You are on page 1of 467

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. relative mass 1 1 1/1836 relative charge +1 0 -1

proton neutron electron

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

The behaviour of protons, neutrons and electrons in electric fields What happens if a beam of each of these particles is passed between two electrically charged plates - one positive and one negative? Opposites will attract. Protons are positively charged and so would be deflected on a curving path towards the negative plate. Electrons are negatively charged and so would be deflected on a curving path towards the positive plate. Neutrons don't have a charge, and so would continue on in a straight line.

Exactly what happens depends on whether the beams of particles enter the

electric field with the various particles having the same speeds or the same energies If the particles have the same energy If beams of the three sorts of particles, all with the same energy, are passed between two electrically charged plates:
y y

Protons are deflected on a curved path towards the negative plate. Electrons are deflected on a curved path towards the positive plate. The amount of deflection is exactly the same in the electron beam as the proton beam if the energies are the same - but, of course, it is in the opposite direction.

Neutrons continue in a straight line.

If the electric field was strong enough, then the electron and proton beams might curve enough to hit their respective plates.

If the particles have the same speeds If beams of the three sorts of particles, all with the same speed, are passed between two electrically charged plates:
y y

Protons are deflected on a curved path towards the negative plate. Electrons are deflected on a curved path towards the positive plate. If the electrons and protons are travelling with the same speed, then the lighter electrons are deflected far more strongly than the heavier protons.

Neutrons continue in a straight line.

Note: This is potentially very confusing! Most chemistry sources that talk about this give either one or the other of these two diagrams without any comment at all - they don't specifically say that they are using constant energy or constant speed beams. But it matters! If this is on your syllabus, it is important that you should know which version your examiners are going to expect, and they probably won't tell you in the syllabus. You should look in detail at past questions, mark schemes and examiner's reports which you can get from your examiners if you are doing a UK-based syllabus. Information about how to do this is on the syllabuses page. If in doubt, I suggest you use the second (constant speed) version. This actually produces more useful information about both masses and charges than the constant energy version.

The nucleus
The nucleus is at the centre of the atom and contains the protons and neutrons. Protons and neutrons are collectively known asnucleons. 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 ofproton 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 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. protons 6 6 6 neutrons 6 7 8 mass number 12 13 14

carbon-12 carbon-13 carbon-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 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
y

y y

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


y y y

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

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:
y

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:

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

ATOMIC ORBITALS
This page explains what atomic orbitals are in a way that makes them understandable for introductory courses such as UK A level and its equivalents. 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 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 says - loosely - that you can't know with certainty both where an electron is and where it's going next. (What it actually says is that it is impossible to define with absolute precision, at the same time, both the position and the momentum of an electron.) 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.
Note: Over the years I have had a steady drip of questions from students in which it is obvious that they still think of electrons as orbiting around a nucleus - which is completely wrong! I have added a page about why the idea of orbits is wrong to try to avoid having to say the same thing over and over again!

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.

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 across-section 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. 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.
Taking chemistry further: 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? At this introductory level 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.

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 the moment, you need to be aware that there are sets of five d orbitals at levels from the third level upwards, but you probably won't be expected to draw them or name them. Apart from a passing reference, you won't come across f orbitals at all.
Note: Some UK-based syllabuses will eventually want you to be able to draw, or at least recognise, the shapes of d orbitals. I am not including them now because I don't want to add confusion to what is already a difficult introductory topic. Check your syllabus and past papers to find out what you need to know. If you are a studying a UK-based syllabus and haven't got these, follow this link to find out how to get hold of them.

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 2pzorbitals); 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 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.
Taking chemistry further: 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. 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 UK-based exam 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 the scope of this site. 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.

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

UK syllabuses for 16 - 18 year olds tend to 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 for chemistry at a higher 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.
Important! You must have a copy of your syllabus and copies of recent exam papers. If you are studying a UK-based syllabus and 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 C N 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.

The next electrons to go in will have to pair up with those already there. O F Ne 1s22s22px22py12pz1 1s22s22px22py22pz1 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 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 [Ne]3s2 [Ne]3s23px23py13pz1 [Ne]3s23px23py23pz2

Mg S Ar

1s22s22p63s2 1s22s22p63s23px23py13pz1 1s22s22p63s23px23py23pz2

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 Ca 1s22s22p63s23p64s1 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

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-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 righthand end of the d-block) 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 Ti 1s22s22p63s23p63d14s2 1s22s22p63s23p63d24s2

V Cr

1s22s22p63s23p63d34s2 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 Fe Co Ni Cu Zn 1s22s22p63s23p63d54s2 1s22s22p63s23p63d64s2 1s22s22p63s23p63d74s2 1s22s22p63s23p63d84s2 1s22s22p63s23p63d104s1 1s22s22p63s23p63d104s2 (back to being tidy again)

(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


y y

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.

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.

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 mol-1) 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 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.

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.

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.

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.

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 B

1s22s2 1s22s22px1

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

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 2-level. Mg Al 1s22s22p63s2 1s22s22p63s23px1 1st I.E. = 736 kJ mol-1 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.
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 O 1s22s22px12py12pz1 1s22s22px22py12pz1 1st I.E. = 1400 kJ mol-1 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:

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 Na 1s22s1 1s22s22p63s1 1st I.E. = 519 kJ mol-1 1st I.E. = 494 kJ mol-1

Lithium's outer electron is in the second level, and only has the 1s2electrons 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

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 Zn [Ar]3d104s1 [Ar]3d104s2 1st I.E. = 745 kJ mol-1 1st I.E. = 908 kJ mol-1

In each case, the electron is coming from the same orbital, with identical screening, but the zinc has one extra proton in the nucleus and so the attraction is greater. There will be a degree of repulsion between the paired up electrons in the 4s orbital, but in this case it obviously isn't enough to outweigh the effect of the extra proton.
Note: This is actually very similar to the increase from, say, sodium to magnesium in the third period. In that case, the outer electronic structure is going from 3s1 to 3s2. Despite the pairing-up of the electrons, the ionisation energy increases because of the extra proton in the nucleus. The repulsion

between the 3s electrons obviously isn't enough to outweigh this either. I don't know why the repulsion between the paired electrons matters less for electrons in s orbitals than in p orbitals (I don't even know whether you can make that generalisation!). I suspect that it has to do with orbital shape and possibly the greater penetration of s electrons towards the nucleus, but I haven't been able to find any reference to this anywhere. In fact, I haven't been able to find anyone who even mentions repulsion in the context of paired s electrons! If you have any hard information on this, could you contact me via the address on the about this site page.

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
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. The first electron affinities of the group 7 elements

F Cl Br I

-328 kJ mol-1 -349 kJ mol-1 -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.

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.
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 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 2- ions. This is more easily seen in symbol terms.

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. 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 structuresyou 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 radiusdepending on the bonding. 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 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 lefthand 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.

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.

IONIC RADIUS
A warning! Ionic radii are difficult to measure with any degree of certainty, and vary according to the environment of the ion. For example, it matters what the co-ordination of the ion is (how many oppositely charged ions are touching it), and what those ions are. There are several different measures of ionic radii in use, and these all differ from each other by varying amounts. It means that if you are going to make reliable comparisons using ionic radii, they have to come from the same source. What you have to remember is that there are quite big uncertainties in the use of ionic radii, and that trying to explain things in fine detail is made difficult by those uncertainties. What follows will be adequate for UK A level (and its various equivalents), but detailed explanations are too complicated for this level.

Trends in ionic radius in the Periodic Table Trends in ionic radius down a group This is the easy bit! As you add extra layers of electrons as you go down a group, the ions are bound to get bigger. The two tables below show this effect in Groups 1 and 7. electronic structure of ion

ionic radius (nm)

Li+ Na+ K+ Rb+ Cs+

2 2, 8 2, 8, 8 2, 8, 18, 8 2, 8, 18, 18, 8

0.076 0.102 0.138 0.152 0.167

electronic structure of ion FClBrI2, 8 2, 8, 8 2, 8, 18, 8 2, 8, 18, 18, 8

ionic radius (nm) 0.133 0.181 0.196 0.220

Note: These figures all come from the Database of Ionic Radii from Imperial College London. I have converted them from Angstroms to nm (nanometres), which are more often used in the data tables that you are likely to come across. If you are interested, 1 Angstrom is 10-10 m; 1 nm = 10-9 m. To convert from Angstroms to nm, you have to divide by 10, so that 1.02 Angstroms becomes 0.102 nm. You may also come across tables listing values in pm (picometres) which are 10-12 m. A value in pm will look like, for example, for chlorine, 181 pm rather than 0.181 nm. Don't worry if you find this confusing. Just use the values you are given in whatever units you are given. For comparison purposes, all the values relate to 6-co-ordinated ions (the same arrangement as in NaCl, for example). CsCl actually crystallises in an 8:8-co-ordinated structure - so you couldn't accurately use these values for CsCl. The 8-co-ordinated ionic radius for Cs is 0.174 nm rather than 0.167 for the 6-co-ordinated version.

Trends in ionic radius across a period

Let's look at the radii of the simple ions formed by elements as you go across Period 3 of the Periodic Table - the elements from Na to Cl. Na+ no of protons electronic structure of ion ionic radius (nm) 11 2,8 0.102 Mg2+ 12 2,8 0.072 Al3+ 13 2,8 0.054 P315 2,8,8 (0.212) S216 2,8,8 0.184 Cl17 2,8,8 0.181

Note: The table misses out silicon which doesn't form a simple ion. The phosphide ion radius is in brackets because it comes from a different data source, and I am not sure whether it is safe to compare it. The values for the oxide and chloride ions agree in the different source, so it is probably OK. The values are again for 6-co-ordination, although I can't guarantee that for the phosphide figure.

First of all, notice the big jump in ionic radius as soon as you get into the negative ions. Is this surprising? Not at all - you have just added a whole extra layer of electrons. Notice that, within the series of positive ions, and the series of negative ions, that the ionic radii fall as you go across the period. We need to look at the positive and negative ions separately. The positive ions In each case, the ions have exactly the same electronic structure - they are said to be isoelectronic. However, the number of protons in the nucleus of the ions is increasing. That will tend to pull the electrons more and more towards the centre of the ion - causing the ionic radii to fall. That is pretty obvious! The negative ions Exactly the same thing is happening here, except that you have an extra layer of electrons. What needs commenting on, though is how similar in size the sulphide ion and the chloride ion are. The additional proton here is making hardly any difference. The difference between the size of similar pairs of ions actually gets even smaller as you go down Groups 6 and 7. For example, the Te2- ion is only 0.001 nm bigger than the I- ion. As far as I am aware there is no simple explanation for this - certainly not one which can be used at this level. This is a good illustration of what I said earlier - explaining things

involving ionic radii in detail is sometimes very difficult.

Trends in ionic radius for some more isoelectronic ions This is only really a variation on what we have just been talking about, but fits negative and positive isoelectronic ions into the same series of results. Remember that isoelectronic ions all have exactly the same electron arrangement. N3no of protons electronic structure of ion ionic radius (nm) 7 2, 8 (0.171) O28 2, 8 0.140 F9 2, 8 0.133 Na+ 11 2, 8 0.102 Mg2+ 12 2, 8 0.072 Al3+ 13 2, 8 0.054

Note: The nitride ion value is in brackets because it came from a different source, and I don't know for certain whether it relates to the same 6-co-ordination as the rest of the ions. This matters. My main source only gave a 4-co-ordinated value for the nitride ion, and that was 0.146 nm. You might also be curious as to how the neutral neon atom fits into this sequence. It would seem logical that its van der Waals radius would fall neatly between that of the fluoride ion and the sodium ion. It doesn't! Its radius is 0.154 or 0.160 nm (depending on which source you look the value up in) - bigger than the fluoride ion. I have no idea why that is!

You can see that as the number of protons in the nucleus of the ion increases, the electrons get pulled in more closely to the nucleus. The radii of the isoelectronic ions therefore fall across this series.

The relative sizes of ions and atoms You probably won't have noticed, but nowhere in what you have read so far has there been any need to talk about the relative sizes of the ions and the atoms they have come from. Neither (as far as I can tell from the syllabuses) do any of the current UK-based exams for 16 - 18 year olds ask for this specifically in their syllabuses. However, it is very common to find statements about the relative sizes of ions and atoms. I

am fairly convinced that these statements are faulty, and I would like to attack the problem head-on rather than just ignoring it. Important! For 10 years, until I rewrote this ionic radius section in August 2010, I included what is in the box below. You will find this same information and explanation in all sorts of books and on any number of websites aimed at this level. At least one non-UK A level syllabus has a statement which specifically asks for this. 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. However, I was challenged by an experienced teacher about the negative ion explanation, and that forced me to think about it carefully for the first time. I am now convinced that the facts and the explanation relating to negative ions are simply illogical. As far as I can tell, no UK-based syllabus mentions the relative sizes of atoms and ions (as of August 2010), but you should check past papers and mark schemes to see whether questions have sneaked in. The rest of this page discusses the problems that I can see, and is really aimed at teachers and others, rather than at students. If you are a student, look carefully at your syllabus, and past exam questions and mark schemes, to find out whether you need to know about

this. If you don't need to know about it, stop reading now (unless, of course, you are interested in a bit of controversy!). If you do need to know it, then you will have to learn what is in the box, even if, as I believe, it is wrong. If you like your chemistry to be simple, ignore the rest of the page, because you risk getting confused about what you need to know. If you have expert knowledge of this topic, and can find any flaws in what I am saying, then please contact me via the address on the about this sitepage.

Choosing the right atomic radius to compare with This is at the heart of the problem. The diagrams in the box above, and similar ones that you will find elsewhere, use the metallic radius as the measure of atomic radius for metals, and the covalent radius for non-metals. I want to focus on the non-metals, because that is where the main problem lies. You are, of course, perfectly free to compare the radius of an ion with whatever measure of atomic radius you choose. The problem comes in relating your choice of atomic radius to the "explanation" of the differences. It is perfectly true that negative ions have radii which are significantly bigger than the covalent radius of the atom in question. And the argument then goes that the reason for this is that if you add one or more extra electrons to the atom, inter-electron repulsions cause the atom to expand. Therefore the negative ion is bigger than the atom. This seems to me to be completely inconsistent. If you add one or more extra electrons to the atom, you aren't adding them to a covalently bound atom. You can't simply add electrons to a covalently-bound chlorine atom, for example - chlorine's existing electrons have reorganised themselves into new molecular orbitals which bind the atoms together. In a covalently-bound atom, there is simply no room to add extra electrons. So if you want to use the electron repulsion explanation, the implication is that you are adding the extra electrons to a raw atom with a simple uncombined electron arrangement. In other words, if you were talking about, say, chlorine, you are adding an extra electron to chlorine with a configuration of 2,8,7 - not to covalently bound chlorine atoms in which the arrangement of the electrons has been altered by sharing.

That means that the comparison that you ought to be making isn't with the shortened covalent radius, but with the much larger van der Waals radius - the only available measure of the radius of an uncombined atom. So what happens if you make that comparison?

Group 7 vdW radius (nm) F Cl Br I 0.147 0.175 0.185 0.198 ionic radius of X- (nm) 0.133 0.181 0.196 0.220

Group 6 vdW radius (nm) O S Se Te 0.152 0.180 0.190 0.206 ionic radius of X2- (nm) 0.140 0.184 0.198 0.221

Group 5 vdW radius (nm) N 0.155 ionic radius of X3- (nm) 0.171

0.180

0.212

As we have already discussed above, measurements of ionic radii are full of uncertainties. That is also true of van der Waals radii. The table uses one particular set of values for comparison purposes. If you use data from different sources, you will find differences in the patterns - including which of the species (ion or atom) is bigger. These ionic radius values are for 6-co-ordinated ions (with a slight question mark over the nitride and phosphide ion figures). But you may remember that I said that ionic radius changes with co-ordination. Nitrogen is a particularly good example of this. 4-co-ordinated nitride ions have a radius of 0.146 nm. In other words if you look at one of the co-ordinations, the nitride ion is bigger than the nitrogen atom; in the other case, it is smaller. Making a general statement that nitride ions are bigger or smaller than nitrogen atoms is impossible.

So what is it safe to say about the facts? For most, but not all, negative ions, the radius of the ion is bigger than that of the atom, but the difference is nothing like as great as is shown if you incorrectly compare ionic radii with covalent radii. There are also important exceptions. I can't see how you can make any real generalisations about this, given the uncertainties in the data.

And what is it safe to say about the explanation? If there are any additional electron-electron repulsions on adding extra electrons, they must be fairly small. This is particularly shown if you consider some pairs of isoelectronic ions. You would have thought that if repulsion was an important factor, then the radius of, say a sulphide ion, with two negative charges would be significantly larger than a chloride ion with only one. The difference should actually be even more marked, because the sulphide electrons are being held by only 16 protons rather than the 17 in the chlorine case. On this repulsion theory, the sulphide ion shouldn't just be a little bit bigger than a chloride ion - it should be a lot bigger. The same effect is shown with selenide and bromide, and with telluride and iodide ions. In the last case, there is virtually no difference in the sizes of the 2- and 1- ions.

So if there is some repulsion playing a part in this, it certainly doesn't look as if it is playing a major part.

What about positive ions? Whether you choose to use van der Waals radii or metallic radii as a measure of the atomic radius, for metals the ionic radius is smaller than either, so the problem doesn't exist to the same extent. It is true that the ionic radius of a metal is less than its atomic radius (however vague you are about defining this). The explanation (at least as long as you only consider positive ions from Groups 1, 2 and 3) in terms of losing a complete layer of electrons is also acceptable.

Conclusion It seems to me that, for negative ions, it is completely illogical to compare ionic radii with covalent radii if you want to use the electron repulsion explanation. If you compare the ionic radii of negative ions with the van der Waals radii of the atoms they come from, the uncertainties in the data make it very difficult to make any reliable generalisations. The similarity in sizes of pairs of isoelectronic ions from Groups 6 and 7 calls into question how important repulsion is in any explanation. Having spent more than a week working on this, and discussing it with input from some very knowledgable people, I don't think there is any explanation which is simple enough to give to most students at this level. It would seem to me to be better that these ideas about relative sizes of atoms and ions are just dropped. At this level, you can describe and explain simple periodic trends in atomic radii in the way I did further up this page, without even thinking about the relative sizes of the atoms and ions. Personally, I would be more than happy never to think about this again for the rest of my life! 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.

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 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

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.

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.

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 dots-and-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-inboxes notation. Only the 2-level electrons are shown. The 1s2electrons 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 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.

The carbon atom is now said to be in an excited state.


Note: People sometimes worry that the promoted electron is drawn as an up-arrow, whereas it started as a down-arrow. The reason for this is actually fairly complicated - well beyond the level we are working at. Just get in the habit of writing it like this because it makes the diagrams look tidy!

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 sp3hybrids (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".

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 covalentlybound molecule.
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 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 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 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 NCl3exists, 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.
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.

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.

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 co-ordinate bond.

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

The second diagram shows another way that you might find coordinate 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 about 180C. If it simply 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 when it sublimes at this relatively low temperature, 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. Measurements of the relative formula mass of aluminium chloride show that its formula in the vapour at the sublimation temperature 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 AlCl3molecules.
Note: Aluminium chloride is complicated because of the way it keeps changing its bonding as the temperature increases. If you are interested in exploring this in more detail, you could have a look at the page about the Period 3 chlorides. It isn't particularly relevant to the present page, though. If you choose to follow this link, use the BACK button on your browser to return quickly to this page later.

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. 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 3-level electrons to leave 1s22s22p6. That means that all the 3-level orbitals are now empty. The aluminium re-organises (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:

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: It looks as if only one current UK 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.

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-andcrosses picture. The bonding involves delocalisation. LECTRONEGATIVITY

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 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 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
y y y

No electronegativity difference between two atoms leads to a pure nonpolar covalent bond. 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.

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.

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.

Explaining the patterns in electronegativity The attraction that a bonding pair of electrons feels for a particular nucleus depends on:
y y y

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 orshielding, 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.

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 3-level as it is in chlorine. If it is closer to the nucleus, the attraction is greater. As you go down a group, electronegativity decreases because the bonding pair of electrons is increasingly distant from the attraction of the nucleus.

Warning! As far as I am aware, none of the UK-based A level (or equivalent) syllabuses any longer want the next bit. It used to be on the AQA syllabus, but has

been removed from their new syllabus. At the time of writing, it does, however, still appear on at least one overseas A level syllabus (Malta, but there may be others that I'm not aware of). If in doubt, check your syllabus. Otherwise, ignore the rest of this page. It is an alternative (and, to my mind, more awkward) way of looking at the formation of a polar bond. Reading it unnecessarily just risks confusing you.

The polarising ability of positive ions


What do we mean by "polarising ability"? In the discussion so far, we've looked at the formation of polar bonds from the point of view of the distortions which occur in a covalent bond if one atom is more electronegative than the other. But you can also look at the formation of polar covalent bonds by imagining that you start from ions. 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.

We say that the aluminium ions polarise the chloride ions. In the case of aluminium chloride, the electron pairs are dragged back towards the aluminium to such an extent that the bonds become covalent. But because the chlorine is more electronegative than aluminium, the electron pairs won't be pulled half way between the two atoms, and so the bond formed will be polar.

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. 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.

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 if you are doing one of the UK-based exams for 16 - 18 year olds.

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 papers. If you are working to a UK-based syllabus for 16 - 18 year olds, and 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:
y

y y

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:
y y

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:
y

Arrange these electron pairs in space to minimise repulsions. How this is done will become clear in the examples which follow.

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 beinglinear.

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 calledtrigonal planar.

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 oncovalent 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.

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. It is my habit to draw diagrams like this with the bond at the top in the plane of the paper, the middle bond at the bottom coming out towards you, and the other two going back in. But that's all it is - a habit! You can equally well draw it differently if you rotate the molecule a bit. This is all described in some detail about half-way down the page about drawing organic molecules. Use the BACK button on your browser to return here later if you choose to follow this link.

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 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 ispyramidal - 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. 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.

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. 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 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.
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.

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.

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.

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. 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. 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 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 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. 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 I'm going to use the same lozenge-shaped diagram now to represent any molecule which could, in fact, be a much more complicated shape. Shape does matter (see below), but keeping the shape simple makes it

a lot easier to both draw the diagrams and understand what is going on. 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.

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 neon argon krypton xenon radon -269C -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. 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 2-methylpropane molecules.

van der Waals forces: dipole-dipole interactions


Warning! There's a bit of a problem here with modern 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, dipole-

dipole 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; dipoledipole 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 are working to a UK-based syllabus for 16 - 18 year olds, but don't have a copy of it, follow this link to find out how to get one.

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. Dipoledipole 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. 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 CCl4 61.2C 76.8C

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. If you plot the boiling points of the compounds of the Group 4 elements with hydrogen, 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.

If you repeat this exercise with the compounds of the elements in Groups 5, 6 and 7 with hydrogen, 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 compound of hydrogen with 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:

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.

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 dipoledipole 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.

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.

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.

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 100C.

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.

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-1-ol 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 double helix in DNA are held together by hydrogen bonds between hydrogen atoms attached to nitrogen on one strand, and lone pairs on another nitrogen or an oxygen on the other one. 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!

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.

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.

THE A'LEVEL VIEW OF IONIC BONDING


y y

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+ Cu2+ Zn2+ Ag+ Pb2+ [Ar]3d5 [Ar]3d9 [Ar]3d10 [Kr]4d10 [Xe]4f145d106s2

Noble gases (apart from helium) have an outer electronic structure ns2np6.
Note: If you aren't happy about writing electronic structuresusing 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.

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 calledionisation 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. 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.

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. It also explains why caesium chloride has a different structure from sodium chloride even though sodium and caesium are both in Group 1 of the Periodic Table.
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.

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.

The different structure of caesium chloride


We'll look first at the arrangement of the ions and then talk about why the structures of sodium chloride and caesium chloride are different afterwards.
Warning: Before you go on with this section, make sure that you actually need it for your syllabus. If you don't, jump down the page to "The physical properties of sodium chloride"

Why are the caesium chloride and sodium chloride structures different? When attractions are set up between two ions of opposite charges, energy is released. The more energy that can be released, the more stable the system becomes. That means that the more contact there is between negative and positive ions, the more stable the crystal should become. If you can surround a positive ion like caesium with eight chloride ions rather than just six (and vice versa for the chloride ions), then you should have a more stable crystal. So why doesn't sodium chloride do the same thing?

the physical properties of sodium chloride


Sodium chloride is taken as typical of ionic compounds, and is chosen rather than, say, caesium chloride, because it is found on every syllabus at this level. 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 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 co-ordinate (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. When it melts, sodium chloride undergoes electrolysis, which involves conduction of electricity because of the movement and discharge of the ions. In the process, sodium and chlorine are produced. This is achemical change rather than a physical process. The positive sodium ions move towards the negatively charged electrode (the cathode). When they get there, each sodium ion picks up an electron from the electrode to form a sodium atom.

The movement of electrons from the cathode onto the sodium ions leaves spaces on the cathode. The power source (the battery or whatever) moves electrons along the wire in the external circuit to fill those spaces. That flow of electrons would be seen as an electric current. (The external circuit is all the rest of the circuit apart from the molten sodium chloride.) Meanwhile, chloride ions are attracted to the positive electrode (the anode). When they get there, each chloride ion loses an electron to the anode to form an atom. These then pair up to make chlorine molecules. Overall, the change is . . .

The new electrons deposited on the anode are pumped off around the external circuit by the power source, eventually ending up on the cathode where they will be transferred to sodium ions. Molten sodium chloride conducts electricity because of the movement of the ions in the melt, and the discharge of the ions at the electrodes. Both of these

have to happen if you are to get electrons flowing in the external circuit. In solid sodium chloride, of course, that ion movement can't happen and that stops any possibility of any current flow in the circuit.

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. The physical properties of diamond Diamond
y

y y

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 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. 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.
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


y

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

y y

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.

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.

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


y

y y

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 silicon-oxygen 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.

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:
y y

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.

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 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. Alloyssuch 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.

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.
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 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


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 are studying a UK-based syllabus and haven't got a copy of your syllabus or copies of recent past papers, follow this link to find out how to get them.

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.

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".

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. 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.

LEMENTS
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


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:
y y y

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

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:

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 P4molecules. 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. 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


y y

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.
y y

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
y y

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 withwriting equations for redox reactions.

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


y y

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
y y

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:

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:
y y y y

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:

y y y y

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! 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. Itis 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 examiners' website. There are links on the syllabuses page for students studying for UK-based exams.

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. 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:
y y y

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:

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 half-equations 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. You will notice that I haven't bothered to include the electrons in the added-up 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.

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 (1and 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:
y y y y

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.

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.

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 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 left-hand 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.
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:

xidation 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-half-equations 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!
y

y y y

y y

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 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:

The reasons for the exceptions

y y

y y y y y

y y

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. 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 chlorinedoesn'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 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.

y y y y y y y y y y y y y

y y y y

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

ATOMIC AND PHYSICAL PROPERTIES OF THE GROUP 1 ELEMENTS


This page explores the trends in some atomic and physical properties of the Group 1 elements - lithium, sodium, potassium, rubidium and caesium. You will find separate sections below covering the trends in atomic radius, first ionisation energy, electronegativity, melting and boiling points, and density. 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.

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
y y

the number of layers of electrons around the nucleus the pull the outer electrons feel from the nucleus.

Compare lithium and sodium: Li Na 1s22s1 1s22s22p63s1


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 outer electron feels a net pull of 1+ from the nucleus. The positive charge on the nucleus is cut down by the negativeness of the inner electrons.

This is equally true for all the other atoms in Group 1. Work it out for potassium 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.

Trends in First Ionisation Energy


First ionisation energy is the energy needed to remove the most loosely held electron from each of 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

y y y

the charge on the nucleus, 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 1+ 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.

All of these elements have a very 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 potassium and rubidium have an electronegativity of 0.8. This can't be

explained just as a rounding error (as I have done elsewhere on the site in thecorresponding Group 2 case), because both elements have the same electronegativity to 2 decimal places as well. Both are 0.82. I'm not clear what the reason for this is! There are various other measures of electronegativity apart from the Pauling one, and on each of these the rubidium value is indeed smaller than the potassium one.

Explaining the decrease in electronegativity Imagine a bond between a sodium 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 because there is a much greater net pull from the chlorine nucleus than from the sodium 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 sodium is. Now compare this with the lithium-chlorine bond. The net pull from each end of the bond is the same as before, but you have to remember that the lithium atom is smaller than a sodium atom. That means that the electron pair is going to be closer to the net 1+ charge from the lithium end, and so more strongly attracted to it.

In some lithium compounds there is often a degree of covalent bonding that isn't there in the rest of the Group. Lithium iodide, for example, will dissolve in organic solvents - a typical property of covalent compounds. The iodine atom is so large that the pull from the iodine nucleus on the pair of electrons is relatively weak, and so a fully ionic bond isn't 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. With the exception of some lithium compounds, these elements all form compounds which we consider as being fully ionic. They are so weakly electronegative that we assume that the electron pair is pulled so far away towards the chlorine (or whatever) that ions are formed.

Trends in Melting and Boiling Points

You will see that both the melting points and boiling points fall as you go down the Group. Explaining the trends in melting and boiling points When you melt any of these metals, the metallic bond is weakened enough for the atoms to move around, and is then broken completely when you boil the metal. The fall in melting and boiling points reflects the fall in the strength of the metallic bond.
Note: If you aren't confident about metallic bonding you should follow this link before you go on. Use the BACK button on your browser to return quickly to this page.

The atoms in a metal are held together by the attraction of the nuclei to the delocalised electrons. As the atoms get bigger, the nuclei get further away from these delocalised electrons, and so the attractions fall. That means that the atoms are more easily pulled apart to make a liquid and finally a gas. In the same way that we have already discussed, each of these atoms has a net pull from the nuclei of 1+. The increased charge on the nucleus as you go down the Group is offset by additional levels of screening electrons. All that matters is the distance between the nucleus and the bonding electrons.
Note: This explanation seems fairly obvious, and works well for the Group 1 metals. However, if you have read about thecorresponding Group 2 case, you will know that life isn't always so simple!

Trends in Density

Notice that these are all light metals - and that the first three in the Group are less dense than water (less than 1 g cm-3). That means that the first three will float on water, while the other two sink. The density tends to increase as you go down the Group (apart from the fluctuation at potassium). Explaining the trend in density It is quite difficult to come up with a simple explanation for this, because the density depends on two factors, both of which are changing as you go down the Group. All of these metals have their atoms packed in the same way, so all you have to consider is how many atoms you can pack in a given volume, and what the mass of the individual atoms is. How many you can pack depends, of course, on their volume - and their volume, in turn, depends on their atomic radius. As you go down the Group, the atomic radius increases, and so the volume of the atoms increases as well. That means that you can't pack as many sodium atoms into a given volume as you can lithium atoms. However, as you go down the Group, the mass of the atoms increases. That means that a particular number of sodium atoms will weigh more than the same number of lithium atoms.

So 1 cm3 of sodium will contain fewer atoms than the same volume of lithium, but each atom will weigh more. What affect will that have on the density? It is completely impossible to say unless you do some sums!

REACTIONS OF THE GROUP 1 ELEMENTS WITH WATER

This page looks at the reactions of the Group 1 elements - lithium, sodium, potassium, rubidium and caesium - with water. It uses these reactions to explore the trend in reactivity in Group 1.

The Facts General All of these metals react vigorously or even explosively with cold water. In each case, a solution of the metal hydroxide is produced together with hydrogen gas.

This equation applies to any of these metals and water - just replace the X by the symbol you want. In each of the following descriptions, I am assuming a very small bit of the metal is dropped into water in a fairly large container.

Details for the individual metals Lithium Lithium's density is only about half that of water so it floats on the surface, gently fizzing and giving off hydrogen. It gradually reacts and disappears, forming a colourless solution of lithium hydroxide. The reaction generates heat too slowly and lithium's melting point is too high for it to melt (see sodium below).

Sodium Sodium also floats on the surface, but enough heat is given off to melt the sodium (sodium has a lower melting point than lithium and the reaction produces heat faster) and it melts almost at once to form a small silvery ball that dashes around the surface. A white trail of sodium hydroxide is seen in the water under the sodium, but this soon dissolves to give a colourless solution of sodium hydroxide. The sodium moves because it is pushed around by the hydrogen which is given off during the reaction. If the sodium becomes trapped on the side of the container, the hydrogen may catch fire to burn with an orange flame. The colour is due to contamination of the normally blue hydrogen flame with sodium compounds. Potassium Potassium behaves rather like sodium except that the reaction is faster and enough heat is given off to set light to the hydrogen. This time the normal hydrogen flame is contaminated by potassium compounds and so is coloured lilac (a faintly bluish pink). Rubidium Rubidium is denser than water and so sinks. It reacts violently and immediately, with everything spitting out of the container again. Rubidium hydroxide solution and hydrogen are formed. Caesium Caesium explodes on contact with water, quite possibly shattering the container. Caesium hydroxide and hydrogen are formed

Summary of the trend in reactivity The Group 1 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 overall enthalpy changes You might think that because the reactions get more dramatic as you go down the Group, the amount of heat given off increases as you go from lithium to caesium. Not so! The table gives estimates of the enthalpy change for each of the elements undergoing the reaction:

Note: That's the same equation as before, but I have divided it by two to show the enthalpy change per mole of metal reacting.

enthalpy change (kJ / mol) Li Na K Rb Cs -222 -184 -196 -195 -203

You will see that there is no pattern at all in these values. They are all fairly similar and, surprisingly, lithium is the metal which releases the most heat during the reaction!
Note: Apart from the lithium value, I haven't been able to confirm these figures. For lithium, sodium and potassium, they are calculated values based on information in the Nuffield Advanced Science Book of Data (page 114 of my 1984 edition). The lithium value agrees almost exactly with a value I found during a web search. The values for rubidium and caesium are

calculated indirectly from the Li, Na and K values and other information which you will find in a later table on this page.

Digging around in the enthalpy changes When these reactions happen, the differences between them lie entirely in what is happening to the metal atoms present. In each case, you start with metal atoms in a solid and end up with metal ions in solution. Overall, what happens to the metal is this:

You can calculate the overall enthalpy change for this process by using Hess's Law and breaking it up into several steps that we know the enthalpy changes for. First, you would need to supply atomisation energy to give gaseous atoms of the metal.

Then ionise the metal by supplying its first ionisation energy.

And finally, you would get hydration enthalpy released when the gaseous ion comes into contact with water.

Note: There is no suggestion that the reaction actually happens by this route. All we are doing is inventing an imaginary route from the start to the end point of the reaction, and using Hess's Law to say that the overall enthalpy change will be exactly the same as we can calculate using this imaginary route. If you don't know about Hess's Law, you probably aren't likely to be making much sense of all this bit of the page anyway. If you aren't happy about enthalpy changes, you might want to explore the energetics section of Chemguide, or my chemistry calculations book.

If we put values for all these steps into a table, they look like this (all values in kJ / mol):

at. energy Li Na K Rb Cs +161 +109 +90 +86 +79

1st IE +519 +494 +418 +402 +376

hydr. enthalpy -519 -406 -322 -301 -276

total +161 +197 +186 +187 +179

Note: Remember that these aren't the overall enthalpy changes for the reactions when the metal reacts with water. They are only for that part of the reaction which involves the metal. There are also changes going on with the water present - turning it into hydrogen gas and hydroxide ions. To get the total enthalpy changes, you would have to add these values in as well. The changes due to the water will, however, be the same for each reaction in each case about -382 kJ / mol. Adding that on to the figures in this table gives the values in the previous one to within a kJ or two. The rubidium and caesium values will agree exactly, because that's how I had to calculate them in the first table. The other three in the previous table were calculated from information from a different source.

So why isn't there any pattern in these values? If you look at the various bits of information, you will find that as you go down the Group each of them decreases:
y

The atomisation energy is a measure of the strength of the metallic bond in each element. This is falling as the atom gets bigger and the metallic bond is getting longer. The delocalised electrons are further from the attraction of the nuclei in the bigger atoms. The first ionisation energy is falling because the electron being removed is getting more distant from the nucleus. The extra protons in the nucleus are screened by additional layers of electrons. The hydration enthalpy is a measure of the attraction between the metal ions and lone pairs on water molecules. As the ions get bigger, the water molecules are further from the attraction of the nucleus. The extra protons in the nucleus are again screened by the extra layers of

electrons. What is happening is that the various factors are falling at different rates. That destroys any overall pattern. It is, however, possible to look at the table again and find a pattern which is useful.

Looking at the activation energies for the reactions Let's take the last table and just look at the energy input terms - the two processes where you have to supply energy to make them work. In other words, we will miss out the hydration enthalpy term and just add up the other two. at. energy Li Na K Rb Cs +161 +109 +90 +86 +79 1st IE +519 +494 +418 +402 +376 total +680 +603 +508 +488 +455

Now you can see that there is a steady fall as you go down the Group. As you go from lithium to caesium, you need to put less energy into the reaction to get a positive ion formed. This energy will be recovered later on (plus quite a lot more!), but has to be supplied initially. This is going to be related to the activation energy of the reaction. The lower the activation energy, the faster the reaction. So although lithium releases most heat during the reaction, it does it relatively slowly - it isn't all released in one short, sharp burst. Caesium, on the other hand, has a significantly lower activation energy, and so although it doesn't release quite as much heat overall, it does it extremely quickly - and you get an explosion.

Note: You need to be a bit careful about how you phrase this! You probably haven't noticed my use of the phrase "This is going to be related to the activation energy of the reaction." In rewriting it, I have emphasised the words "related to". The reaction certainly won't involve exactly the energy terms we are talking about. The metal won't first convert to gaseous atoms which then lose an electron. But at some point, atoms will have to break away from the metal structure and they will have to lose electrons. However, other energy releasing processes may happen at exactly the same time - for example, if the metal atom loses an electron, something almost certainly picks it up simultaneously. The electron is never likely to be totally free. That will have the effect of reducing the height of the real activation energy barrier. The values we have calculated by adding up the atomisation and ionisation energies are very big in activation energy terms and the reactions would be extremely slow if they were for real.

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 in part due to a decrease in ionisation energy as you go down the Group, and in part to a fall in atomisation energy reflecting weaker metallic bonds as you go from lithium to caesium. This leads to lower activation energies, and therefore faster reactions. REACTIONS OF THE GROUP 1 ELEMENTS WITH OXYGEN AND CHLORINE

This page mainly looks at the reactions of the Group 1 elements (lithium, sodium, potassium, rubidium and caesium) with oxygen - including the simple reactions of the various kinds of oxides formed. It also deals very briefly with the reactions of the elements with chlorine.

The Reactions with Air or Oxygen General These are all very reactive metals and have to be stored out of contact with air

to prevent their oxidation. Reactivity increases as you go down the Group. Lithium, sodium and potassium are stored in oil. (Lithium in fact floats on the oil, but there will be enough oil coating it to give it some protection. It is, anyway, less reactive than the rest of the Group.) Rubidium and caesium are normally stored in sealed glass tubes to prevent air getting at them. They are stored either in a vacuum or in an inert atmosphere of, say, argon. The tubes are broken open when the metal is used. Depending on how far down the Group you are, different kinds of oxide are formed when the metals burn (details below). Reaction with oxygen is just a more dramatic version of the reaction with air. Lithium is unique in the Group because it also reacts with the nitrogen in the air to form lithium nitride (again, see below).

Details for the individual metals Lithium Lithium burns with a strongly red-tinged flame if heated in air. It reacts with oxygen in the air to give white lithium oxide. With pure oxygen, the flame would simply be more intense.

For the record, it also reacts with the nitrogen in the air to give lithium nitride. Lithium is the only element in this Group to form a nitride in this way.

Note: You will find the reason why lithium forms a nitride on the page about reactions of Group 2 elements with air or oxygen. Lithium's reactions are often rather like those of the Group 2 metals. You will find what you want about 3/4 of the way down that page. If you choose to follow this link (more out of interest than because it is essential for UK A level purposes), use the BACK button on your browser to return to this page later.

Sodium Small pieces of sodium burn in air with often little more than an orange glow. Using larger amounts of sodium or burning it in oxygen gives a strong orange flame. You get a white solid mixture of sodium oxide and sodium peroxide. The equation for the formation of the simple oxide is just like the lithium one.

The peroxide equation is:

Potassium Small pieces of potassium heated in air tend to just melt and turn instantly into a mixture of potassium peroxide and potassium superoxide without any flame being seen. Larger pieces of potassium burn with a lilac flame. The equation for the formation of the peroxide is just like the sodium one above:

. . . and for the superoxide:

Note: Potassium peroxide and superoxide are described as being somewhere between yellow and orange depending on what source you look at. I have a bit of a problem with this, because over my teaching career I have heated potassium in air many times and, if memory serves correctly, it always leaves a greyish white film on the bit of porcelain you are heating it on. I don't recall ever seeing it yellow or orange! The formula for a peroxide doesn't look too stange, because most people are familiar with the similar formula for hydrogen peroxide. The formula for a superoxide always looks wrong! There is more about these oxides later on.

Rubidium and caesium Both metals catch fire in air and produce superoxides, RbO2 and CsO2. The equations are the same as the equivalent potassium one.
Note: In a lifetime in teaching chemistry, I have never actually handled (or even seen in real life!) either of these metals. I haven't even seen video or film clips of them being burnt. That means that I don't have much confidence in this next bit.

Both superoxides are described in most sources as being either orange or yellow. One major web source describes rubidium superoxide as being dark brown on one page and orange on another! I don't know what the flames look like either. You can't necessarily be sure that the flame that a metal burns with will be the same as the flame colour of its compounds.

Why are different oxides formed as you go down the Group?


y y y

Lithium (and to some extent sodium) form simple oxides, X2O, which contain the common O2- ion. Sodium (and to some extent potassium) form peroxides, X2O2, containing the more complicated O22- ion (discussed below). Potassium, rubidium and caesium form superoxides, XO2. The structure of the superoxide ion, O2-, is too difficult to discuss at this level, needing a good knowledge of molecular orbital theory to make sense of it.

The more complicated ions aren't stable in the presence of a small positive ion. Consider the peroxide ion, for example. The peroxide ion, O22- looks like this:

The covalent bond between the two oxygen atoms is relatively weak.

Now imagine bringing a small positive 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.
Note: A high charge density simply means that you have a lot of charge packed into a small volume.

Even though it only has one charge, the lithium ion at the top of the Group is so small and has such a high charge density that any peroxide ion near it falls to pieces to give an oxide and oxygen. As you go down the Group to sodium and potassium the positive ions get bigger and they don't have so much effect on the peroxide ion. The superoxide ions are even more easily pulled apart, and these are only stable in the presence of the big ions towards the bottom of the Group. So why do any of the metals form the more complicated oxides? It is a matter of energetics. In the presence of sufficient oxygen, they produce the compound whose formation gives out most energy. That gives the most stable compound. The amount of heat evolved per mole of rubidium in forming its various oxides is: enthalpy change (kJ / mol of Rb) Rb2O -169.5

Rb2O2 RbO2

-236 -278.7

Note: These figures are based on a thermodynamic properties table from Gazi University in Turkey. It was the only place I could track down a value for the enthalpy of formation of rubidium superoxide. The enthalpy of formation values for rubidium oxide and peroxide have been divided by two to give results per mole of rubidium in order to make them comparable with the superoxide value.

The values for the various potassium oxides show exactly the same trends. As long as you have enough oxygen, forming the peroxide releases more energy per mole of metal than forming the simple oxide. Forming the superoxide releases even more. I assume the same thing to be true of the caesium oxides, although I couldn't find all the figures to be able to check it.

Summary Forming the more complicated oxides from the metals releases more energy and makes the system more energetically stable. BUT . . . this only works for the metals in the lower half of the Group where the metal ions are big and have a low charge density. At the top of the Group, the small ions with a higher charge density tend to polarise the more complicated oxide ions to the point of destruction.

Reactions of the Oxides


The simple oxides, X2O Reaction with water These are simple basic oxides, reacting with water to give the metal hydroxide. For example, lithium oxide reacts with water to give a colourless solution of

lithium hydroxide.

Note: I'm going to use "X" for all the rest of the equations in this section. There is no difference between the equations for the various elements in the Group whichever metal oxide (or peroxide or superoxide) you are using.

Reaction with dilute acids These simple oxides all react with an acid to give a salt and water. For example, sodium oxide will react with dilute hydrochloric acid to give colourless sodium chloride solution and water.

The peroxides, X2O2 Reaction with water If the reaction is done ice cold (and the temperature controlled so that it doesn't rise even though these reactions are strongly exothermic), a solution of the metal hydroxide and hydrogen peroxide is formed.

If the temperature increases (as it inevitably will unless the peroxide is added to water very, very, very slowly!), the hydrogen peroxide produced decomposes into water and oxygen. The reaction can be very violent overall.

Reaction with dilute acids These reactions are even more exothermic than the ones with water. A solution containing a salt and hydrogen peroxide is formed. The hydrogen peroxide will decompose to give water and oxygen if the temperature rises again, it is almost impossible to avoid this. Another potentially violent reaction!

The superoxides, XO2 Reaction with water This time, a solution of the metal hydroxide and hydrogen peroxide is formed, but oxygen gas is given off as well. Once again, these are strongly exothermic reactions and the heat produced will inevitably decompose the hydrogen peroxide to water and more oxygen. Again violent!

Reaction with dilute acids Again, these reactions are even more exothermic than the ones with water. A solution containing a salt and hydrogen peroxide is formed together with oxygen gas. The hydrogen peroxide will again decompose to give water and oxygen as the temperature rises. Violent!

The Reactions of the elements with Chlorine


This is included on this page because of the similarity in appearance between the reactions of the Group 1 metals with chlorine and with oxygen. Sodium, for example, burns with an intense orange flame in chlorine in exactly the same way that it does in pure oxygen. The rest also behave the same in both gases. In each case, there is a white solid residue which is the simple chloride, XCl. There is nothing in any way complicated about these reactions!

OME COMPOUNDS OF THE GROUP 1

ELEMENTS
This page looks at some compounds of the Group 1 elements (lithium, sodium, potassium, rubidium and caesium) - limited to various bits and pieces required by various UK A level syllabuses. You will find some information about the nitrates, carbonates, hydrogencarbonates and hydrides of the metals. We will first look at what happens to some of the compounds on heating, and then their solubility. At the end, you will find a section about the preparation and reactions of the metal hydrides.

The effect of heat on Group 1 compounds


The facts Group 1 compounds are more stable to heat than the corresponding compounds in Group 2. You will often find that the lithium compounds behave similarly to Group 2 compounds, but the rest of Group 1 are in some way different.

Heating the nitrates Most nitrates tend to decompose on heating to give the metal oxide, brown fumes of nitrogen dioxide, and oxygen. For example, a typical Group 2 nitrate like magnesium nitrate decomposes like this:

In Group 1, lithium nitrate behaves in the same way - producing lithium oxide, nitrogen dioxide and oxygen.

The rest of the Group, however, don't decompose so completely (at least not at Bunsen temperatures) - producing the metal nitrite and oxygen, but no nitrogen dioxide.

All the nitrates from sodium to caesium decomposes in this same way, the only difference being how hot they have to be to undergo the reaction. As you go down the Group, the decomposition gets more difficult, and you have to use higher temperatures.
Note: The more modern name for sodium nitrite is sodium nitrate(III). On this basis, sodium nitrate should properly be called sodium nitrate(V). Most people still call nitrates and nitrites by the older names.

Heating the carbonates Most carbonates tend to decompose on heating to give the metal oxide and carbon dioxde. For example, a typical Group 2 carbonate like calcium carbonate decomposes like this:

In Group 1, lithium carbonate behaves in the same way - producing lithium oxide and carbon dioxide.

The rest of the Group 1 carbonates don't decompose at Bunsen temperatures, although at higher temperatures they will. The decomposition temperatures again increase as you go down the Group.
Note: I have severe problems with this - and what I have said is in line with what UK examiners are likely to expect, but whether it is the truth, I don't know! Various data sources give a decomposition temperature for lithium carbonate as 1310C - well above Bunsen temperatures (about 1000C maximum if something is heated directly with no glass getting in the way). Heslop and Robinson's Inorganic Chemistry (my copy published in 1960) says that it will decompose on heating in a stream of hydrogen at 800C. I'm not sure what the purpose of the hydrogen is. If it was simply to sweep away the carbon dioxide to prevent it recombining with the oxide, it seems an unnecessarily hazardous way of doing it! It is also difficult to get reliable results if you heat these carbonates in the lab. They all tend to react with water vapour and carbon dioxide in the air to produce hydrogencarbonates - and these decompose easily on heating,

releasing the carbon dioxide again. Therefore heating a normal lab sample of, say, sodium carbonate does often produce some carbon dioxide because of this contamination. It is difficult to say categorically that no carbon dioxide is being produced from the sodium carbonate.

The thermal stability of the hydrogencarbonates The Group 2 hydrogencarbonates like calcium hydrogencarbonate are so unstable to heat that they only exist in solution. Any attempt to get them out of solution causes them to decompose to give the carbonate, carbon dioxide and water.

By contrast, the Group 1 hydrogencarbonates are stable enough to exist as solids, although they do decompose easily on heating. For example, for sodium hydrogencarbonate:

Note: There is complete disagreement in various sources about lithium hydrogencarbonate. Some say it only exists in solution; some quote it as a solid. The only reasonably definitive information I managed to track down was from the Handbook of Inorganic Compounds edited by Perry and Phillips. This quotes a colour for lithium hydrogencarbonate as "white", and a solubility in water of 5.5g per 100 ml at 13C. Both of these statements imply to me that it is a solid.

Explanations for the trends in thermal stability Detailed explanations are given for the carbonates because the diagrams are easier to draw. Exactly the same arguments apply to the nitrates or hydrogencarbonates. There are two ways of explaining the increase in thermal stability as you go down the Group. The hard way is in terms of the energetics of the process; the simple way is to look at the polarising ability of the positive ions.
Note: The UK A level syllabuses which talk about Group 1 chemistry only want the simpler way. If you are interested in the energetics arguments, you will find them discussed at length for Group 2 compounds by following this link. Be prepared for some seriously hard work!

The explanation below on the polarising ability of the positive ions is taken from that page with only minor modifications.

Explaining the trend in terms of the polarising ability of the positive ion A small positive ion has a lot of charge packed into a small volume of space especially if it has more than one positive charge. 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 positive ion has the same charge spread over a larger volume of 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-and-crosses" 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 aredelocalised. 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 tobenzene or to organic acids. Either of these links is likely to involve you in a fairly timeconsuming detour!

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. The diagram shows what happens with an ion from Group 2, carrying two positive charges

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. If the positive ion only had one positive charge, the polarising effect would be less. That is why the Group 1 compounds are more thermally stable than those in Group 2. You have to heat the Group 1 compound more because the carbonate ions are less polarised by singly charged positive ions.

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. In other words, as you go down the Group, the carbonates become more thermally stable.

What about the nitrates and hydrogencarbonates? The argument is exactly the same here. The small positive ions at the top of the Group polarise the nitrate or hydrogencarbonate ions more than the larger positive ions at the bottom. And, again, the Group 1 compounds will need to be heated more strongly than those in Group 2 because the Group 1 ions are less polarising.
Note: The reason for drawing the diagrams for a 2+ ion polarising a carbonate ion is that they are much easier than any other combination. For everything else you have more complicated interactions involving more than one positive or negative ion. The principle is the same - it's just a lot more difficult to make it easy to understand because the diagrams would be so confusing. Don't worry about this. For UK A level purposes all you would need to do is talk about how the polarising ability of the positive ion increases as it gets smaller or more charged. The diagrams and lengthy explanation above are just to help you to understand what that means.

The solubility of Group 1 compounds


The facts For UK A level purposes, the important thing to remember is that Group 1 compounds tend to be more soluble than the corresponding ones in Group 2.

The carbonates

For example, Group 2 carbonates are virtually insoluble in water. 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. By contrast, the least soluble Group 1 carbonate is lithium carbonate. A saturated solution of it has a concentration of about 1.3 g per 100 g of water at 20C. The other carbonates in the Group all count as very soluble - increasing to an astonishing 261.5 g per 100 g of water at this temperature for caesium carbonate. Solubility of the carbonates increases as you go down Group 1.

The hydroxides The least soluble hydroxide in Group 1 is lithium hydroxide - but it is still possible to make a solution with a concentration of 12.8 g per 100 g of water at 20C. The other hydroxides in the Group are even more soluble. Solubility of the hydroxides increases as you go down Group 1. In Group 2, the most soluble one is barium hydroxide - and it is only possible to make a solution of concentration around 3.9 g per 100 g of water at the same temperature.

I'm not even going to attempt an explanation of these trends! Trying to explain trends in solubility is a complete nightmare. If you have read the section on Group 2 of the Periodic Table, you may know that I have shown why the usual explanations given for these trends at this level don't work.
Note: If you are an absolute glutton for punishment, you can read about this by following this link to the page about why the normal explanations for Group 2 solubility trends don't work! Don't waste time doing this unless you know about entropy.

Explaining the trends in Group 2 was difficult enough. Comparing them with Group 1 is going to be even more difficult - particularly in the case of the carbonates, because the trends in the two Groups are in opposite directions. The carbonates get more soluble as you go down Group 1, but tend to get less soluble down Group 2.

This is too difficult to talk about at this level - and I'm not going to do it! You should not need it for UK A level purposes for Group 1. Just learn that Group 1 compounds tend to be more soluble than their Group 2 equivalents.

The Group 1 hydrides


Saline (salt-like) hydrides The hydrides of Group 1 metals are white crystalline solids which contain the metal ions and hydride ions, H-. They have exactly the same crystal structure as sodium chloride - that's why they are called saline or salt-like hydrides. Because they can react violently with water or moist air, they are normally supplied as suspensions in mineral oil.
Note: You will find the crystal structure of sodium chloride if you follow this link. Use the BACK button on your browser to return to this page.

Preparation of the Group 1 hydrides These are made by passing hydrogen gas over the heated metal. For example, for lithium hydride:

Reactions of the Group 1 hydrides These are limited to the two reactions most likely to be wanted by UK A level syllabuses. Electrolysis On heating, most of these hydrides decompose back into the metal and hydrogen before they melt. It is, however, possible to melt lithium hydride and to electrolyse the melt. The metal is released at the cathode as you would expect. Hydrogen is given

off at the anode (the positive electrode) and this is evidence for the presence of the negative hydride ion in lithium hydride. The anode equation is:

The other Group 1 hydrides can be electrolysed in solution in various molten mixtures such as a mixture of lithium chloride and potassium chloride. Mixtures such as these melt at lower temperatures than the pure chlorides. Reaction with water These hydrides react violently with water releasing hydrogen gas and producing the metal hydroxide. For example, sodium hydride reacts with water to produce a solution of sodium hydroxide and hydrogen gas.

The colours The colours in the table are just a guide. Almost everybody sees and describes colours differently. I have, for example, used the word "red" several times to describe colours which can be quite different from each other. Other people use words like "carmine" or "crimson" or "scarlet", but not everyone knows the differences between these words - particularly if their first language isn't English.

flame colour Li Na K Rb Cs red strong persistent orange lilac (pink) red (reddish-violet) blue? violet? (see below)

Ca Sr Ba Cu Pb

orange-red red pale green blue-green (often with white flashes) greyish-white

The origin of flame colours


Flame colours are produced from the movement of the electrons in the metal ions present in the compounds. For example, a sodium ion in an unexcited state has the structure 1s22s22p6. When you heat it, the electrons gain energy and can jump into any of the empty orbitals at higher levels for example, into the 7s or 6p or 4d or whatever, depending on how much energy a particular electron happens to absorb from the flame. Because the electrons are now at a higher and more energetically unstable level, they tend to fall back down to where they were before - but not necessarily all in one go. An electron which had been excited from the 2p level to an orbital in the 7 level, for example, might jump back to the 2p level in one go. That would release a certain amount of energy which would be seen as light of a particular colour.

TOMIC 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 physical properties. Even if you aren't currently interested in all these things, it would probably pay you to read most of this page. The same ideas tend to recur throughout the atomic properties, and you may find that earlier explanations help to you understand later ones. The physical properties

are extremely difficult to explain, however, and you might not want to read about those unless you have to.

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.

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
y y

the number of layers of electrons around the nucleus the pull the outer electrons feel from the nucleus.

Compare beryllium and magnesium: Be Mg 1s22s2 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.

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.

Trends in First Ionisation Energy


First ionisation energy is the energy needed to remove the most loosely held electron from each of 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
y y y

the charge on the nucleus, 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.

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 appear to have an electronegativity of 1.0. This is probably most easily explained by the fact that electronegativities are often only recorded to 1 decimal place. To two decimal places, calcium is 1.00 and strontium is 0.95. When these numbers are rounded to 1 decimal place, both would appear to have an electronegativity of 1.0.

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.

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, Boiling Point, and Atomisation Energy


The facts Melting points

You will see that (apart from where the smooth trend is broken by magnesium) the melting point falls as you go down the Group. Boiling points

You will see that there is no obvious pattern in boiling points. It would be quite wrong to suggest that there is any trend here whatsoever. 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. It looks similar to, but notexactly the same as, the boiling point chart.
Warning! Unless you are doing a syllabus which expects you to be able to explain this, or unless your chemistry is good and you like a good argument, don't read on beyond this point.

Trying to explain this The only explanations you are likely to ever come across relate to the melting points. I will give you the most common explanation (the one required by what I think is the only UK A level syllabus to expect you to know it - AQA), and then explain why I think it is completely wrong! The faulty explanation All of these elements are held together by metallic bonds. The melting points get lower as you go down the Group because the metallic bonds get weaker. The oddity of magnesium has to be explained separately. The atoms in a metal are held together by the attraction of the nuclei to the delocalised electrons. As the atoms get bigger, the nuclei get further away from these delocalised electrons, and so the attractions fall. That means that the atoms are more easily separated to make a liquid and finally a gas. As you go down the Group, the arrangement of the atoms in the various solid metals changes. Beryllium and magnesium are both hexagonal close-packed; calcium and strontium are facecentred cubic; barium is body-centred cubic. Don't worry if you don't know what this means. All that matters is that there is a change in crystal structure between magnesium and calcium. That is supposed to account for the fact that magnesium is out of line with the rest of the Group.

Why I don't believe this explanation The odd position of magnesium Let's take this first, because that argument is relatively easy to demolish. Despite the fact that the first four elements have two different structures, those structures are both 12-co-ordinated. Each atom is touched by 12 surrounding atoms. In that case, you would expect the metallic bond to be similar in each case, because the orbitals are going to overlap and delocalise in the same sort of way. Any differences just due to the structures should only be minor. By contrast, barium is 8-co-ordinated (like the Group 1 metals). That's a less efficient packing, and you might expect that to be reflected in a much weaker metallic bond. Although the barium melting point is lower than that of strontium, it isn't dramatically lower. It just follows the general trend - suggesting that the major change of structure isn't making much difference. You can't have it both ways! If a minor change of structure at magnesium-calcium makes a huge difference, then a major one at barium should make an even bigger difference. It

obviously doesn't. The strength of the metallic bonds Melting point isn't a good guide to the strength of the metallic bonds. When a metal melts, the bonds aren't completely broken - only loosened enough for the atoms to move around. Metallic bonds are still present in the molten metal, and aren't entirely broken until it boils. That means that boiling point, or the size of the atomisation energy, is a much better guide to the real strengths of the metallic bonds. With both of those measures, you are ending up with free atoms in the gas state with the metallic bond completely broken. Cotton and Wilkinson, in their classic degree level book Advanced Inorganic Chemistry say "The strength of binding between the atoms in metals can conveniently be measured by the energies of atomization of the metallic elements." (Third edition, page 68.) If you look back at the atomisation energy chart above, you will see that magnesium still has the lowest value, but there is no obvious trend in atomisation energies as you go down the Group. The explanation about weaker metallic bonds as you go down the Group can't be accurate either. If you look at figures for Group 1 rather than Group 2, then the trends for all the various measures (melting point, boiling point and atomisation energy) work almost perfectly as you go down the Group. There is obviously something happening in Group 2 which is causing the problem. I have no idea at all what it might be.

A final comment I have had a request for solid information about this on Chemguide since 2002, during which time this page will have been read by hundreds of thousands, if not millions, of visitors. In all that time, nobody has suggested an explanation which would account for the low melting point value for magnesium, or the lack of any pattern with the other two properties. If you can see flaws in what I have said above, please get in touch with me. I would also be grateful to anyone who could point me towards an explanation, even if it is too difficult to use at this level, or even too difficult for me to understand. But that explanation has to be capable of accounting for all the variations in the data. There is one book that I have come across which is honest enough to admit the difficulty. A.G.Sharpe, in his degree level bookInorganic Chemistry admits that there is no easy explanation for the variations in the physical data in Group 2. If that is indeed the case, as looks pretty likely, it is a pity that Exam Boards should encourage faulty explanations like the one above. Much better to have no explanation than a deeply flawed one.

EACTIONS 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 white magnesium oxide and hydrogen gas. Very clean magnesium ribbon has a very slight reaction with cold water. After several minutes, some bubbles of hydrogen form on its surface, and the coil of magnesium ribbon usually floats to the surface. However, 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 cold water, you get the metal hydroxide. If it reacts with steam, the metal oxide is formed. This is because the metal hydroxides thermally decompose (split up on heating) to give the oxide and 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. Calcium, for example, reacts fairly vigorously with cold water in an exothermic reaction. Bubbles of hydrogen gas are given off, and a white precipitate (of calcium hydroxide) is formed, together with an alkaline solution (also of calcium hydroxide - calcium hydroxide is slightly soluble). 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


Beryllium as a special case There is an additional reason for the lack of reactivity of beryllium compared with the rest of the Group. Beryllium has a strong resistant layer of oxide on its surface which lowers its reactivity. (This is just like the aluminium case that you are probably familiar with.) If you add that to the trends explained below, beryllium turns out to be very unreactive.

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 enthalpy changes, you might want to explore the energetics section of Chemguide, or my chemistry calculations book.

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.
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:
y y

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.

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. 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!

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 cando 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 the metals look like when they burn is a bit problematical!
y

y y y

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 only seen this burn on video. It is also reluctant to start burning, but then burns with an intense almost white flame with red tinges especially around the outside. Barium: I have also 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.

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 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.
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 avery 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.

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.
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 N3- ions. 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 N3ions. 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.
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.

OLUBILITY 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 UK 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
y

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 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


y

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
y

The carbonates tend to become less soluble as you go down the Group.

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.

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 gas. Thermal decomposition is the term given to splitting up a compound by heating it. All of these carbonates are white solids, and the oxides that are produced are also white solids. 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.
y

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. The nitrates are white solids, and the oxides produced are also white solids. Brown nitrogen dioxide gas is given off together with oxygen. Magnesium and calcium nitrates normally have water of crystallisation, and the solid may dissolve in its own water of crystallisation to make a colourless solution before it starts to decompose. Again, if "X" represents any one of the elements:

As you go down the Group, the nitrates also have to be heated more strongly before they will decompose.
y

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 are working towards a UK-based exam (A level or its equivalent) and 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 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-and-crosses" 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 aredelocalised. 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 tobenzene or to organic acids. Either of these links is likely to involve you in a fairly timeconsuming detour!

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. 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 decomposition 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.

Note: If you aren't happy about enthalpy changes, you might want to explore the energetics section of Chemguide, or my chemistry calculations book.

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 CaCO3 SrCO3 BaCO3

+117 +178 +235 +267

You can see that the reactions become more endothermic as you go 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. 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 UK 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:

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.

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. 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+ Ca2+ O2CO320.065 0.099 0.140 ?

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%. 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.

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.

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
y y

the number of layers of electrons around the nucleus the pull the outer electrons feel from the nucleus.

Compare fluorine and chlorine: F Cl 2,7 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.

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.

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. 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 -349 kJ mol-1. By convention, the negative

sign shows a release of energy.

The first electron affinities of the Group 7 elements


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.

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

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.
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.

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.

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.

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!

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.

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. 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.

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
y

y y

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


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 oversimplified 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.
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:
y y y y

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.

Note: For the next bit, if you aren't happy about enthalpy changes, you might want to explore the energetics section of Chemguide, or my chemistry calculations book.

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: 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) F Cl Br I +79 +121 +112 +107

electron affinity (kJ mol-1) -328 -349 -324 -295

hydration enthalpy (kJ mol-1) -506 -364 -335 -293

overall (kJ mol1) -755 -592 -547 -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.

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) +25 +29 change in hydration enthalpy (kJ mol-1) +29 +42

going from Cl to Br Br to I

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!

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.
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.

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
y y y

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:

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 I2molecules 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


y y y y

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

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 youneed 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.

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:
y

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:

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 want to explore

theenergetics section of Chemguide, or my chemistry calculations book.

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) F Cl Br I +902 +771 +733 +684 heat needed to remove electron from halide ion (kJ mol-1) +328 +349 +324 +295 heat released in forming halogen molecules (kJ mol-1) -79 -121 -112 -107 sum of these (kJ mol1) +1151 +999 +945 +872

Note: There is likely to be some error in these figures. They come from a variety of sources - some more reliable than others!

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! 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 actuallyeasier 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.

sing 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 FClBrIobservation no precipitate white precipitate very pale cream precipitate very pale yellow precipitate

Confirming the precipitate using ammonia solution

Carrying out the confirmation Ammonia solution is added to the precipitates. original precipitate AgCl

observation precipitate dissolves to give a colourless solution precipitate is almost unchanged using dilute ammonia solution, but dissolves in concentrated ammonia solution to give a colourless solution precipitate is insoluble in ammonia solution of any concentration

AgBr

AgI

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 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 mychemistry 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 AgBr AgI 1.8 x 10-10 7.7 x 10-13 8.3 x 10-17

Note: These figures come from the Chemistry Data Book by Stark and Wallace.

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: ion present FClBrIsteamy acidic fumes (of HF) steamy acidic fumes (of HCl) steamy acidic fumes (of HBr) contaminated with brown bromine vapour Some steamy fumes (of HI), but lots of purple iodine vapour (plus various red colours in the tube)

observation

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.

The Ideal Gas Equation


The ideal gas equation is:

pV = nRT

N INTRODUCTION TO CHEMICAL ENERGETICS


This page deals with the basic ideas about energy changes during chemical reactions, including simple energy diagrams and the terms exothermic and endothermic.

Energy changes during chemical reactions


Obviously, lots of chemical reactions give out energy as heat. Getting heat by burning a fuel is a simple example, but you will probably have come across lots of others in the lab. Other reactions need a continuous supply of heat to make them work. Splitting calcium carbonate into calcium oxide and carbon dioxide is a simple example of this. Any chemical reaction will involve breaking some bonds and making new ones. Energy is needed to break bonds, and is given out when the new bonds are formed. It is very unlikely that these two processes will involve exactly the same amount of energy - and so some energy will either be absorbed or released during a reaction. You will find this discussed in more detail in the page about bond enthalpies.

Simple energy diagrams A reaction in which heat energy is given off is said to beexothermic. A reaction in which heat energy is absorbed is said to beendothermic. You can show this on simple energy diagrams. For an exothermic change:

Notice that in an exothermic change, the products have a lower energy than the reactants. The energy that the system loses is given out as heat. The surroundings warm up.

For an endothermic change:

This time the products have a higher energy than the reactants. The system absorbs this extra energy as heat from the surroundings.

Expressing exothermic and endothermic changes in numbers Here is an exothermic reaction, showing the amount of heat evolved:

This shows that 394 kJ of heat energy are evolved when equation quantities of carbon and oxygen combine to give carbon dioxide. The mol-1 (per mole) refers to the whole equation in mole quantities. How do you know that heat is evolved? That is shown by the negative sign. You always think of the energy change during a reaction from the point of view of the reactants. The reactants (carbon and oxygen) have lost energy during the reaction. When you burn carbon in oxygen, that is the energy which is causing the surroundings to get hotter. And here is an endothermic change:

In this case, 178 kJ of heat are absorbed when 1 mole of calcium carbonate reacts to give 1 mole of calcium oxide and 1 mole of carbon dioxide. You can tell that energy is being absorbed because of the plus sign. A simple energy diagram for the reaction looks like this:

The products have a higher energy than the reactants. Energy has been gained by the system - hence the plus sign. Whenever you write values for any energy change, you must always write a plus or a minus sign in front of it. If you have an endothermic change, and write, say, 178 kJ mol-1 instead of +178 kJ mol-1, you risk losing a mark in an exam.

Energetic stability

You are likely to come across statements that say that something is energetically more stable than something else. For example, in the next page in this section you will find that I have said that oxygen, O2, is more energetically stable than ozone, O3. What does this mean? If you plot the positions of oxygen and ozone on an energy diagram, it looks like this:

The lower down the energy diagram something is, the more energetically stable it is. If ozone converted into ordinary oxygen, heat energy would be released, and the oxygen would be in a more energetically stable form than it was before. So why doesn't ozone immediately convert into the more energetically stable oxygen? Similarly, if you mix petrol (gasoline) and air at ordinary temperatures (when you are filling up a car, for example), why doesn't it immediately convert into carbon dioxide and water? It would be much more energetically stable if it turned into carbon dioxide and water - you can tell that, because lots of heat is given out when petrol burns in air. But there is no reaction when you mix the two. For any reaction to happen, bonds have to be broken, and new ones made. Breaking bonds takes energy. There is a minimum amount of energy needed before a reaction can start - activation energy. If the molecules don't, for example, hit each other with enough energy, then nothing happens. We say that the mixture iskinetically stable, even though it may be energetically unstablewith respect to its possible products. So a petrol and air mixture at ordinary temperatures doesn't react, even though a lot of energy would be released if the reaction took place. Petrol and air are energetically unstable with respect to carbon dioxide and water - they are much higher up the energy diagram. But a petrol and air mixture is kinetically stable at ordinary temperatures, because the activation energy barrier is too high. If you expose the mixture to a flame or a spark, then you get a major fire or explosion. The initial flame supplies activation energy. The heat given out by the molecules that react first is more than enough to supply the activation energy for the next molecules to react - and so on. The moral of all this is that you should be very careful using the word "stable" in chemistry!

VARIOUS ENTHALPY CHANGE DEFINITIONS


This page explains what an enthalpy change is, and then gives a definition and brief comment for three of the various kinds of enthalpy change that you will come across. You will find some more definitions on other pages in this section. It is essential that you learn the definitions. You aren't going to be able to do any calculations successfully if you don't know exactly what all the terms mean.

Enthalpy changes
Enthalpy change is the name given to the amount of heat evolved or absorbed in a reaction carried out at constant pressure. It is given the symbol H, read as "delta H".
Note: The term "enthalpy change" only applies to reactions done at constant pressure. That is actually how most lab reactions are done - in tubes or flasks (or whatever) open to the atmosphere, so that the pressure is constant at atmospheric pressure. The phrase "at constant pressure" is an essential part of the definition but, apart from that, you are unlikely to need to worry about it if you are doing a UK-based exam at the equivalent of A level.

Standard enthalpy changes Standard enthalpy changes refer to reactions done understandard conditions, and with everything present in theirstandard states. Standard states are sometimes referred to as "reference states". Standard conditions Standard conditions are:
y

298 K (25C)

y y

a pressure of 1 bar (100 kPa). where solutions are involved, a concentration of 1 mol dm-3

Warning! Standard pressure was originally defined as 1 atmosphere (101.325 kPa), and you will still find that in older books (including my calculations book). At the time of writing (August 2010) there was at least one UK-based syllabus that was still talking in terms of "1 atmosphere". It is essential to check your syllabus to find out exactly what you need to learn.

Standard states For a standard enthalpy change everything has to be present in its standard state. That is the physical and chemical state that you would expect to find it in under standard conditions. That means that the standard state for water, for example, is liquid water, H2O(l) - not steam or water vapour or ice. Oxygen's standard state is the gas, O2(g) - not liquid oxygen or oxygen atoms. For elements which have allotropes (two different forms of the element in the same physical state), the standard state is the most energetically stable of the allotropes. For example, carbon exists in the solid state as both diamond and graphite. Graphite is energetically slightly more stable than diamond, and so graphite is taken as the standard state of carbon. Similarly, under standard conditions, oxygen can exist as O2(simply called oxygen) or as O3 (called ozone - but it is just an allotrope of oxygen). The O2 form is far more energetically stable than O3, so the standard state for oxygen is the common O2(g).

The symbol for standard enthalpy changes The symbol for a standard enthalpy change is H, read as "delta H standard" or, perhaps more commonly, as "delta H nought".
Note: Technically, the "o" in the symbol should have a horizontal line through it, extending out at each side. This is such a bother to produce convincingly without the risk of different computers producing unreliable results, that I shall use the common practice of simplifying it to "o".

Standard enthalpy change of reaction, Hr Remember that an enthalpy change is the heat evolved or absorbed when a reaction takes place at constant pressure. The standard enthalpy change of a reaction is the enthalpy change which occurs when equation quantities of materials react under standard conditions, and with everything in its standard state. That needs exploring a bit. Here is a simple reaction between hydrogen and oxygen to make water:

y y

First, notice that the symbol for a standard enthalpy change of reaction is Hr. For enthalpy changes of reaction, the "r" (for reaction) is often missed off - it is just assumed. The "kJ mol-1" (kilojoules per mole) doesn't refer to any particular substance in the equation. Instead it refers to the quantities of all the substances given in the equation. In this case, 572 kJ of heat is evolved when 2 moles of hydrogen gas react with 1 mole of oxygen gas to form 2 moles of liquid water. Notice that everything is in its standard state. In particular, the water has to be formed as a liquid. And there is a hidden problem! The figure quoted is for the reaction under standard conditions, but hydrogen and oxygen don't react under standard conditions. Whenever a standard enthalpy change is quoted, standard conditions are assumed. If the reaction has to be done under different conditions, a different enthalpy change would be recorded. That has to be calculated back to what it would be under standard conditions. Fortunately, you don't have to know how to do that at this level.

Some important types of enthalpy change

Standard enthalpy change of formation, Hf The standard enthalpy change of formation of a compound is the enthalpy change which occurs when one mole of the compound is formed from its elements under standard conditions, and with everything in its standard state.

Note: When you are trying to learn these definitions, you can make life easier for yourself by picking out the key bit, and adding the other bits on afterwards. The key bit about this definition is that you are forming 1 mole of a compound from its elements. All the stuff about enthalpy change and standard conditions and standard states is common to most of these definitions.

The equation showing the standard enthalpy change of formation for water is:

When you are writing one of these equations for enthalpy change of formation, you must end up with 1 mole of the compound. If that needs you to write fractions on the left-hand side of the equation, that is OK. (In fact, it is not just OK, it is essential, because otherwise you will end up with more than 1 mole of compound, or else the equation won't balance!) The equation shows that 286 kJ of heat energy is given out when 1 mole of liquid water is formed from its elements under standard conditions.

Standard enthalpy changes of formation can be written for any compound, even if you can't make it directly from the elements. For example, the standard enthalpy change of formation for liquid benzene is +49 kJ mol-1. The equation is:

If carbon won't react with hydrogen to make benzene, what is the point of this, and how does anybody know what the enthalpy change is?

What the figure of +49 shows is the relative positions of benzene and its elements on an energy diagram:

How do we know this if the reaction doesn't happen? It is actually very simple to calculate it from other values which we canmeasure - for example, from enthalpy changes of combustion (coming up next). We will come back to this again when we look at calculations on another page. Knowing the enthalpy changes of formation of compounds enables you to calculate the enthalpy changes in a whole host of reactions and, again, we will explore that in a bit more detail on another page.

And one final comment about enthalpy changes of formation: The standard enthalpy change of formation of an element in its standard state is zero. That's an important fact. The reason is obvious . . . For example, if you "make" one mole of hydrogen gas starting from one mole of hydrogen gas you aren't changing it in any way, so you wouldn't expect any enthalpy change. That is equally true of any other element. The enthalpy change of formation of any element has to be zero because of the way enthalpy change of formation is defined.
Note: Some sources say that the enthalpy change of formation of elements is taken as zero by convention. That is simply nonsense! The standard enthalpy change of formation of elements is zero because of the way the enthalpy change is defined. If this confuses you, ignore it!

Standard enthalpy change of combustion, Hc The standard enthalpy change of combustion of a compound is the enthalpy change which occurs when one mole of the compound is burned completely in oxygen under standard conditions, and with everything in its standard state. The enthalpy change of combustion will always have a negative value, of course, because burning always releases heat. Two examples:

Notice:
y y

Enthalpy of combustion equations will often contain fractions, because you must start with only 1 mole of whatever you are burning. If you are talking about standard enthalpy changes of combustion, everything must be in its standard state. One important result of this is that any water you write amongst the products must be there as liquid water. Similarly, if you are burning something like ethanol, which is a liquid under standard conditions, you must show it as a liquid in any equation you use.

Notice also that the equation and amount of heat evolved in the hydrogen case is exactly the same as you have already come across further up the page. At that time, it was illustrating the enthalpy of formation of water. That can happen in some simple cases. Talking about the enthalpy change of formation of water is exactly the same as talking about the enthalpy change of combustion of hydrogen.

HESS'S LAW AND ENTHALPY CHANGE CALCULATIONS

This page explains Hess's Law, and uses it to do some simple enthalpy change calculations involving enthalpy changes of reaction, formation and combustion.

Hess's Law
Stating Hess's Law Hess's Law is the most important law in this part of chemistry. Most calculations follow from it. It says . . . The enthalpy change accompanying a chemical change is independent of the route by which the chemical change occurs.

Explaining Hess's Law Hess's Law is saying that if you convert reactants A into products B, the overall enthalpy change will be exactly the same whether you do it in one step or two steps or however many steps. If you look at the change on an enthalpy diagram, that is actually fairly obvious.

This shows the enthalpy changes for an exothermic reaction using two different ways of getting from reactants A to products B. In one case, you do a direct conversion; in the other, you use a two-step process involving some intermediates. In either case, the overall enthalpy change must be the same, because it is governed by the relative positions of the reactants and products on the enthalpy diagram. If you go via the intermediates, you do have to put in some extra heat energy to start with, but you get it

back again in the second stage of the reaction sequence. However many stages the reaction is done in, ultimately the overall enthalpy change will be the same, because the positions of the reactants and products on an enthalpy diagram will always be the same.
Note: It is possibly confusing that I am switching between the terms enthalpy and energy. Enthalpy change is simply a particular measure of energy change. You will remember that the enthalpy change is the heat evolved or absorbed during a reaction happening at constant pressure. I have labelled the vertical scale on this particular diagram as enthalpy rather than energy, because we are specifically thinking about enthalpy changes. I could have just kept to the more general term "energy", but I prefer to be accurate.

You can do calculations by setting them out as enthalpy diagrams as above, but there is a much simpler way of doing it which needs virtually no thought. You could set out the above diagram as:

Hess's Law says that the overall enthalpy change in these two routes will be the same. That means that if you already know two of the values of enthalpy change for the three separate reactions shown on this diagram (the three black arrows), you can easily calculate the third - as you will see below. The big advantage of doing it this way is that you don't have to worry about the relative positions of everything on an enthalpy diagram. It is completely irrelevant whether a particular enthalpy change is positive or negative. Warnings! Although most calculations you will come across will fit into a triangular diagram like the above, you may also come across other slightly more complex cases needing more steps. That doesn't make it any harder! You need to take care in choosing your two routes. The pattern willnot always look like the one above. You

will see that in the examples below.

Enthalpy change calculations using Hess's Law cycles


I can only give a brief introduction here, because this is covered in careful, step-by-step detail in my chemistry calculations book.

Working out an enthalpy change of formation from enthalpy changes of combustion If you have read an earlier page in this section, you may remember that I mentioned that the standard enthalpy change of formation of benzene was impossible to measure directly. That is because carbon and hydrogen won't react to make benzene.
Important: If you don't know (without thinking about it too much) exactly what is meant by standard enthalpy change of formation or combustion, you must get this sorted out now. Re-read the page about enthalpy change definitions before you go any further - and learn them!

Standard enthalpy changes of combustion, Hc are relatively easy to measure. For benzene, carbon and hydrogen, these are: Hc (kJ mol-1) C6H6(l) C(s) H2(g) First you have to design your cycle.
y

-3267 -394 -286

y y

Write down the enthalpy change you want to find as a simple horizontal equation, and write H over the top of the arrow. (In diagrams of this sort, we often miss off the standard symbol just to avoid clutter.) Then fit the other information you have onto the same diagram to make a Hess's Law cycle, writing the known enthalpy changes over the arrows for each of the other changes. Finally, find two routes around the diagram, always going with the flow of the various arrows. You must never have one of your route arrows going in the opposite direction to one of the equation

arrows underneath it. In this case, what we are trying to find is the standard enthalpy change of formation of benzene, so that equation goes horizontally.

You will notice that I haven't bothered to include the oxygen that the various things are burning in. The amount of oxygen isn't critical because you just use an excess anyway, and including it really confuses the diagram. Why have I drawn a box around the carbon dioxide and water at the bottom of the cycle? I tend to do this if I can't get all the arrows to point to exactly the right things. In this case, there is no obvious way of getting the arrow from the benzene to point at both the carbon dioxide and the water. Drawing the box isn't essential - I just find that it helps me to see what is going on more easily. Notice that you may have to multiply the figures you are using. For example, standard enthalpy changes of combustion start with 1 mole of the substance you are burning. In this case, the equations need you to burn 6 moles of carbon, and 3 moles of hydrogen molecules. Forgetting to do this is probably the most common mistake you are likely to make. How were the two routes chosen? Remember that you have to go with the flow of the arrows. Choose your starting point as the corner that only has arrows leaving from it. Choose your end point as the corner which only has arrows arriving. Now do the calculation: Hess's Law says that the enthalpy changes on the two routes are the same. That means that: H - 3267 = 6(-394) + 3(-286) Rearranging and solving: H = 3267 + 6(-394) + 3(-286) H = +45 kJ mol-1

Note: If you have a good memory, you might remember that I gave a figure of +49 kJ mol-1 for the standard enthalpy change of formation of benzene on an earlier page in this section. So why is this answer different? The main problem here is that I have taken values of the enthalpies of combustion of hydrogen and carbon to 3 significant figures (commonly done in calculations at this level). That introduces small errors if you are just taking each figure once. However, here you are multiplying the error in the carbon value by 6, and the error in the hydrogen value by 3. If you are interested, you could rework the calculation using a value of -393.5 for the carbon and -285.8 for the hydrogen. That gives an answer of +48.6. So why didn't I use more accurate values in the first place? Because I wanted to illustrate this problem! Answers you get to questions like this are often a bit out. The reason usually lies either in rounding errors (as in this case), or the fact that the data may have come from a different source or sources. Trying to get consistent data can be a bit of a nightmare.

Working out an enthalpy change of reaction from enthalpy changes of formation This is the commonest use of simple Hess's Law cycles that you are likely to come across. In this case, we are going to calculate the enthalpy change for the reaction between ethene and hydrogen chloride gases to make chloroethane gas from the standard enthalpy of formation values in the table. If you have never come across this reaction before, it makes no difference. Hf (kJ mol-1) C2H4(g) HCl(g) C2H5Cl(g) +52.2 -92.3 -109

Note: I'm not too happy about the value for chloroethane! The data sources I normally use give a wide range of values. The one I have chosen is an average value from the NIST Chemistry WebBook. This uncertainty doesn't affect how you do the calculation in any way, but the answer may not be exactly right - don't quote it as if it was right.

In the cycle below, this reaction has been written horizontally, and the enthalpy of formation values added to complete the cycle.

Again, notice the box drawn around the elements at the bottom, because it isn't possible to connect all the individual elements to the compounds they are forming in any tidy way. Be careful to count up all the atoms you need to use, and make sure they are written as they occur in the elements in their standard state. You mustn't, for example, write the hydrogens as 5H(g), because the standard state for hydrogen is H2.
Note: In truth, if I am doing this type of enthalpy sum myself (with nobody watching!), I tend to just write the word "elements" in the bottom box to save the bother of working out exactly how many of everything I need. I would be wary of doing that in an exam, though.

And now the calculation. Just write down all the enthalpy changes which make up the two routes, and equate them. +52.2 - 92.3 + H = -109 Rearranging and solving: H = -52.2 + 92.3 - 109 H = -68.9 kJ mol-1 BOND ENTHALPY (BOND ENERGY)

This page introduces bond enthalpies (bond energies) and looks at some simple calculations involving them. One of the most confusing things about this is the way the words are used. These days, the term "bond enthalpy" is normally used, but you will also find it described as "bond energy" - sometimes in the same article. An even older term is "bond strength". So you can take all these terms as being interchangeable. As you will see below, though, "bond enthalpy" is used in several different ways, and you might need to be

careful about this.


Note: Bond enthalpies quoted from different sources often vary by a few kilojoules, even if they are referring to exactly the same thing. Don't worry if you come across slightly different values. As you will see later on this page, calculations involving bond enthalpies hardly ever give accurate answers anyway. You may even find differences in values between different pages of Chemguide, or differences between Chemguide and my calculations book. If so, I apologise, but I tend to use a lot of different data sources which have varied over the years.

Explaining the terms


Bond dissociation enthalpy and mean bond enthalpy Simple diatomic molecules A diatomic molecule is one that only contains two atoms. They could be the same (for example, Cl2) or different (for example, HCl). The bond dissociation enthalpy is the energy needed to break one mole of the bond to give separated atoms - everything being in the gas state. Important! The point about everything being in the gas state isessential. You cannot use bond enthalpies to do calculations directly from substances starting in the liquid or solid state. As an example of bond dissociation enthalpy, to break up 1 mole of gaseous hydrogen chloride molecules into separate gaseous hydrogen and chlorine atoms takes 432 kJ. The bond dissociation enthalpy for the H-Cl bond is +432 kJ mol-1.

More complicated molecules What happens if the molecule has several bonds, rather than just 1? Consider methane, CH4. It contains four identical C-H bonds, and it seems reasonable that they should all have the same bond enthalpy. However, if you took methane to pieces one hydrogen at a time, it needs a different amount of energy to break each of the four C-H bonds. Every time you break a hydrogen off the carbon, the environment of those left behind changes. And the strength of a bond is affected by what else is around it.

In cases like this, the bond enthalpy quoted is an average value. In the methane case, you can work out how much energy is needed to break a mole of methane gas into gaseous carbon and hydrogen atoms. That comes to +1662 kJ and involves breaking 4 moles of C-H bonds. The average bond energy is therefore +1662/4 kJ, which is +415.5 kJ per mole of bonds. That means that many bond enthalpies are actually quoted asmean (or average) bond enthalpies, although it might not actually say so. Mean bond enthalpies are sometimes referred to as "bond enthalpy terms". In fact, tables of bond enthalpies give average values in another sense as well, particularly in organic chemistry. The bond enthalpy of, say, the C-H bond varies depending on what is around it in the molecule. So data tables use average values which will work well enough in most cases. That means that if you use the C-H value in some calculation, you can't be sure that it exactly fits the molecule you are working with. So don't expect calculations using mean bond enthalpies to give very reliable answers. You may well have to know the difference between a bond dissociation enthalpy and a mean bond enthalpy, and you should be aware that the word mean (or average) is used in two slightly different senses. But for calculation purposes, it isn't something you need to worry about. Just use the values you are given.
Important: The rest of this page assumes that you have already read the page about Hess's Law and enthalpy change calculations. If you have come straight to the current page from a search engine, you won't make sense of the way the calculations are set out unless you first read the Hess's Law page.

Finding enthalpy changes of reaction from bond enthalpies


I can only give a brief introduction here, because this is covered in careful, step-by-step detail in my chemistry calculations book.

Cases where everything present is gaseous Remember that you can only use bond enthalpies directly if everything you are working with is in the gas state. Using the same method as for other enthalpy sums We are going to estimate the enthalpy change of reaction for the reaction between carbon monoxide and steam. This is a part of the manufacturing process for hydrogen.

The bond enthalpies are: bond enthalpy (kJ mol-1) C-O in carbon monoxide C-O in carbon dioxide O-H H-H +1077 +805 +464 +436

So let's do the sum. Here is the cycle - make sure that you understand exactly why it is the way it is.

And now equate the two routes, and solve the equation to find the enthalpy change of reaction. H + 2(805) + 436 = 1077 + 2(464) H = 1077 + 2(464) - 2(805) - 436 H = -41 kJ mol-1

Using a short-cut method for simple cases You could do any bond enthalpy sum by the method above - taking the molecules completely to pieces and then remaking the bonds. If you are happy doing it that way, just go on doing it that way. However, if you are prepared to give it some thought, you can save a bit of time - although only in very

simple cases where the changes in a molecule are very small. For example, chlorine reacts with ethane to give chloroethane and hydrogen chloride gases.

(All of these are gases. I have left the state symbols out this time to avoid cluttering the diagram.) It is always a good idea to draw full structural formulae when you are doing bond enthalpy calculations. It makes it much easier to count up how many of each type of bond you have to break and make. If you look at the equation carefully, you can see what I mean by a "simple case". Hardly anything has changed in this reaction. You could work out how much energy is needed to break every bond, and how much is given out in making the new ones, but quite a lot of the time, you are just remaking the same bond. All that has actually changed is that you have broken a C-H bond and a Cl-Cl bond, and made a new C-Cl bond and a new H-Cl bond. So you can just work those out. bond enthalpy (kJ mol-1) C-H Cl-Cl C-Cl H-Cl Work out the energy needed to break C-H and Cl-Cl: +413 + 243 = +656 kJ mol-1 Work out the energy released when you make C-Cl and H-Cl: -346 - 432 = -778 kJ mol-1 So the net change is +656 - 778 = -122 kJ mol-1
Note: Even if you choose not to use this method, it might be a good idea to be aware of it. It is possible to imagine an examiner setting a question which assumes that you will use this method, and therefore doesn't give a particular

+413 +243 +346 +432

bond enthalpy value that you would need if you did it by the longer method. For example, in the case above, you don't actually need to know the C-C bond enthalpy.

Cases where you have a liquid present I have to keep on saying this! Remember that you can only use bond enthalpies directly if everything you are working with is in the gas state. If you have one or more liquids present, you need an extra energy term to work out the enthalpy change when you convert from liquid to gas, or vice versa. That term is the enthalpy change of vaporisation, and is given the symbol Hvap or Hv. This is the enthalpy change when 1 mole of the liquid converts to gas at its boiling point with a pressure of 1 bar (100 kPa). (Older sources might quote 1 atmosphere rather than 1 bar.) For water, the enthalpy change of vaporisation is +41 kJ mol-1. That means that it take 41 kJ to change 1 mole of water into steam. If 1 mole of steam condenses into water, the enthalpy change would be -41 kJ. Changing from liquid to gas needs heat; changing gas back to liquid releases exactly the same amount of heat. To see how this fits into bond enthalpy calculations, we will estimate the enthalpy change of combustion of methane - in other words, the enthalpy change for this reaction:

Notice that the product is liquid water. You cannot apply bond enthalpies to this. You must first convert it into steam. To do this you have to supply 41 kJ mol-1. The bond enthalpies you need are: bond enthalpy (kJ mol-1) C-H O=O +413 +498

C=O in carbon dioxide O-H The cycle looks like this:

+805 +464

This obviously looks more confusing than the cycles we've looked at before, but apart from the extra enthalpy change of vaporisation stage, it isn't really any more difficult. Before you go on, make sure that you can see why every single number and arrow on this diagram is there. In particular, make sure that you can see why the first 4 appears in the expression "4(+464)". That is an easy thing to get wrong. (In fact, when I first drew this diagram, I carelessly wrote 2 instead of 4 at that point!) That's the hard bit done - now the calculation: H + 2(805) + 2(41) + 4(464) = 4(413) + 2(498) H = 4(413) + 2(498) - 2(805) - 2(41) - 4(464) H = -900 kJ mol-1 The measured enthalpy change of combustion is -890 kJ mol-1, and so this answer agrees to within about 1%. As bond enthalpy calculations go, that's a pretty good estimate.

LATTICE ENTHALPY (LATTICE ENERGY)

This page introduces lattice enthalpies (lattice energies) and Born-Haber cycles. Lattice enthalpy and lattice energy are commonly used as if they mean exactly the same thing - you will often find both terms used within the same textbook article or web site, including on university sites. In fact, there is a difference between them which relates to the conditions under which they are calculated. However, the difference is small, and negligible compared with the differing values for lattice enthalpy that you will find from different data sources. Unless you go on to do chemistry at degree level, the difference between the two terms isn't likely to worry you.
Note: While I have been writing this section, the different values for the same piece of data from different data sources has driven me crazy, because there is no easy way of knowing which is the most recent or most accurate data. In the Born-Haber cycles below, I have used numbers which give a consistent answer, but please don't assume that they are necessarily the most accurate ones. If you are doing a course for 16 - 18 year olds, none of this really matters - you just use the numbers you are given. If you use my chemistry calculations book, you will find a slightly different set of numbers. These came from the Chemistry Data Book

edited by Stark and Wallace, published by John Murray. Values from this now fairly old book often differ slightly from more recent sources. Don't worry about this. It doesn't affect the principles in any way. Just don't assume that any bit of data you are given (even by me) is necessarily "right"!

What is lattice enthalpy?


Two different ways of defining lattice enthalpy There are two different ways of defining lattice enthalpy which directly contradict each other, and you will find both in common use. In fact, there is a simple way of sorting this out, but many sources don't use it. I will explain how you can do this in a moment, but first let's look at how the problem arises. Lattice enthalpy is a measure of the strength of the forces between the ions in an ionic solid. The greater the lattice enthalpy, the stronger the forces. Those forces are only completely broken when the ions are present as gaseous ions, scattered so far apart that there is negligible attraction between them. You can show this on a simple enthalpy diagram.

For sodium chloride, the solid is more stable than the gaseous ions by 787 kJ mol-1, and that is a measure of the strength of the attractions between the ions in the solid. Remember that energy (in this case heat energy) is given out when bonds are made, and is needed to break bonds. So lattice enthalpy could be described in either of two ways.
y

You could describe it as the enthalpy change when 1 mole of sodium chloride (or whatever) was formed from its scattered gaseous ions. In other words, you are looking at a downward arrow on the diagram. In the sodium chloride case, that would be -787 kJ mol-1.

Or, you could describe it as the enthalpy change when 1 mole of sodium chloride (or whatever) is broken up to form its scattered gaseous ions. In other words, you are looking at an upward arrow on the diagram. In the sodium chloride case, that would be +787 kJ mol-1.

Both refer to the same enthalpy diagram, but one

looks at it from the point of view of making the lattice, and the other from the point of view of breaking it up. Unfortunately, both of these are often described as "lattice enthalpy".

This is an absurdly confusing situation which is easily resolved. I suggest that you never use the term "lattice enthalpy" without qualifying it.
y

You should talk about "lattice dissociation enthalpy" if you want to talk about the amount of energy needed to split up a lattice into its scattered gaseous ions. For NaCl, the lattice dissociation enthalpy is +787 kJ mol-1.

You should talk about "lattice formation enthalpy" if you want to talk about the amount of energy released when a lattice is formed from its scattered gaseous ions. For NaCl, the lattice formation enthalpy is 787 kJ mol-1.

That immediately removes any possibility of confusion. So . . . ice dissociation enthalpy is the enthalpy change neede 1 mole of solid crystal into its scattered gaseous ions. L tion enthalpies are always positive.

ice formation enthalpy is the enthalpy change when 1 m ystal is formed from its scattered gaseous ions. Lattice n enthalpies are always negative.

Note: Find out which of these versions your syllabus is likely to want you to know (even if they just call it "lattice enthalpy") and concentrate on that one, but be aware of the confusion! Incidentally, if you are ever uncertain about which version is being used, you can tell from the sign of the enthalpy change being discussed. If the sign is positive, for example, it must refer to breaking bonds, and therefore to a lattice dissociation enthalpy.

Factors affecting lattice enthalpy


The two main factors affecting lattice enthalpy are the charges on the ions and the ionic radii (which affects the distance between the ions). The charges on the ions Sodium chloride and magnesium oxide have exactly the same arrangements of ions in the crystal lattice, but the lattice enthalpies are very different.

Note: In this diagram, and similar diagrams below, I am not interested in whether the lattice enthalpy is defined as a positive or a negative number - I am just interested in their relative sizes. Strictly speaking, because I haven't added a sign to the vertical axis, the values are for lattice dissociation enthalpies. If you prefer lattice formation enthalpies, just mentally put a negative sign in front of each number.

You can see that the lattice enthalpy of magnesium oxide is much greater than that of sodium chloride. That's because in magnesium oxide, 2+ ions are attracting 2- ions; in sodium chloride, the attraction is only between 1+ and 1- ions.

The radius of the ions

The lattice enthalpy of magnesium oxide is also increased relative to sodium chloride because magnesium ions are smaller than sodium ions, and oxide ions are smaller than chloride ions. That means that the ions are closer together in the lattice, and that increases the strength of the attractions. You can also see this effect of ion size on lattice enthalpy as you go down a Group in the Periodic Table. For example, as you go down Group 7 of the Periodic Table from fluorine to iodine, you would expect the lattice enthalpies of their sodium salts to fall as the negative ions get bigger - and that is the case:

Attractions are governed by the distances between the centres of the oppositely charged ions, and that distance is obviously greater as the negative ion gets bigger. And you can see exactly the same effect if as you go down Group 1. The next bar chart shows the lattice enthalpies of the Group 1 chlorides.

Note: To save anyone the bother of getting in touch with me to point it out, it's not strictly fair to include caesium chloride in this list. Caesium chloride has a different packing arrangement of ions in its crystal, and that has a small effect on the lattice enthalpy. The effect is small enough that it doesn't actually affect the trend.

Calculating lattice enthalpy


It is impossible to measure the enthalpy change starting from a solid crystal and converting it into its scattered gaseous ions. It is even more difficult to imagine how you could do the reverse - start with scattered gaseous ions and measure the enthalpy change when these convert to a solid crystal. Instead, lattice enthalpies always have to be calculated, and there are two entirely different ways in which this can be done. You can can use a Hess's Law cycle (in this case called a Born-Haber cycle) involving enthalpy changes which can be measured. Lattice

enthalpies calculated in this way are described as experimental values. Or you can do physics-style calculations working out how much energy would be released, for example, when ions considered as point charges come together to make a lattice. These are described as theoretical values. In fact, in this case, what you are actually calculating are properly described as lattice energies.
Note: If you aren't confident about Hess's Law cycles, it isessential that you follow this link before you go on.

Experimental values - Born-Haber cycles Standard atomisation enthalpies Before we start talking about Born-Haber cycles, there is an extra term which we need to define. That is atomisation enthalpy, Ha. ndard atomisation enthalpy is the enthalpy change whe gaseous atoms is formed from the element in its stand nthalpy change of atomisation is always positive. You are always going to have to supply energy to break an element into its separate gaseous atoms. All of the following equations represent changes involving atomisation enthalpy:

Notice particularly that the "mol-1" is per mole of

atoms formed - NOT per mole of element that you start with. You will quite commonly have to write fractions into the left-hand side of the equation. Getting this wrong is a common mistake.

Born-Haber cycles I am going to start by drawing a Born-Haber cycle for sodium chloride, and then talk it through carefully afterwards. You will see that I have arbitrarily decided to draw this for lattice formation enthalpy. If you wanted to draw it for lattice dissociation enthalpy, the red arrow would be reversed - pointing upwards.

Focus to start with on the higher of the two thicker horizontal lines. We are starting here with the elements sodium and chlorine in their standard states. Notice that we only need half a mole of chlorine gas in order to end up with 1 mole of NaCl. The arrow pointing down from this to the lower thick line represents the enthalpy change of formation of sodium chloride. The Born-Haber cycle now imagines this formation of sodium chloride as happening in a whole set of small changes, most of which we know the enthalpy changes for - except, of course, for the

lattice enthalpy that we want to calculate.


y

The +107 is the atomisation enthalpy of sodium. We have to produce gaseous atoms so that we can use the next stage in the cycle. The +496 is the first ionisation energy of sodium. Remember that first ionisation energies go from gaseous atoms to gaseous singly charged positive ions. The +122 is the atomisation enthalpy of chlorine. Again, we have to produce gaseous atoms so that we can use the next stage in the cycle. The -349 is the first electron affinity of chlorine. Remember that first electron affinities go from gaseous atoms to gaseous singly charged negative ions. And finally, we have the positive and negative gaseous ions that we can convert into the solid sodium chloride using the lattice formation enthalpy.

Note: If you have forgotten about ionisation energies orelectron affinities follow these links before you go on.

Now we can use Hess's Law and find two different routes around the diagram which we can equate. As I have drawn it, the two routes are obvious. The diagram is set up to provide two different routes between the thick lines. So, here is the cycle again, with the calculation directly underneath it . . .

-411 = +107 + 496 + 122 - 349 + LE LE = -411 - 107 - 496 - 122 + 349 LE = -787 kJ mol-1
Note: Notice that in the calculation, we aren't making any assumptions about the sign of the lattice enthalpy (despite the fact that it is obviously negative because the arrow is pointing downwards). In the first line of the calculation, I have just written "+ LE", and have left it to the calculation to work out that it is a negative answer.

How would this be different if you had drawn a lattice dissociation enthalpy in your diagram? (Perhaps because that is what your syllabus wants.) Your diagram would now look like this:

The only difference in the diagram is the direction the lattice enthalpy arrow is pointing. It does, of course, mean that you have to find two new routes. You can't use the original one, because that would go against the flow of the lattice enthalpy arrow. This time both routes would start from the elements in their standard states, and finish at the

gaseous ions. -411 + LE = +107 + 496 + 122 - 349 LE = +107 + 496 + 122 - 349 + 411 LE = +787 kJ mol-1 Once again, the cycle sorts out the sign of the lattice enthalpy for you.
Note: You will find more examples of calculations involving Born-Haber cycles in my chemistry calculations book. This includes rather more complicated cycles involving, for example, oxides. If you compare the figures in the book with the figures for NaCl above, you will find slight differences - the main culprit being the electron affinity of chlorine, although there are other small differences as well. Don't worry about this - the values in the book come from an older data source. In an exam, you will just use the values you are given, so it isn't a problem.

Theoretical values for lattice energy Let's assume that a compound is fully ionic. Let's also assume that the ions are point charges - in other words that the charge is concentrated at the centre of the ion. By doing physics-style calculations, it is possible to calculate a theoretical value for what you would expect the lattice energy to be.

And no - I am not being careless about this! Calculations of this sort end up with values of lattice energy, and not lattice enthalpy. If you know how to do it, you can then fairly easily convert between the two. There are several different equations, of various degrees of complication, for calculating lattice energy in this way. You won't be expected to be able to do these calculations at this level, but you might be expected to comment on the results of them. There are two possibilities:
y

There is reasonable agreement between the experimental value (calculated from a Born-Haber cycle) and the theoretical value. Sodium chloride is a case like this - the theoretical and experimental values agree to within a few percent. That means that for sodium chloride, the assumptions about the solid being ionic are fairly good.

The experimental and theoretical values don't agree. A commonly quoted example of this is silver chloride, AgCl. Depending on where you get your data from, the theoretical value for lattice enthalpy for AgCl is about 150 kJ mol-1 less than the value that comes from a Born-Haber cycle. In other words, treating the AgCl as 100% ionic underestimates its lattice enthalpy by quite a lot. The explanation is that silver chloride actually has a significant amount of covalent bonding between the silver and the chlorine, because there isn't enough electronegativity difference between the

two to allow for complete transfer of an electron from the silver to the chlorine. Comparing experimental (Born-Haber cycle) and theoretical values for lattice enthalpy is a good way of judging how purely ionic a crystal is.
Note: If you have forgotten about electronegativity it might pay you to revise it now by following this link.

Why is magnesium chloride MgCl2?


This section may well go beyond what your syllabus requires. Before you spend time on it, check your syllabus (and past exam papers as well if possible) to make sure. The question arises as to why, from an energetics point of view, magnesium chloride is MgCl2 rather than MgCl or MgCl3 (or any other formula you might like to choose). It turns out that MgCl2 is the formula of the compound which has the most negative enthalpy change of formation - in other words, it is the most stable one relative to the elements magnesium and chlorine. Let's look at this in terms of Born-Haber cycles. In the cycles this time, we are interested in working out what the enthalpy change of formation would be for the imaginary compounds MgCl and MgCl3. That means that we will have to use theoretical values of their lattice enthalpies. We can't use experimental ones, because these compounds obviously don't exist!

I'm taking theoretical values for lattice enthalpies for these compounds that I found on the web. I can't confirm these, but all the other values used by that source were accurate. The exact values don't matter too much anyway, because the results are so dramatically clear-cut. We will start with the compound MgCl, because that cycle is just like the NaCl one we have already looked at.

The Born-Haber cycle for MgCl

Find two routes around this without going against the flow of any arrows. That's easy:

Hf = +148 + 738 + 122 - 349 - 753 Hf = -94 kJ mol-1 So the compound MgCl is definitely energetically more stable than its elements. I have drawn this cycle very roughly to scale, but that is going to become more and more difficult as we look at the other two possible formulae. So I am going to rewrite it as a table. You can see from the diagram that the enthalpy change of formation can be found just by adding up all the other numbers in the cycle, and we can do this just as well in a table. kJ atomisation enthalpy of Mg 1st IE of Mg atomisation enthalpy of Cl electron affinity of Cl lattice enthalpy calculated Hf +148 +738 +122 -349 -753 -94

The Born-Haber cycle for MgCl2 The equation for the enthalpy change of formation this time is

So how does that change the numbers in the BornHaber cycle?

You need to add in the second ionisation energy of magnesium, because you are making a 2+ ion. You need to multiply the atomisation enthalpy of chlorine by 2, because you need 2 moles of gaseous chlorine atoms. You need to multiply the electron affinity of chlorine by 2, because you are making 2 moles of chloride ions. You obviously need a different value for lattice enthalpy. kJ

atomisation enthalpy of Mg 1st IE of Mg 2nd IE of Mg atomisation enthalpy of Cl (x 2) electron affinity of Cl (x 2) lattice enthalpy calculated Hf

+148 +738 +1451 +244 -698 -2526 -643

You can see that much more energy is released when you make MgCl2 than when you make MgCl. Why is that? You need to put in more energy to ionise the magnesium to give a 2+ ion, but a lot more energy is released as lattice enthalpy. That is because there are stronger ionic attractions between 1- ions and 2+ ions than between the 1- and 1+ ions in MgCl. So what about MgCl3? The lattice energy here

would be even greater.

The Born-Haber cycle for MgCl3 The equation for the enthalpy change of formation this time is

So how does that change the numbers in the BornHaber cycle this time?
y

You need to add in the third ionisation energy of magnesium, because you are making a 3+ ion. You need to multiply the atomisation enthalpy of chlorine by 3, because you need 3 moles of gaseous chlorine atoms. You need to multiply the electron affinity of chlorine by 3, because you are making 3 moles of chloride ions. You again need a different value for lattice enthalpy. kJ

atomisation enthalpy of Mg 1st IE of Mg 2nd IE of Mg 3rd IE of Mg atomisation enthalpy of Cl (x 3) electron affinity of Cl (x 3) lattice enthalpy

+148 +738 +1451 +7733 +366 -1047 -5440

calculated Hf

+3949

This time, the compound is hugely energetically unstable, both with respect to its elements, and also to other compounds that could be formed. You would need to supply nearly 4000 kJ to get 1 mole of MgCl3 to form! Look carefully at the reason for this. The lattice enthalpy is the highest for all these possible compounds, but it isn't high enough to make up for the very large third ionisation energy of magnesium. Why is the third ionisation energy so big? The first two electrons to be removed from magnesium come from the 3s level. The third one comes from the 2p. That is closer to the nucleus, and lacks a layer of screening as well - and so much more energy is needed to remove it. The 3s electrons are screened from the nucleus by the 1 level and 2 level electrons. The 2p electrons are only screened by the 1 level (plus a bit of help from the 2s electrons).

Conclusion Magnesium chloride is MgCl2 because this is the combination of magnesium and chlorine which produces the most energetically stable compound the one with the most negative enthalpy change of formation.

ENTHALPIES OF SOLUTION AND HYDRATION


This page looks at the relationship

between enthalpies of solution, hydration enthalpies and lattice enthalpies.


Note: You really ought to have read the page s about Hess 's Law cycle s and lattic e entha lpies befor e you conti nue with this page.

Enthalpy change of solution


Defining enthalpy change of solution change of solution is the enthalpy change ance dissolves in water to give a solution o

Enthalpies of solution may be either positive or negative - in other words, some ionic substances dissolved

endothermically (for example, NaCl); others dissolve exothermically (for example NaOH). An infinitely dilute solution is one where there is a sufficiently large excess of water that adding any more doesn't cause any further heat to be absorbed or evolved. So, when 1 mole of sodium chloride crystals are dissolved in an excess of water, the enthalpy change of solution is found to be +3.9 kJ mol-1. The change is slightly endothermic, and so the temperature of the solution will be slightly lower than that of the original water.

Thinking about dissolving as an energy cycle Why is heat sometimes evolved and sometimes absorbed when a substance dissolves in water? To answer that it is useful to think about the various enthalpy changes that are involved in the process. You can think of an imaginary process where the crystal lattice is first broken up into its separate gaseous ions, and then those ions have water molecules wrapped around them. That is how they exist in the final solution. The heat energy needed to break up 1 mole of the crystal lattice is the lattice dissociation enthalpy. The heat energy released when new bonds are made between the ions and water molecules is known as

the hydration enthalpy of the ion. enthalpy is the enthalpy change when 1 m dissolve in sufficient water to give an infini ation enthalpies are always negative.

Factors affecting the size of hydration enthalpy Hydration enthalpy is a measure of the energy released when attractions are set up between positive or negative ions and water molecules. With positive ions, there may only be loose attractions between the slightly negative oxygen atoms in the water molecules and the positive ions, or there may be formal dative covalent (co-ordinate covalent) bonds. With negative ions, hydrogen bonds are formed between lone pairs of electrons on the negative ions and the slightly positive hydrogens in water molecules.
Note: You will find the attrac tions betw een water mole cules and positi ve ions discu ssed

on the page about dativ e coval ent bondi ng. You will find the attrac tions betw een negat ive ions and water mole cules discu ssed on the page about hydr ogen bondi ng.

The size of the hydration enthalpy is governed by the amount of attraction between the ions and the water molecules.
y

The attractions are stronger the smaller the ion. For example, hydration enthalpies fall as you go down a group in the Periodic Table. The small lithium ion has by far the highest hydration enthalpy in Group1, and the small fluoride

ion has by far the highest hydration enthalpy in Group 7. In both groups, hydration enthalpy falls as the ions get bigger. The attractions are stronger the more highly charged the ion. For example, the hydration enthalpies of Group 2 ions (like Mg2+) are much higher than those of Group 1 ions (like Na+).

Estimating enthalpies of solution from lattice enthalpies and hydration enthalpies The hydration enthalpies for calcium and chloride ions are given by the equations:

The following cycle is for calcium chloride, and includes a lattice dissociation enthalpy of +2258 kJ mol1. We have to use double the hydration enthalpy of the chloride ion because we are hydrating 2 moles of chloride ions. Make sure you understand exactly how the cycle works.

So . . . Hsol = +2258 - 1650 + 2(-364) Hsol = -120 kJ mol-1 Whether an enthalpy of solution turns out to be negative or positive depends on the relative sizes of the lattice enthalpy and the hydration enthalpies. In this particular case, the negative hydration enthalpies more than made up for the positive lattice dissociation enthalpy.

ENTHALPY CHANGE OF NEUTRALISATION


This page looks briefly at enthalpy changes of neutralisation. In common with my experience with most of the other pages in this section, searches for reliable data throw up various values for the same reaction. Don't worry too much about this. It doesn't actually affect the arguments.

Enthalpy change of neutralisation

Defining standard enthalpy change of neutralisation The standard enthalpy change of neutralisation is the enthalpy change when solutions of an acid and an alkali react together under standard conditions to produce 1 mole of water.

Notice that enthalpy change of solution is always measured per mole of water formed. Enthalpy changes of neutralisation are always negative - heat is given out when an acid and and alkali react. For reactions involving strong acids and alkalis, the values are always very closely similar, with values between -57 and -58 kJ mol-1. That varies slightly depending on the acid-alkali combination (and also on what source you look it up in!).

Why do strong acids reacting with strong alkalis give closely similar values? We make the assumption that strong acids and strong alkalis are fully ionised in solution, and that the ions behave independently of each other. For example, dilute hydrochloric acid contains hydrogen ions and chloride ions in solution. Sodium hydroxide solution consists of sodium ions and hydroxide ions in solution. The equation for any strong acid being neutralised by a strong alkali is essentially just a reaction between hydrogen ions and hydroxide ions to make water. The other ions present (sodium and chloride, for example) are just spectator ions, taking no part in the reaction. The full equation for the reaction between hydrochloric acid and sodium hydroxide solution is:

. . . but what is actually happening is:

If the reaction is the same in each case of a strong acid and a strong alkali, it isn't surprising that the enthalpy change is similar.
Note: Actually, of course, the enthalpy changes should be the same, not similar, if the assumptions we are making are exactly true! The small differences between strong acid-strong base combinations are almost invariably glossed over at this level. In fact, I can't remember ever seeing this discussed in any source - textbook or web. It isn't uncommon to find a list of enthalpy changes of neutralisation showing some variability in the strong acidstrong alkali cases, and then a few lines later on, this is ignored completely with a statement that in these cases, the enthalpy changes of neutralisation are the same, because . . . I have decide not to waste time trying to sort out the exact reasons for the problem, because I suspect it will take ages and ages, and it is never going to get asked at this level anyway.

Why do weak acids or weak alkalis give different values? In a weak acid, such as ethanoic acid, at ordinary concentrations, something like 99% of the acid isn't actually ionised. That means that the enthalpy change of neutralisation will include other enthalpy terms involved in ionising the acid as well as the reaction between the hydrogen ions and hydroxide ions. And in a weak alkali like ammonia solution, the ammonia is also present mainly as ammonia molecules in solution. Again, there

will be other enthalpy changes involved apart from the simple formation of water from hydrogen ions and hydroxide ions. For reactions involving ethanoic acid or ammonia, the measured enthalpy change of neutralisation is a few kilojoules less exothermic than with strong acids and bases. For example, one source which gives the enthalpy change of neutralisation of sodium hydroxide solution with HCl as -57.9 kJ mol-1, gives a value of -56.1 kJ mol-1 for sodium hydroxide solution being neutralised by ethanoic acid. For very weak acids, like hydrogen cyanide solution, the enthalpy change of neutralisation may be much less. A different source gives the value for hydrogen cyanide solution being neutralised by potassium hydroxide solution as -11.7 kJ mol-1, for example.

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. (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.

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.

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:

Note: The only difference if the reaction was endothermic

would be the relative positions of the reactants and productslin es. 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 Maxwell-Boltzmann Distribution which is a plot of the number of particles having each particular energy.
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!

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 MaxwellBoltzmann 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.

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.

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. 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.

he 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).

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 pressureon 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.

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.

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 aboutcollision 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 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 fastA splits up to make X and Y. This is described as the rate determining step of the 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.

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.

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.
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.

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

aboutcollision 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.

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 Maxwell-Boltzmann 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.
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, 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 MaxwellBoltzmann 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. 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.

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.

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.

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 Maxwell-Boltzmann 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: reaction Decomposition of hydrogen peroxide catalyst manganese(IV) oxide, MnO2 concentrated sulphuric acid iron vanadium(V) oxide, V2O5 nickel

Nitration of benzene Manufacture of ammonia by the Haber Process Conversion of SO2 into SO3 during the Contact Process to make sulphuric acid Hydrogenation of a C=C double bond

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.
Note: What follows assumes you have a reasonable idea about activation energy and its relationship with the Maxwell-Boltzmann 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 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:

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 that the majority of particles will react via the easier catalysed route.

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".

However, for this more formal and mathematical look at rates of reaction, the rate is usually measured by looking at how fast theconcentration 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 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 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. 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 toB. 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 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.
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.

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 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!

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?

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 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.

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.

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 termmolecularity. The only confusion is that you may sometimes find it used as if it meant the same as order. It doesn't!

NERGY PROFILES FOR SIMPLE REACTIONS

This page takes a closer look at simple energy profiles for reactions, and shows how they are slightly different for reactions involving an intermediate or just a transition state. Both of those terms are explained as well.

Types of Energy Profile What is an energy profile? If you have done any work involving activation energy or catalysis, you will have come across diagrams like this:

This diagram shows that, overall, the reaction is exothermic. The products have a lower energy than the reactants, and so energy is released when the reaction happens. It also shows that the molecules have to possess enough energy (called activation energy) to get the reactants over what we think of as the "activation energy barrier". In this example of a reaction profile, you can see that a catalyst offers a route for the reaction to follow which needs less activation energy. That, of course, causes the reaction to happen faster.
Note: If you aren't very happy about this, read the page about catalysts 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.

Diagrams like this are described as energy profiles. In the diagram above, you can clearly see that you need an input of energy to get the reaction going. Once the activation energy barrier has been passed, you can also see that you get even more energy released, and so the reaction is overall exothermic. If you had an endothermic reaction, a simple energy profile for a non-catalysed reaction would look like this:

Unfortunately, for many reactions, the real shapes of the energy profiles are slightly different from these, and the rest of this page explores some simple differences. What matters is whether the reaction goes via a single transition state or an intermediate. We will look at these two different cases in some detail.

Energy profiles for reactions which go via a single transition state only This is much easier to talk about with a real example. The equation below shows an organic chemistry reaction in which a bromine atom is being replaced by an OH group in an organic compound. The starting compound is bromoethane, and the organic product is ethanol.

During the reaction one of the lone pairs of electrons on the negatively charged oxygen in the -OH group is attracted to the carbon atom with the bromine attached. That's because the bromine is more electronegative than carbon, and so the electron pair in the C-Br bond is slightly closer to the bromine. The carbon atom becomes slightly positively charged and the bromine slightly negative.

As the hydroxide ion approaches the slightly positive carbon, a new bond starts to be set up between the oxygen and the carbon. At the same time, the bond between the carbon and bromine starts to break as the electrons in the bond are repelled towards the bromine. At some point, the process is exactly half complete. The carbon atom now has the oxygen half-attached, the bromine half-attached, and the three other groups still there, of course.

And then the process completes:

Note: These diagrams have been simplified in various ways to make the process clearer. For example, the true arrangement of the lone pairs of electrons around the oxygen in the first diagram has been simplified for clarity. The bromine also has 3 lone pairs as well as the bonding pair, but they play no part. And, of course, the other groups attached to the carbon have been left out in order to concentrate on what is important.

The second diagram where the bonds are half-made and half-broken is called the transition state, and it is at this point that the energy of the system is at its

maximum. This is what is at the top of the activation energy barrier.

But the transition state is entirely unstable. Any tiny change in either direction will send it either forward to make the products or back to the reactants again. Neither is there anything special about a transition state except that it has this maximum energy. You can't isolate it, even for a very short time. The situation is entirely different if the reaction goes through an intermediate. Again, we'll look at a specific example.

Energy profiles for reactions which go via an intermediate For reasons which you may well meet in the organic chemistry part of your course, a different organic bromine-containing compound reacts with hydroxide ions in an entirely different way. In this case, the organic compound ionises slightly in a slow reaction to produce an intermediate positive organic ion. This then goes on to react very rapidly with hydroxide ions.

Note: If you haven't come across the use of curly arrows in organic chemistry yet, all you need to know for now is that they show the movement of a pair of electrons. In the first equation, for example, the bonding pair of electrons in the C-Br bond moves entirely on to the bromine to make a bromide ion. In the second equation, a lone pair on the hydroxide ion moves towards the positive carbon to form a covalent bond.

The big difference in this case is that the positively charged organic ion can actually be detected in the mixture. It is very unstable, and soon reacts with a hydroxide ion (or picks up its bromide ion again). But, for however short a time, it does have a real presence in the system. That shows itself in the energy profile.

The stability (however temporary and slight) of the intermediate is shown by the fact that there are small activation barriers to its conversion either into the products or back into the reactants again. Notice that the barrier on the product side of the intermediate is lower than that on the reactant side. That means that there is a greater chance of it finding the extra bit of energy to convert into products. It would need a greater amount of energy to convert back to the reactants again. I've labelled these peaks "ts1" and "ts2" - they both represent transition states between the intermediate and either the reactants or the products. During either conversion, there will be some arrangement of the atoms which causes an energy maximum - that's all a transition state is.

ATE 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 menufirst.

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:

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. 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 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 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 mol1. 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.

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. 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.

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. 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. 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 carbon-carbon 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 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.

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.

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.

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.

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.

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.

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. This can happen in two ways depending on whether the ClO radical hits an ozone molecule or an oxygen radical. If it hits an oxygen radical (produced from one of the reactions we've looked at previously):

Or if it hits an ozone molecule:

Because the chlorine radical keeps on being regenerated, each one can destroy thousands of ozone molecules.
Note: If you are a UK A level student it is probable that you will only want one of these last two equations depending on what your syllabus says. Unfortunately, at the time of writing, different A level syllabuses were quoting different final equations. You need to check your current syllabus. If you don't have a copy of your syllabus, follow this link to find out how to get one.

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).

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:

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 necessarilyshows 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.

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.

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 adynamic 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. 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).

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%.

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 dynamicshows 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:
y y

"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 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".

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.

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.

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. 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 theintroductory 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
y

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.

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 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 amore 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. 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 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.

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.

Summary
y

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. 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.

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:

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


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. 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 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.

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.
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:
y y y

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 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:

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 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 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 - 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 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.

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 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.
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 heterogeneousmean, you would find it useful to follow this link and read thebeginning 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.
y

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 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.

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.

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 Kcexpression.

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:

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 right-hand 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:

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.

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 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. 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 Kcpage.

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.

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 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 nitrogen hydrogen ammonia mole fraction 20/100 = 0.2 60/100 = 0.6 20/100 = 0.2 partial pressure 0.2 x 200 = 40 atm 0.6 x 200 = 120 atm 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 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.

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. 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. 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.

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 Kcexpression 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. 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 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, Kpwill 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?

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 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:

The Kp expression is:

Two values for Kp are:

temperature 500 K 700 K

Kp 160 54

You can see that as the temperature increases, the value of Kpfalls.
Note: You might possibly be wondering what the units of Kpare. 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 mychemistry calculations book.

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 (or equivalent) 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.

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. 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
y y

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:

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
y y

An acid is a proton (hydrogen ion) donor. A base is a proton (hydrogen ion) acceptor.

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. Coordinate 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.

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:

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:

y y

The HA is an acid because it is donating a proton (hydrogen ion) to the water. 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 A- ion:
y y

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 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.

Note: You might also come across the term amphiprotic in this context. The two words are related and easily confused. An amphiprotic substance is one which can both donate hydrogen ions (protons) and also accept them. Water is a good example of such a compound. The water acts as both an acid (donating hydrogen ions) and as a base (by accepting them). The "protic" part of the word refers to the hydrogen ions (protons) either being donated or accepted. Other examples of amphiprotic compounds are amino acids, and ions like HSO4- (which can lose a hydrogen ion to form sulphate ions or accept one to form sulphuric acid). But as well as being amphiprotic, these compounds are alsoamphoteric. Amphoteric means that they have reactions as both acids and bases. So what is the difference between the two terms? All amphiprotic substances are also amphoteric - but the reverse isn't true. There are amphoteric substances which don't either donate or accept hydrogen ions when they act as acids or bases. There is a whole new definition of acid-base

behaviour that you are just about to meet (the Lewis theory) which doesn't necessarily involve hydrogen ions at all. A Lewis acid is an electron pair acceptor; a Lewis base is an electron pair donor (see below). Some metal oxides (like aluminium oxide) are amphoteric - they react both as acids and bases. For example, they react as bases because the oxide ions accept hydrogen ions to make water. That's not a problem as far as the definition of amphiprotic is concerned - but the reaction as an acid is. The aluminium oxide doesn't contain any hydrogen ions to donate! But aluminium oxide reacts with bases like sodium hydroxide solution to form complex aluminate ions. You can think of lone pairs on hydroxide ions as forming dative covalent (coordinate) bonds with empty orbitals in the aluminium ions. The aluminium ions are accepting lone pairs (acting as a Lewis acid). So aluminium oxide can act as both an acid and a base - and so is amphoteric. But it isn'tamphiprotic because both of the acid reaction and the base reaction don't involve hydrogen ions. I have gone through 40-odd years of teaching (in the lab, and via books and the internet) without once using the term amphiprotic! I simply don't see the point of it. The term amphoteric takes in all the cases of substances functioning as both acids and bases without exception. The term amphiprotic can only be used where both of these functions involve transference of hydrogen ions - in other words, it can only be used if you are limited to talking about the Bronsted-Lowry theory. Personally, I would stick to the older, more useful, term "amphoteric" unless your syllabus demands that you use the word "amphiprotic".

The Lewis Theory of acids and bases


This theory extends well beyond the things you normally think of as acids and bases. The theory
y y

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 Bronsted-Lowry 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.

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 with coordinate 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 co-ordinate 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.

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 hydrogenchlorine bond will be attracted towards the chlorine end, leaving the hydrogen slightly positive and the chlorine slightly negative.

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.

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:
y y

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. 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, Kaand pKa. It is important that you don't confuse the words strong and weakwith 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 baseson 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.

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+.
Note: You should find out what your 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 are doing a UKbased exam and haven't got copies of yoursyllabus 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.

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.

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.

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 equilibriumfollow 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:

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.

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:

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 hydrofluoric acid methanoic acid ethanoic acid hydrogen sulphide Ka (mol dm-3) 5.6 x 10-4 1.6 x 10-4 1.7 x 10-5 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 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: acid hydrofluoric acid methanoic acid ethanoic acid hydrogen sulphide Ka (mol dm-3) 5.6 x 10-4 1.6 x 10-4 1.7 x 10-5 8.9 x 10-8 pKa 3.3 3.8 4.8 7.1

Note: Notice that unlike Ka, pKa doesn't have any units.

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:
y y

The lower the value for pKa, the stronger the acid. The higher the value for pKa, the weaker the acid.

HE 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.

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. 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-6at 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 pKwdoesn'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 1014 mol2 dm-6. 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 Kwexpression 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:

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) 0 Kw (mol2 dm-6) 0.114 x 10-14 pH 7.47

10 20 25 30 40 50 100

0.293 x 10-14 0.681 x 10-14 1.008 x 10-14 1.471 x 10-14 2.916 x 10-14 5.476 x 10-14 51.3 x 10-14

7.27 7.08 7.00 6.92 6.77 6.63 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. 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.

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.

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
y

Work out the concentration of the hydroxide ions.

y y

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. [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.

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. 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. base C6H5NH2 NH3 CH3NH2 CH3CH2NH2 Kb (mol dm-3) 4.17 x 10-10 1.78 x 10-5 4.37 x 10-4 5.37 x 10-4 pKb 9.38 4.75 3.36 3.27

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. 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 pointof 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 cm3of 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 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:
y y y

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.

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.

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

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.

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.

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

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.

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.

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 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 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 adiprotic 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:

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. 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 commonly quoted acid-base combinations, 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.

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:

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. In an alkaline solution, methyl orange is yellow and the structure is:

Now, you might think that when you add an acid, the hydrogen ion would be picked up by the negatively charged oxygen. That's the obvious place for it to go. Not so! In fact, the hydrogen ion attaches to one of the nitrogens in the nitrogennitrogen double bond to give a structure which might be drawn like this:

Note: You may find other structures for this with different arrangements of the bonds (although always with the hydrogen attached to that same nitrogen). The truth is that there is delocalisation over the entire structure, and no simple picture will show it properly. Don't worry about this exact structure - it is just to show a real case where the colour of a compound is drastically changed by the presence or absence of a hydrogen ion.

You have the same sort of equilibrium between the two forms of methyl orange 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.

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!
Note: If you are interested in understanding the reason for the colour changes in methyl orange and phenolphthalein, they are discussed on a page in the analysis section of the site about UV-visible spectroscopy. This is quite difficult stuff, and if you are coming at this from scratch you will have to explore at least one other page before you can make sense of what is on that page. There is a link to help you to do that.Don't start this lightly! Use the BACK button (or more likely the HISTORY file or GO menu) on your browser to return to this page much later.

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.

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 rearrange the last equation so that the hydrogen ion concentration is on the lefthand 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 litmus methyl orange phenolphthalein 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. pKind 6.5 3.7 9.3

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 pKindvalue. The exact values for the three indicators we've looked at are: indicator litmus methyl orange phenolphthalein pKind 6.5 3.7 9.3 pH range 5-8 3.1 - 4.4 8.3 - 10.0

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. 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.

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

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

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 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.

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.

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.

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:


y y y

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. 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 oneway, 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.

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:
y y y

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. 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. 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. 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 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.

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. The diagram shows the stretching that happens in a carbon-oxygen single bond. There will, of course, be other atoms attached to both the carbon and the oxygen. For example, it could be the carbon-oxygen bond in methanol, CH3OH.

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. The diagram shows the bending of the bonds in a water molecule. The effect of this, of course, is that the bond angle between the two hydrogen-oxygen bonds fluctuates slightly around its average value. Imagine a lab model of a water molecule where the atoms are joined together with springs. These bending vibrations are what you would see if you shook the model gently.

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.

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. 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:

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 bondsis 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 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 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. WHAT IS C-13 NMR?

This page describes what a C-13 NMR spectrum is and how it tells you useful things about the carbon atoms in organic molecules.

The background to C-13 NMR spectroscopy Nuclear magnetic resonance is concerned with the magnetic properties of certain nuclei. On this page we are focussing on the magnetic behaviour of carbon-13 nuclei.

Carbon-13 nuclei as little magnets About 1% of all carbon atoms are the C-13 isotope; the rest (apart from tiny amounts of the radioactive C-14) is C-12. C-13 NMR relies on the magnetic properties of the C-13 nuclei. Carbon-13 nuclei fall into a class known as "spin " nuclei for reasons which don't really need to concern us at the introductory level this page is aimed at (UK A level and its equivalents). The effect of this is that a C-13 nucleus can behave as a little magnet. C-12 nuclei don't have this property. 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.

Because a C-13 nucleus behaves like a little magnet, it means that it can also be aligned with an external magnetic field or opposed to it. 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 25 - 100 MHz. (BBC Radio 4 is found between 92 - 95 MHz!) If you have also looked at proton-NMR, the frequency is about a quarter of that used to flip a hydrogen nucleus for a given magnetic field strength. It's possible to detect this interaction between the radio waves of just the right frequency and the carbon-13 nucleus as it flips from one orientation to the other as a peak on a graph. This flipping of the carbon-13 nucleus from one magnetic alignment to the other by the radio waves is known as the resonance condition.

The importance of the carbon's environment What we've said so far would apply to an isolated carbon-13 nucleus, but real carbon atoms in real bonds 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 carbon-13 nucleus.

Suppose you were using a radio frequency of 25 MHz, and you adjusted the size of the magnetic field so that an isolated carbon-13 atom was in the resonance condition. If you replaced the isolated carbon with the more realistic case of it being surrounded by bonding electrons, 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 shielding effect of the electrons. Now suppose that you attached the carbon to something more electronegative. The electrons in the bond would be further away from the carbon nucleus, and so would have less of a lowering effect on the magnetic field around the carbon nucleus.

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!

The external magnetic field needed to bring the carbon into resonance will be smaller if it is attached to a more electronegative element, because the C-13 nucleus feels more of the field. Even small differences in the electronegativities of the attached atoms will make a difference to the magnetic field needed to achieve resonance.

Summary For a given radio frequency (say, 25 MHz) each carbon-13 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 carbon atom's environment in the molecule.

Features of a C-13 NMR spectrum


The C-13 NMR spectrum for ethanol This is a simple example of a C-13 NMR spectrum. Don't worry about the scale for now - we'll look at that in a minute.

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.

There are two peaks because there are two different environments for the carbons. The carbon in the CH3 group is attached to 3 hydrogens and a carbon. The carbon in the CH2 group is attached to 2 hydrogens, a carbon and an oxygen. The two lines are in different places in the NMR spectrum because they need different external magnetic fields to bring them in to resonance at a particular radio frequency.

The C-13 NMR spectrum for a more complicated compound This is the C-13 NMR spectrum for 1-methylethyl propanoate (also known as isopropyl propanoate or isopropyl propionate).

This time there are 5 lines in the spectrum. That means that there must be 5 different environments for the carbon atoms in the compound. Is that reasonable from the structure?

Well - if you count the carbon atoms, there are 6 of them. So why only 5 lines? In this case, two of the carbons are in exactly the same environment. They are attached to exactly the same things. Look at the two CH3 groups on the righthand side of the molecule. You might reasonably ask why the carbon in the CH3 on the left isn't also in the same environment. Just like the ones on the right, the carbon is attached to 3 hydrogens and another carbon. But the similarity isn't exact - you have to chase the similarity along the rest of the molecule as well to be sure. The carbon in the left-hand CH3 group is attached to a carbon atom which in turn is attached to a carbon with two oxygens on it - and so on down the molecule. That's not exactly the same environment as the carbons in the right-hand CH3 groups. They are attached to a carbon which is attached to a single oxygen - and so on down the molecule. We'll look at this spectrum again in detail on the next page - and look at some more similar examples as well. This all gets easier the more examples you look at. For now, all you need to realise is that each line in a C-13 NMR spectrum recognises a carbon atom in one particular environment in the compound. If two (or more) carbon atoms in a compound have exactly the same environment, they will be represented by a single line.
Note: If you are fairly wide-awake, you might wonder why all this works, since only about 1% of carbon atoms are C-13. These are the only ones picked up by this form of NMR. If you had a single molecule of ethanol, then the chances are only about 1 in 50 of there being one C-13 atom in it, and only about 1 in 10,000 of both being C-13. But you have got to remember that you will be working with a sample containing huge numbers of molecules. The instrument can pick up the magnetic effect of the C-13 nuclei in the carbon of the CH3 group and the carbon of the CH2group even if they are in separate molecules. There's no need for them to be in the same one.

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 carbon-13 atoms in tetramethylsilane usually called TMS.Everything else is compared with this.

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:
y

It has 4 carbon 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 carbon atoms all doing the same thing). The electrons in the C-Si bonds are closer to the carbons in this compound than in almost any other one. That means that these carbon 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 carbons 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). measured in parts per million - ppm. is called the chemical shift and is

A peak at a chemical shift of, say, 60 means that the carbon atoms which caused that peak need a magnetic field 60 millionths lessthan the field needed by TMS to produce resonance.

A peak at a chemical shift of 60 is said to be downfield of TMS. The further to the left a peak is, the more downfield it is.
Note: If you are familiar with proton-NMR, you will notice that the chemical shifts for C-13 NMR are much bigger than for protonNMR. In C-13 NMR, they range up to about 200 ppm. In protonNMR they only go up to about 12 ppm. You don't need to worry about the reasons for this at this level.

Solvents for NMR spectroscopy NMR spectra are usually measured using solutions of the substance being investigated. A commonly used solvent is CDCl3. This is a trichloromethane (chloroform) molecule in which the hydrogen has been replaced by its isotope, deuterium. CDCl3 is also commonly used as the solvent in proton-NMR because it doesn't have any ordinary hydrogen nuclei (protons) which would give a line in a proton-NMR spectrum. It does, of course, have a carbon atom - so why doesn't it give a potentially confusing line in a C-13 NMR spectrum? In fact it does give a line, but the line has an easily recognisable chemical shift and so can be removed from the final spectrum. All of the spectra from the SDBS have this line removed to avoid any confusion.
NTERPRETING C-13 NMR SPECTRA? This page takes an introductory look at how you can get useful information from a C-13 NMR spectrum. Important: If you have come straight to this page via a search engine (including the Google site search on the Main Menu of Chemguide), you should be aware that this is the second of two pages about C-13 NMR. Unless you are familiar with C-13 NMR, you should read the introduction to C-13 NMR first by following this link.

Taking a close look at three C-13 NMR spectra


The C-13 NMR spectrum for ethanol

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.

Remember that each peak identifies a carbon atom in a different environment within the molecule. In this case there are two peaks because there are two different environments for the carbons. The carbon in the CH3 group is attached to 3 hydrogens and a carbon. The carbon in the CH2 group is attached to 2 hydrogens, a carbon and an oxygen. So which peak is which? You might remember from the introductory page that the external magnetic field experienced by the carbon nuclei is affected by the electronegativity of the atoms attached to them. The effect of this is that the chemical shift of the carbon increases if you attach an atom like oxygen to it. That means that the peak at about 60 (the larger chemical shift) is due to the CH2 group because it has a more electronegative atom attached.
Note: In principle, you should be able to work out the fact that the carbon attached to the oxygen will have the larger chemical shift. In practice, given the level I am aiming at (16 - 18 year old chemistry students), you always work from tables of chemical shift values for different groups (see below). What if you needed to work it out? The electronegative oxygen pulls

electrons away from the carbon nucleus leaving it more exposed to any external magnetic field. That means that you will need a smaller external magnetic field to bring the nucleus into the resonance condition than if it was attached to less electronegative things. The smaller the magnetic field needed, the higher the chemical shift. All this is covered in more detail on the introduction to C-13 NMR page mentioned above.

A table of typical chemical shifts in C-13 NMR spectra carbon environment C=O (in ketones) C=O (in aldehydes) C=O (in acids and esters) C in aromatic rings C=C (in alkenes) RCH2OH RCH2Cl RCH2NH2 R3CH CH3COR2CH2 RCH3 chemical shift (ppm) 205 - 220 190 - 200 170 - 185 125 - 150 115 - 140 50 - 65 40 - 45 37 - 45 25 - 35 20 - 30 16 - 25 10 - 15

In the table, the "R" groups won't necessarily be simple alkyl groups. In each case there will be a carbon atom attached to the one shown in red, but there may well be other things substituted into the "R" group. If a substituent is very close to the carbon in question, and very electronegative, that might affect the values given in the table slightly. For example, ethanol has a peak at about 60 because of theCH2OH group. No

problem! It also has a peak due to the RCH3 group. The "R" group this time is CH2OH. The electron pulling effect of the oxygen atom increases the chemical shift slightly from the one shown in the table to a value of about 18.

A simplification of the table At the time of writing, a draft UK syllabus (Cambridge pre-U) was expecting their students to learn the following simplification: carbon environment chemical shift (ppm) C-C C-O C=C C=O 0 - 50 50 - 100 100 - 150 150 - 200

This may, of course, change and other syllabuses might want something similar. The only way to find out is to check your syllabus, and recent question papers to see whether you are given tables of chemical shifts or not.
Note: If you are following a UK-based syllabus, and haven't got a copies of your syllabus and past papers, follow this link to find out how to get hold of them.

The C-13 NMR spectrum for but-3-en-2-one This is also known as 3-buten-2-one (amongst many other things!)

Here is the structure for the compound:

You can pick out all the peaks in this compound using the simplified table above. The peak at just under 200 is due to a carbon-oxygen double bond. The two peaks at 137 and 129 are due to the carbons at either end of the carboncarbon double bond. And the peak at 26 is the methyl group which, of course, is joined to the rest of the molecule by a carbon-carbon single bond. If you want to use the more accurate table, you have to put a bit more thought into it - and, in particular, worry about the values which don't always exactly match those in the table! The carbon-oxygen double bond in the peak for the ketone group has a slightly lower value than the table suggests for a ketone. There is an interaction between the carbon-oxygen and carbon-carbon double bonds in the molecule which affects the value slightly. This isn't something which we need to look at in detail for the purposes of this topic. You must be prepared to find small discrepancies of this sort in more complicated molecules - but don't worry about this for exam purposes at this level. Your examiners should give you shift values which exactly match the compound you are given. The two peaks for the carbons in the carbon-carbon double bond are exactly where they would be expected to be. Notice that they aren't in exactly the same environment, and so don't have the same shift values. The one closer to

the carbon-oxygen double bond has the larger value. And the methyl group on the end has exactly the sort of value you would expect for one attached to C=O. The table gives a range of 20 - 30, and that's where it is. One final important thing to notice. There are four carbons in the molecule and four peaks because they are all in different environments. But they aren't all the same height. In C-13 NMR, you can't draw any simple conclusions from the heights of the various peaks.

The C-13 NMR spectrum for 1-methylethyl propanoate 1-methylethyl propanoate is also known as isopropyl propanoate or isopropyl propionate.

Here is the structure for 1-methylethyl propanoate:

Two simple peaks There are two very simple peaks in the spectrum which could be identified easily from the second table above. The peak at 174 is due to a carbon in a carbon-oxygen double bond. (Looking at the more detailed table, this peak is due to the carbon in a carbon-oxygen double bond in an acid or ester.)

The peak at 67 is due to a different carbon singly bonded to an oxygen. Those two peaks are therefore due to:

If you look back at the more detailed table of chemical shifts, you will find that a carbon singly bonded to an oxygen has a range of 50 - 65. 67 is, of course, a little bit higher than that. As before, you must expect these small differences. No table can account for all the fine differences in environment of a carbon in a molecule. Different tables will quote slightly different ranges. At this level, you can just ignore that problem! Before we go on to look at the other peaks, notice the heights of these two peaks we've been talking about. They are both due to a single carbon atom in the molecule, and yet they have different heights. Again, you can't read any reliable information directly from peak heights in these spectra.

The three right-hand peaks From the simplified table, all you can say is that these are due to carbons attached to other carbon atoms by single bonds. But because there are three peaks, the carbons must be in three different environments. The more detailed table is more helpful. Here are the structure and the spectrum again:

The easiest peak to sort out is the one at 28. If you look back at the table, that could well be a carbon attached to a carbon-oxygen double bond. The table quotes the group as CH3CO-, but replacing one of the hydrogens by a simple CH3 group won't make much difference to the shift value. The right-hand peak is also fairly easy. This is the left-hand methyl group in the molecule. It is attached to an admittedly complicated R group (the rest of the molecule). It is the bottom value given in the detailed table. The tall peak at 22 must be due to the two methyl groups at the right-hand end of the molecule - because that's all that's left. These combine to give a single peak because they are both in exactlythe same environment. If you are looking at the detailed table, you need to think very carefully which of the environments you should be looking at. Without thinking, it is tempting to go for the R2CH2 with peaks in the 16 - 25 region. But you would be wrong! The carbons we are interested in are the ones in the methyl group, not in the R groups. These carbons are again in the environment: RCH3. The R is the rest of the molecule. The table says that these should have peaks in the range 10 - 15, but our peak is a bit higher. This is because of the presence of the nearby oxygen atom. Its electronegativity is pulling electrons away from the methyl groups - and, as we've seen above, this tends to increase the chemical shift slightly. Once again, don't worry about the discrepancies. In an exam, perhaps your examiners will just want you to have learnt the simple table above - in which case, they can't expect you to work out which peak is which in a complicated spectrum of this sort. Or they will give you tables of chemical shifts - in which case, they will give you values which match the peaks in the spectra. Remember that you are only doing an introduction to C-13 NMR at this level. It

isn't going to be that hard in an exam!

Working out structures from C-13 NMR spectra


So far, we've just been trying to see the relationship between carbons in particular environments in a molecule and the spectrum produced. We've had all the information necessary. Now let's make it a little more difficult - but we'll work from much easier examples! In each example, try to work it out for yourself before you read the explanation.

Example 1 How could you tell from just a quick look at a C-13 NMR spectrum (and without worrying about chemical shifts) whether you had propanone or propanal (assuming those were the only options)?

Because these are isomers, each has the same number of carbon atoms, but there is a difference between the environments of the carbons which will make a big impact on the spectra. In propanone, the two carbons in the methyl groups are in exactly the same environment, and so will produce only a single peak. That means that the propanone spectrum will have only 2 peaks - one for the methyl groups and one for the carbon in the C=O group. However, in propanal, all the carbons are in completely different environments, and the spectrum will have three peaks.

Example 2 Thare are four alcohols with the molecular formula C4H10O.

Which one produced the C-13 NMR spectrum below?

You can do this perfectly well without referring to chemical shift tables at all. In the spectrum there are a total of three peaks - that means that there are only three different environments for the carbons, despite there being four carbon atoms. In A and B, there are four totally different environments. Both of these would produce four peaks. In D, there are only two different environments - all the methyl groups are exactly equivalent. D would only produce two peaks. That leaves C. Two of the methyl groups are in exactly the same environment - attached to the rest of the molecule in exactly the same way. They would only produce one peak. With the other two carbon atoms, that would make a total of three. The alcohol is C.

Example 3 This follows on from Example 2, and also involves an isomer of C4H10O but which isn't an alcohol. Its C-13 NMR spectrum is below. Work out what its

structure is.

Because we don't know what sort of structure we are looking at, this time it would be a good idea to look at the shift values. The approximations are perfectly good, and we will work from this table: carbon environment chemical shift (ppm) C-C C-O C=C C=O 0 - 50 50 - 100 100 - 150 150 - 200

There is a peak for carbon(s) in a carbon-oxygen single bond and one for carbon(s) in a carbon-carbon single bond. That would be consistent with C-CO in the structure. It isn't an alcohol (you are told that in the question), and so there must be another carbon on the right-hand side of the oxygen in the structure in the last paragraph. The molecular formula is C4H10O, and there are only two peaks. The only solution to that is to have two identical ethyl groups either side of the oxygen. The compound is ethoxyethane (diethyl ether), CH3CH2OCH2CH3.

Example 4 Using the simplified table of chemical shifts above, work out the structure of the compound with the following C-13 NMR spectrum. Its molecular formula is C4H6O2.

Let's sort out what we've got.


y y y y

There are four peaks and four carbons. No two carbons are in exactly the same environment. The peak at just over 50 must be a carbon attached to an oxygen by a single bond. The two peaks around 130 must be the two carbons at either end of a carbon-carbon double bond. The peak at just less than 170 is the carbon in a carbon-oxygen double bond.

Putting this together is a matter of playing around with the structures until you have come up with something reasonable. But you can't be sure that you have got the right structure using this simplified table. In this particular case, the spectrum was for the compound:

If you refer back to the more accurate table of chemical shifts towards the top of the page, you will get some better confirmation of this. The relatively low value of the carbon-oxygen double bond peak suggests an ester or acid rather than an aldehyde or ketone.

It can't be an acid because there has to be a carbon attached to an oxygen by a single bond somewhere - apart from the one in the -COOH group. We've already accounted for that carbon atom from the peak at about 170. If it was an acid, you would already have used up both oxygens in the structure in the -COOH group. Without this information, though, you could probably come up with reasonable alternative structures. If you were working from the simplified table in an exam, your examiners would have to allow any valid alternatives. WHAT IS NUCLEAR MAGNETIC RESONANCE (NMR)?

This page describes what a proton NMR spectrum is and 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. On this page we are focussing 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. 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.

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.

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!

The external magnetic field needed to bring the hydrogen into resonance will be smaller if it is attached to a more electronegative element, because the hydrogen nucleus feels more of the field. Even small differences in the electronegativities of the attached atom or groups of atoms will make a difference to the magnetic field needed to achieve resonance.

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:

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.

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:
y

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). measured in parts per million - ppm. is called the chemical shift and is

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.

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. 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.

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 CH2group, the other to an oxygen. 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 are doing a UK-based exam and 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:

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 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:

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. 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 CH2group and the COOH group. There would be three peaks in total with areas in the ratio 9:2:1 IGH 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:
y y y

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: 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 eachcluster 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 quartet next door to carbon with no hydrogens attached next door to a CH group next door to a CH2 group next door to a CH3 group
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.

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 CH3group. 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:

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:
y y y y

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.

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).

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.

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 CH2group.
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!

Equivalent hydrogen atoms Hydrogen atoms attached to the same carbon atom are said to beequivalent. 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.

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.

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.

You might also like