMR so that you understand what an NMR
spectrum looks like and the use of the term "chemical shift". It also
assumes that you know how to interpret simple low resolution spectra.

Note: If you haven't read the background page on NMR or


the page on low resolution NMR, you really ought to read
them before you go on.

The difference between high and low resolution spectra


What a low resolution NMR spectrum tells you
Remember:

The number of peaks tells you the number of different


environments the hydrogen atoms are in.
The ratio of the areas under the peaks tells you the ratio of the
numbers of hydrogen atoms in each of these environments.
The chemical shifts give you important information about the sort
of environment the hydrogen atoms are in.

High resolution NMR spectra


In a high resolution spectrum, you find that many of what looked like
single peaks in the low resolution spectrum are split into clusters of
peaks.
For A'level purposes, you will only need to consider these possibilities:

http://www.chemguide.co.uk/analysis/nmr/highres.html (1 of 10)30/12/2004 11:42:27

high resolution nuclear magnetic resonance (nmr) spectra

1 peak
2 peaks in the cluster
3 peaks in the cluster
4 peaks in the cluster

a singlet
a doublet
a triplet
a quartet

You can get exactly the same information from a high resolution
spectrum as from a low resolution one - you simply treat each cluster of
peaks as if it were a single one in a low resolution spectrum.
But in addition, the amount of splitting of the peaks gives you important
extra information.

Interpreting a high resolution spectrum


The n+1 rule
The amount of splitting tells you about the number of hydrogens attached
to the carbon atom or atoms next door to the one you are currently
interested in.
The number of sub-peaks in a cluster is one more than the number of
hydrogens attached to the next door carbon(s).
So - on the assumption that there is only one carbon atom with
hydrogens on next door to the carbon we're interested in (usually true at
A'level!):
singlet
doublet
triplet

next door to carbon with no hydrogens attached


next door to a CH group
next door to a CH2 group

quartet

next door to a CH3 group

http://www.chemguide.co.uk/analysis/nmr/highres.html (2 of 10)30/12/2004 11:42:27

high resolution nuclear magnetic resonance (nmr) spectra

Note: You probably won't need to know the origin of the n+1
rule, but if you are interested there is a page on the reasons
for splitting which you could look at.

Using the n+1 rule


What information can you get from this NMR spectrum?

Note: The nmr spectra on this page have been produced


from data taken from the Spectral Data Base System for
Organic Compounds (SDBS) at the National Institute of
Materials and Chemical Research in Japan.
Any small errors that I've introduced during the process of
converting them for use on this site won't affect the argument
in any way.

http://www.chemguide.co.uk/analysis/nmr/highres.html (3 of 10)30/12/2004 11:42:27

high resolution nuclear magnetic resonance (nmr) spectra

Assume that you know that the compound above has the molecular
formula C4H8O2.
Treating this as a low resolution spectrum to start with, there are three
clusters of peaks and so three different environments for the hydrogens.
The hydrogens in those three environments are in the ratio 2:3:3. Since
there are 8 hydrogens altogether, this represents a CH2 group and two
CH3 groups.
What about the splitting?
The CH2 group at about 4.1 ppm is a quartet. That tells you that it is next
door to a carbon with three hydrogens attached - a CH3 group.
The CH3 group at about 1.3 ppm is a triplet. That must be next door to a
CH2 group.
This combination of these two clusters of peaks - one a quartet and the
other a triplet - is typical of an ethyl group, CH3CH2. It is very common.
Get to recognise it!
Finally, the CH3 group at about 2.0 ppm is a singlet. That means that the
carbon next door doesn't have any hydrogens attached.
So what is this compound? You would also use chemical shift data to
help to identify the environment each group was in, and eventually you
would come up with:

http://www.chemguide.co.uk/analysis/nmr/highres.html (4 of 10)30/12/2004 11:42:27

high resolution nuclear magnetic resonance (nmr) spectra

Note: You now know how to get the information you need
from NMR spectra, but it often isn't easy to fit all that
information together into a final formula. You simply need to
practise! Go through all the examples in past papers from
your Exam Board. How complicated they are will vary
markedly from Board to Board. Some of the compounds you
will come across may be very unfamiliar. Don't forget to use
the information in chemical shift tables - if your examiners
include some obscure group, it's almost certain you will need
to use it. Take all the hints that are going!

Two special cases


Alcohols
Where is the -O-H peak?
This is very confusing! Different sources quote totally different chemical
shifts for the hydrogen atom in the -OH group in alcohols - often
inconsistently. For example:

The Nuffield Data Book quotes 2.0 - 4.0, but the Nuffield text book
shows a peak at about 5.4.
The OCR Data Sheet for use in their exams quotes 3.5 - 5.5.
A reliable degree level organic chemistry text book quotes1.0 - 5.0,
but then shows an NMR spectrum for ethanol with a peak at about
6.1.
The SDBS database (used throughout this site) gives the -OH
peak in ethanol at about 2.6.

http://www.chemguide.co.uk/analysis/nmr/highres.html (5 of 10)30/12/2004 11:42:27

high resolution nuclear magnetic resonance (nmr) spectra

The problem seems to be that the position of the -OH peak varies
dramatically depending on the conditions - for example, what solvent is
used, the concentration, and the purity of the alcohol - especially on
whether or not it is totally dry.

Help! Do you need to worry about this? Not really - you can
assume that in an exam question, any NMR spectrum will be
consistent with the chemical shift data you are given.

A clever way of picking out the -OH peak


If you measure an NMR spectrum for an alcohol like ethanol, and then
add a few drops of deuterium oxide, D2O, to the solution, allow it to settle
and then re-measure the spectrum, the -OH peak disappears! By
comparing the two spectra, you can tell immediately which peak was due
to the -OH group.

Note: Deuterium oxide (sometimes called "heavy water") is


simply water in which all the normal hydrogen-1 atoms are
replaced by its isotope, hydrogen-2 (or deuterium).

http://www.chemguide.co.uk/analysis/nmr/highres.html (6 of 10)30/12/2004 11:42:27

high resolution nuclear magnetic resonance (nmr) spectra

The reason for the loss of the peak lies in the interaction between the
deuterium oxide and the alcohol. All alcohols, such as ethanol, are very,
very slightly acidic. The hydrogen on the -OH group transfers to one of
the lone pairs on the oxygen of the water molecule. The fact that here
we've got "heavy water" makes no difference to that.

The negative ion formed is most likely to bump into a simple deuterium
oxide molecule to regenerate the alcohol - except that now the -OH
group has turned into an -OD group.

Deuterium atoms don't produce peaks in the same region of an NMR


spectrum as ordinary hydrogen atoms, and so the peak disappears.
You might wonder what happens to the positive ion in the first equation
and the OD- in the second one. These get lost into the normal equilibrium
which exists wherever you have water molecules - heavy or otherwise.

The lack of splitting with -OH groups


Unless the alcohol is absolutely free of any water, the hydrogen on the OH group and any hydrogens on the next door carbon don't interact to
produce any splitting. The -OH peak is a singlet and you don't have to
worry about its effect on the next door hydrogens.

http://www.chemguide.co.uk/analysis/nmr/highres.html (7 of 10)30/12/2004 11:42:27

high resolution nuclear magnetic resonance (nmr) spectra

The left-hand cluster of peaks is due to the CH2 group. It is a quartet


because of the 3 hydrogens on the next door CH3 group. You can ignore
the effect of the -OH hydrogen.
Similarly, the -OH peak in the middle of the spectrum is a singlet. It hasn't
turned into a triplet because of the influence of the CH2 group.

Note: The reason for this is quite complex, and certainly


goes beyond A'level. It lies in the very rapid interchange that
occurs between the hydrogen atoms on the -OH group and
either water molecules or other alcohol molecules. To find out
about it you will have to read either a degree level organic
chemistry book or one specifically about NMR.
For A'level purposes just accept the fact that -OH produces a
singlet and has no effect on neighbouring groups!

http://www.chemguide.co.uk/analysis/nmr/highres.html (8 of 10)30/12/2004 11:42:27

high resolution nuclear magnetic resonance (nmr) spectra

Equivalent hydrogen atoms


Hydrogen atoms attached to the same carbon atom are said to be
equivalent. Equivalent hydrogen atoms have no effect on each other so that one hydrogen atom in a CH2 group doesn't cause any splitting in
the spectrum of the other one.
But hydrogen atoms on neighbouring carbon atoms can also be
equivalent if they are in exactly the same environment. For example:

These four hydrogens are all exactly equivalent. You would get a single
peak with no splitting at all.
You only have to change the molecule very slightly for this no longer to
be true.

Because the molecule now contains different atoms at each end, the
hydrogens are no longer all in the same environment. This compound
would give two separate peaks on a low resolution NMR spectrum. The
high resolution spectrum would show that both peaks subdivided into
triplets - because each is next door to a differently placed CH2 group.

http://www.chemguide.co.uk/analysis/nmr/highres.html (9 of 10)30/12/2004 11:42:27

high resolution nuclear magnetic resonance (nmr) spectra

Where would you like to go now?


To the NMR menu . . .
To the instrumental analysis menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/analysis/nmr/highres.html (10 of 10)30/12/2004 11:42:27

origin of splitting in nmr spectra

SPLITTING IN HIGH RESOLUTION NMR


SPECTRA

This page describes the reason that you get clusters of peaks in a high
resolution NMR spectrum in place of simple peaks in the low resolution
spectrum. The effect is known as spin-spin coupling or spin-spin
splitting.

Note: If you have come straight to this page via a search


engine, you should realise that it is only a "footnote" to a
page on high resolution NMR spectra.

Spin-spin coupling
The origin of a doublet
Consider what the high resolution NMR spectrum of the compound
CH2Cl-CHCl2 would look like.

http://www.chemguide.co.uk/analysis/nmr/splitting.html (1 of 5)30/12/2004 11:42:30

origin of splitting in nmr spectra

Note: I'm sorry about the absence of a real spectrum. The


SDBS database didn't have a suitable NMR spectrum for this
compound.

Focus on the CH2 group. Why is that a doublet?


Remember that the peaks in an NMR spectrum are in different places
because the hydrogens are experiencing different magnetic fields due to
their different environments.
Two peaks close together must mean that those particular hydrogens are
experiencing two slightly different magnetic fields. Those two slightly
different fields are caused by the hydrogen in the CH group next door.
The hydrogen next door has a small magnetic field of its own, which
could be aligned with the external magnetic field or opposed to it.
Depending on which way around it is aligned, it will either strengthen or
weaken the field felt by the CH2 hydrogens.

There is an equal chance of either of these arrangements happening and


so there will be two peaks due to the CH2 hydrogens, close together and
with equal areas under them (because of the 50/50 chance of either
arrangement).

The origin of a triplet

http://www.chemguide.co.uk/analysis/nmr/splitting.html (2 of 5)30/12/2004 11:42:30

origin of splitting in nmr spectra

Now focus on the CH group in the compound CH2Cl-CHCl2. Why is that


a triplet? It must be a triplet because that hydrogen is experiencing any
one of three slightly different magnetic fields.
Think about the magnetic alignments of the hydrogens on the next door
CH2 group. These are the various possibilities:

The two arrangements in the centre of the diagram produce the same
field (exactly the same as the external field). So . . . there are three
possible magnetic fields that the CH hydrogen could feel, and so there
are three peaks close together - a triplet.
The areas under the peaks are in the ratio of 1:2:1 because that
represents the chances of these various magnetic fields occurring.

The origin of a quartet


If you apply the same sort of argument to hydrogens next door to a CH3
group, you will find that they could be experiencing any one of four
different magnetic fields depending on the alignment of the CH3
hydrogens.

http://www.chemguide.co.uk/analysis/nmr/splitting.html (3 of 5)30/12/2004 11:42:30

origin of splitting in nmr spectra

All the arrangements in the second line produce the same field. All the
alignments in the third line also produce the same field, but this time a bit
smaller. There are four different possible fields, with the chances of them
arising in the ratio 1:3:3:1.
So a CH3 group produces a quartet in the spectrum of the hydrogens of
the next door group, with the peak sizes in the ratio 1:3:3:1.

Where would you like to go now?


Return to the article on high resolution NMR . . .
To the NMR menu . . .
To the instrumental analysis menu . . .
To Main Menu . . .

http://www.chemguide.co.uk/analysis/nmr/splitting.html (4 of 5)30/12/2004 11:42:30

origin of splitting in nmr spectra

Jim Clark 2000

http://www.chemguide.co.uk/analysis/nmr/splitting.html (5 of 5)30/12/2004 11:42:30

mass spectra - the molecular ion (M+) peak

MASS SPECTRA - THE MOLECULAR ION (M+)


PEAK

This page explains how to find the relative formula mass (relative
molecular mass) of an organic compound from its mass spectrum. It also
shows how high resolution mass spectra can be used to find the
molecular formula for a compound.

Using a mass spectrum to find relative formula mass


The formation of molecular ions
When the vaporised organic sample passes into the ionisation chamber
of a mass spectrometer, it is bombarded by a stream of electrons. These
electrons have a high enough energy to knock an electron off an organic
molecule to form a positive ion. This ion is called the molecular ion.

Note: If you aren't sure about how a mass spectrum is


produced, it might be worth taking a quick look at the page
describing how a mass spectrometer works.

The molecular ion is often given the symbol M+ or


- the dot in this
second version represents the fact that somewhere in the ion there will
be a single unpaired electron. That's one half of what was originally a
pair of electrons - the other half is the electron which was removed in the
ionisation process.
The molecular ions tend to be unstable and some of them break into
smaller fragments. These fragments produce the familiar stick diagram.
Fragmentation is irrelevant to what we are talking about on this page - all
we're interested in is the molecular ion.

http://www.chemguide.co.uk/analysis/masspec/mplus.html (1 of 6)30/12/2004 11:42:34

mass spectra - the molecular ion (M+) peak

Note: If you are interested in a detailed look at fragmentation


patterns you could follow this link.

Using the molecular ion to find the relative formula mass


In the mass spectrum, the heaviest ion (the one with the greatest m/z
value) is likely to be the molecular ion. A few compounds have mass
spectra which don't contain a molecular ion peak, because all the
molecular ions break into fragments. That isn't a problem you are likely to
meet at A'level.
For example, in the mass spectrum of pentane, the heaviest ion has an
m/z value of 72.

http://www.chemguide.co.uk/analysis/masspec/mplus.html (2 of 6)30/12/2004 11:42:34

mass spectra - the molecular ion (M+) peak

Note: This mass spectrum has been drawn using data from
the Spectral Data Base System for Organic Compounds
(SDBS) at the National Institute of Materials and Chemical
Research in Japan.
It has been simplified by omitting all the minor lines with peak
heights of 2% or less of the base peak (the tallest peak).

Because the largest m/z value is 72, that represents the largest ion going
through the mass spectrometer - and you can reasonably assume that
this is the molecular ion. The relative formula mass of the compound is
therefore 72.

Note: This assumes that the charge on the ion is 1+. That's
always the case when you are interpreting these mass
spectra.

Finding the relative formula mass (relative molecular mass) from a mass
spectrum is therefore trivial. Look for the peak with the highest value for
m/z, and that value is the relative formula mass of the compound.
There are, however, complications which arise because of the possibility
of different isotopes (either of carbon or of chlorine or bromine) in the
molecular ion. These cases are dealt with on separate pages.

http://www.chemguide.co.uk/analysis/masspec/mplus.html (3 of 6)30/12/2004 11:42:34

mass spectra - the molecular ion (M+) peak

Note: The presence of the carbon-13 isotope in a molecular


ion causes a little peak 1 unit to the right of the M+ peak. This
is called the M+1 peak.
The presence of a chlorine atom in a compound causes two
peaks in the molecular ion region - the M+ peak and the M+2
peak depending on whether the particular molecular ion
contains a chlorine-35 or chlorine-37 isotope. Bromine
creates a similar problem. Follow these links if you are
interested - or explore them later via the mass spectrometry
menu.

Using a mass spectrum to find a molecular formula


So far we've been looking at m/z values in a mass spectrum as whole
numbers, but it's possible to get far more accurate results using a high
resolution mass spectrometer. You can use that more accurate
information about the mass of the molecular ion to work out the
molecular formula of the compound.
Accurate isotopic masses
For normal calculation purposes, you tend to use rounded-off relative
isotopic masses. For example, you are familiar with the numbers:
1H

12C

12

14N

14

16O

16

To 4 decimal places, however, these are the relative isotopic masses:


1H
12C

1.0078
12.0000

http://www.chemguide.co.uk/analysis/masspec/mplus.html (4 of 6)30/12/2004 11:42:34

mass spectra - the molecular ion (M+) peak

14N

14.0031

16O

15.9949

The carbon value is 12.0000, of course, because all the other masses
are measured on the carbon-12 scale which is based on the carbon-12
isotope having a mass of exactly 12.

Using these accurate values to find a molecular formula


Two simple organic compounds have a relative formula mass of 44 propene, C3H8, and ethanal, CH3CHO. Using a high reolution mass
spectrometer, you could easily decide which of these you had.
On a high resolution mass spectrometer, the molecular ion peaks for the
two compounds give the following m/z values:
C 3 H8

44.0624

CH3CHO

44.0261

You can easily check that by adding up numbers from the table of
accurate relative isotopic masses above.
A possible exam question
A gas was known to contain only elements from the following list:
1H

1.0078

12C

12.0000

14N

14.0031

16O

15.9949

The gas had a molecular ion peak at m/z = 28.0312 in a high resolution
mass spectrometer. What was the gas?
After a bit of playing around, you might reasonably come up with 3 gases
which had relative formula masses of approximately 28 and which
http://www.chemguide.co.uk/analysis/masspec/mplus.html (5 of 6)30/12/2004 11:42:34

mass spectra - the molecular ion (M+) peak

contained the elements from the list. They are N2, CO and C2H4.
Working out their accurate relative formula masses gives:
N2

28.0062

CO
C2H4

27.9949
28.0312

The gas is obviously C2H4.


In an exam, you would hope that - apart from the most simple cases you would be given the possible formulae to work from. Trying to work
out all the possible things which might add up to the value you want is
quite time-consuming - and it's easy to miss an important possibility!

Where would you like to go now?


To the mass spectrometry menu . . .
To the instrumental analysis menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/analysis/masspec/mplus.html (6 of 6)30/12/2004 11:42:34

mass spectra - the M+1 peak

MASS SPECTRA - THE M+1 PEAK

This page explains how the M+1 peak in a mass spectrum can be used
to estimate the number of carbon atoms in an organic compound.

Note: This is a small corner of mass spectrometry. It would


be a good idea not to attack this page unless you have a
reasonable idea about how a mass spectrum is produced
and the sort of information you can get from it. If you haven't
already done so, explore the mass spectrometry menu before
you go on.

What causes the M+1 peak?


What is an M+1 peak?
If you had a complete (rather than a simplified) mass spectrum, you will
find a small line 1 m/z unit to the right of the main molecular ion peak.
This small peak is called the M+1 peak.

http://www.chemguide.co.uk/analysis/masspec/mplus1.html (1 of 5)30/12/2004 11:42:36

mass spectra - the M+1 peak

In A'level questions, the M+1 peak is often left out to avoid confusion particularly if you were being asked to find the relative formula mass of
the compound from the molecular ion peak.
The carbon-13 isotope
The M+1 peak is caused by the presence of the 13C isotope in the
molecule. 13C is a stable isotope of carbon - don't confuse it with the 14C
isotope which is radioactive. Carbon-13 makes up 1.11% of all carbon
atoms.
If you had a simple compound like methane, CH4, approximately 1 in
every 100 of these molecules will contain carbon-13 rather than the more
common carbon-12. That means that 1 in every 100 of the molecules will
have a mass of 17 (13 + 4) rather than 16 (12 + 4).
The mass spectrum will therefore have a line corresponding to the
molecular ion [13CH4]+ as well as [12CH4]+.
The line at m/z = 17 will be much smaller than the line at m/z = 16
because the carbon-13 isotope is much less common. Statistically you
will have a ratio of approximately 1 of the heavier ions to every 99 of the
lighter ones. That's why the M+1 peak is much smaller than the M+ peak.

Using the M+1 peak


What happens when there is more than 1 carbon atom in the
compound?
Imagine a compound containing 2 carbon atoms. Either of them has an
approximately 1 in 100 chance of being 13C.

http://www.chemguide.co.uk/analysis/masspec/mplus1.html (2 of 5)30/12/2004 11:42:36

mass spectra - the M+1 peak

There's therefore a 2 in 100 chance of the molecule as a whole


containing one 13C atom rather than a 12C atom - which leaves a 98 in
100 chance of both atoms being 12C.
That means that the ratio of the height of the M+1 peak to the M+ peak
will be approximately 2 : 98. That's pretty close to having an M+1 peak
approximately 2% of the height of the M+ peak.

Note: You might wonder why both atoms can't be carbon-13,


giving you an M+2 peak. They can - and do! But statistically
the chance of both carbons being 13C is approximately 1 in
10,000. The M+2 peak will be so small that you couldn't
observe it.

Using the relative peak heights to predict the number of carbon


atoms
If there are small numbers of carbon atoms
If you measure the peak height of the M+1 peak as a percentage of the
peak height of the M+ peak, that gives you the number of carbon atoms
in the compound.
We've just seen that a compound with 2 carbons will have an M+1 peak
approximately 2% of the height of the M+ peak.
Similarly, you could show that a compound with 3 carbons will have the M
+1 peak at about 3% of the height of the M+ peak.
With larger numbers of carbon atoms

http://www.chemguide.co.uk/analysis/masspec/mplus1.html (3 of 5)30/12/2004 11:42:36

mass spectra - the M+1 peak

The approximations we are making won't hold with more than 2 or 3


carbons. The proportion of carbon atoms which are 13C isn't 1% - it's
1.11%. And the appoximation that a ratio of 2 : 98 is about 2% doesn't
hold as the small number increases.
Consider a molecule with 5 carbons in it. You could work out that 5.55 (5
x 1.11) molecules will contain 1 13C to every 94.45 (100 - 5.55) which
contain only 12C atoms. If you convert that to how tall the M+1 peak is as
a percentage of the M+ peak, you get an answer of 5.9% (5.55/94.45 x
100). That's close enough to 6% that you might assume wrongly that
there are 6 carbon atoms.
Above 3 carbon atoms, then, you shouldn't really be making the
approximation that the height of the M+1 peak as a percentage of the
height of the M+ peak tells you the number of carbons - you will need to
do some fiddly sums!

Important! Only one of the current A'level syllabuses


actually mentions the M+1 peak. Before you get too bogged
down in all this, check your syllabus and recent past papers
to see whether you need to bother. This is a case where you
really need to know what line your examiners are currently
taking - for example, how far are they pushing the
approximation.
If you haven't got a syllabus and past papers, follow this link
to find out how to get them.

http://www.chemguide.co.uk/analysis/masspec/mplus1.html (4 of 5)30/12/2004 11:42:36

mass spectra - the M+1 peak

Where would you like to go now?


To the mass spectrometry menu . . .
To the instrumental analysis menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/analysis/masspec/mplus1.html (5 of 5)30/12/2004 11:42:36

mass spectra - the M+2 peak

MASS SPECTRA - THE M+2 PEAK

This page explains how the M+2 peak in a mass spectrum arises from
the presence of chlorine or bromine atoms in an organic compound. It
also deals briefly with the origin of the M+4 peak in compounds
containing two chlorine atoms.

Note: Before you start this page, it would be a good idea to


have a reasonable understanding about how a mass
spectrum is produced and the sort of information you can get
from it. If you haven't already done so, explore the mass
spectrometry menu before you go on.

The effect of chlorine or bromine atoms on the mass


spectrum of an organic compound
Compounds containing chlorine atoms
One chlorine atom in a compound

http://www.chemguide.co.uk/analysis/masspec/mplus2.html (1 of 6)30/12/2004 11:42:41

mass spectra - the M+2 peak

Note: All the mass spectra on this page have been drawn
using data from the Spectral Data Base System for Organic
Compounds (SDBS) at the National Institute of Materials and
Chemical Research in Japan.
With one exception, they have been simplified by omitting all
the minor lines with peak heights of 2% or less of the base
peak (the tallest peak).

The molecular ion peaks (M+ and M+2) each contain one chlorine atom but the chlorine can be either of the two chlorine isotopes, 35Cl and 37Cl.
The molecular ion containing the 35Cl isotope has a relative formula
mass of 78. The one containing 37Cl has a relative formula mass of 80 hence the two lines at m/z = 78 and m/z = 80.
Notice that the peak heights are in the ratio of 3 : 1. That reflects the fact
that chlorine contains 3 times as much of the 35Cl isotope as the 37Cl
one. That means that there will be 3 times more molecules containing the
lighter isotope than the heavier one.
So . . . if you look at the molecular ion region, and find two peaks
separated by 2 m/z units and with a ratio of 3 : 1 in the peak heights, that
tells you that the molecule contains 1 chlorine atom.
You might also have noticed the same pattern at m/z = 63 and m/z = 65
in the mass spectrum above. That pattern is due to fragment ions also
containing one chlorine atom - which could either be 35Cl or 37Cl. The
fragmentation that produced those ions was:

http://www.chemguide.co.uk/analysis/masspec/mplus2.html (2 of 6)30/12/2004 11:42:41

mass spectra - the M+2 peak

Note: If you aren't sure about fragmentation you might like to


have a look at this link.

Two chlorine atoms in a compound

Note: This spectrum has been simplified by omitting all the


minor lines with peak heights of less than 1% of the base
peak (the tallest peak). This contains more minor lines than
other mass spectra in this section. It was necessary because
otherwise an important line in the molecular ion region would
have been missing.

http://www.chemguide.co.uk/analysis/masspec/mplus2.html (3 of 6)30/12/2004 11:42:41

mass spectra - the M+2 peak

The lines in the molecular ion region (at m/z values of 98, 100 ands 102)
arise because of the various combinations of chlorine isotopes that are
possible. The carbons and hydrogens add up to 28 - so the various
possible molecular ions could be:
28 + 35 + 35 = 98
28 + 35 + 37 = 100
28 + 37 + 37 = 102
If you have the necessary maths, you could show that the chances of
these arrangements occurring are in the ratio of 9:6:1 - and this is the
ratio of the peak heights. If you don't know the right bit of maths, just
learn this ratio!
So . . . if you have 3 lines in the molecular ion region (M+, M+2 and M+4)
with gaps of 2 m/z units between them, and with peak heights in the ratio
of 9:6:1, the compound contains 2 chlorine atoms.

Compounds containing bromine atoms


Bromine has two isotopes, 79Br and 81Br in an approximately 1:1 ratio
(50.5 : 49.5 if you want to be fussy!). That means that a compound
containing 1 bromine atom will have two peaks in the molecular ion
region, depending on which bromine isotope the molecular ion contains.
Unlike compounds containing chlorine, though, the two peaks will be very
similar in height.

http://www.chemguide.co.uk/analysis/masspec/mplus2.html (4 of 6)30/12/2004 11:42:41

mass spectra - the M+2 peak

The carbons and hydrogens add up to 29. The M+ and M+2 peaks are
therefore at m/z values given by:
29 + 79 = 108
29 + 81 = 110
So . . . if you have two lines in the molecular ion region with a gap of 2 m/
z units between them and with almost equal heights, this shows the
presence of a bromine atom in the molecule.

Where would you like to go now?


To the mass spectrometry menu . . .
To the instrumental analysis menu . . .
To Main Menu . . .

Jim Clark 2000

http://www.chemguide.co.uk/analysis/masspec/mplus2.html (5 of 6)30/12/2004 11:42:41

mass spectra - the M+2 peak

http://www.chemguide.co.uk/analysis/masspec/mplus2.html (6 of 6)30/12/2004 11:42:41

chemistry a level textbook suggestions

Understanding Chemistry

CHEMISTRY TEXTBOOK SUGGESTIONS

You can get almost all the books on this page directly from Amazon.co.
uk, and you will find links to the Amazon site where appropriate. You can,
of course, also buy the books from normal bookshops.
The suggestions are mainly suitable for UK A level chemistry courses. If
you want to make life as easy as possible for yourself, I think there is a
lot to be said for using books designed especially to cover your particular
syllabus. The content is then exactly what you need - and no more.
The alternative is to use a short revision guide for all the facts, and to get
your explanations in detail from a website like this one.

The author's books


You will find more details about my two books (one GCSE and one A
level) on their own pages on this site.

Calculations in AS / A Level Chemistry

Longman GCSE Chemistry

Revision notes covering all syllabuses

http://www.chemguide.co.uk/suggestions.html (1 of 5)30/12/2004 11:42:47

chemistry a level textbook suggestions

There are lots of these around. Of the ones I have looked at, the books in
the Letts series seem to me to be the easiest and most logical to find
your way around - with clear coding to show what is needed for each
syllabus.

Letts: Revise AS Chemistry

Letts: Revise A2 Chemistry

Letts: Chemistry AS / A2 Exam Practice

Books for specific syllabuses


These are written in cooperation with the various Exam Boards. I am not
personally familiar with all the following books, but most of them have
excellent 4 or 5 star reviews by users on the Amazon.co.uk site.

AQA
Textbooks
Collins and Heinemann both publish textbooks written specifically for the
AQA AS and A2 syllabuses.

Collins Advanced Modular Science: Chemistry AS

Collins Advanced Modular Science: Chemistry A2

Advanced Level Chemistry for AQA: AS Student Book

Advanced Level Chemistry for AQA: A2 Student Book

Revision books
The same publishers both produce a set of revision notes.

http://www.chemguide.co.uk/suggestions.html (2 of 5)30/12/2004 11:42:47

chemistry a level textbook suggestions

The Collins Student Support Materials provide one booklet for each
module.

Atomic Structure, Bonding and Periodicity

Foundation Physical and Inorganic Chemistry

Introduction to Organic Chemistry

Further Physical and Organic Chemistry

Thermodynamics and Further Inorganic Chemistry

The Heinemann material comes in two books - one for AS and one for
A2.

Heinemann: Revise AS Chemistry for AQA

Heinemann: Revise A2 Chemistry for AQA

Edexcel
Nelson Advanced Science produces 4 books to cover each of the theory
units in the AS and A2 syllabuses, and an overall revision guide.
Textbooks

Structure, Bonding and Main Group Chemistry

Organic Chemistry, Energetics, Kinetics and Equilibrium

Periodicity, Quantitative Equilibria and Functional Group Chemistry

Transition Metals, Quantitative Kinetics and Applied Organic


Chemistry

Revision books

http://www.chemguide.co.uk/suggestions.html (3 of 5)30/12/2004 11:42:47

chemistry a level textbook suggestions

Make the Grade in AS and A Chemistry

OCR
Textbooks
There are two textbooks for the OCR syllabus in the Cambridge
Advanced Science series.

Cambridge Advanced Science: Chemistry 1. This is the AS book.

Cambridge Advanced Science: Chemistry 2 (the A2 book).

The OCR A2 syllabus has an options section. The following books in the
Cambridge Advanced Science series cover these options:

Biochemistry

Environmental Chemistry

Methods of Analysis and Detection

Gases, Liquids and Solids

Transition Elements

Revision books

Letts: Chemistry (OCR Endorsed AS Revision Notes).

At the time of writing, there wasn't a set of equivalent A2 revision notes


from Letts.

Nuffield and Salters

http://www.chemguide.co.uk/suggestions.html (4 of 5)30/12/2004 11:42:47

chemistry a level textbook suggestions

I don't have enough experience of either of these two syllabuses to be


able to make any useful suggestions.
The current Nuffield Students Book is very poorly reviewed on the
Amazon.co.uk site. I have a previous edition of this book, and I know that
it is probably the least helpful book I possess in terms of finding any
simple information. That is because it was never intended to be used as
a source of information, but as a work book or study guide for the
Nuffield A level chemistry course.
If you want an easy source of facts, you could use the Letts Revision
guides further up the page.

Go to Main Menu . . .

Jim Clark 2003

http://www.chemguide.co.uk/suggestions.html (5 of 5)30/12/2004 11:42:47

S-ar putea să vă placă și