Sunteți pe pagina 1din 152

REDOX MENU

Definitions of oxidation and reduction . . . Covers definitions of oxidation and reduction in terms of transfer of oxygen, hydrogen and electrons. Writing equations for redox reactions . . . How to construct ionic equations for redox reactions by working out electron half equations and then combining them. Oxidation states (oxidation numbers) . . . How to work out oxidation states (oxidation numbers) and use them to decide simply what is being oxidised and what is being reduced in a reaction.

DEFINITIONS OF OXIDATION AND REDUCTION (REDOX)


This page looks at the various definitions of oxidation and reduction (redox) in terms of the transfer of oxygen, hydrogen and electrons. It also explains the terms oxidising agent and reducing agent.

Oxidation and reduction in terms of oxygen transfer


Definitions

Oxidation is gain of oxygen. Reduction is loss of oxygen.

For example, in the extraction of iron from its ore:

Because both reduction and oxidation are going on side-by-side, this is known as a redox reaction. Oxidising and reducing agents An oxidising agent is substance which oxidises something else. In the above example, the iron(III) oxide is the oxidising agent.

A reducing agent reduces something else. In the equation, the carbon monoxide is the reducing agent.

Oxidising agents give oxygen to another substance. Reducing agents remove oxygen from another substance.

Oxidation and reduction in terms of hydrogen transfer


These are old definitions which aren't used very much nowadays. The most likely place you will come across them is in organic chemistry. Definitions

Oxidation is loss of hydrogen. Reduction is gain of hydrogen.

Notice that these are exactly the opposite of the oxygen definitions. For example, ethanol can be oxidised to ethanal:

You would need to use an oxidising agent to remove the hydrogen from the ethanol. A commonly used oxidising agent is potassium dichromate(VI) solution acidified with dilute sulphuric acid.

Note: The equation for this is rather complicated for this introductory page. If you are interested, you will find a similar example (ethanol to ethanoic acid) on the page dealing with writing equations for redox reactions.

Ethanal can also be reduced back to ethanol again by adding hydrogen to it. A possible reducing agent is sodium tetrahydridoborate, NaBH4. Again the equation is too complicated to be worth bothering about at this point.

An update on oxidising and reducing agents

Oxidising agents give oxygen to another substance or remove hydrogen from it.

Reducing agents remove oxygen from another substance or give hydrogen to it.

Oxidation and reduction in terms of electron transfer


This is easily the most important use of the terms oxidation and reduction at A' level. Definitions

Oxidation is loss of electrons. Reduction is gain of electrons.

It is essential that you remember these definitions. There is a very easy way to do this. As long as you remember that you are talking about electron transfer:

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:

An oxidising agent oxidises something else. Oxidation is loss of electrons (OIL RIG). That means that an oxidising agent takes electrons from that other substance. So an oxidising agent must gain electrons.

Or you could think it out like this:


An oxidising agent oxidises something else. That means that the oxidising agent must be being reduced. Reduction is gain of electrons (OIL RIG). So an oxidising agent must gain electrons.

Understanding is a lot safer than thoughtless learning!

WRITING IONIC EQUATIONS FOR REDOX REACTIONS


This page explains how to work out electron-half-reactions for oxidation and reduction processes, and then how to combine them to give the overall ionic equation for a redox reaction. This is an important skill in inorganic chemistry. Don't worry if it seems to take you a long time in the early stages. It is a fairly slow process even with experience. Take your time and practise 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 syllabusespage 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 "half-equations" 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:

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 lefthand side.

Now that all the atoms are balanced, all you need to do is balance the charges. At the moment there are a net 7+ charges on the left-hand side (1- and 8+), but only 2+ on the right. Add 5 electrons to the left-hand side to reduce the 7+ to 2+.

This is the typical sort of half-equation which you will have to be able to work out. The sequence is usually:

Balance the atoms apart from oxygen and hydrogen. Balance the oxygens by adding water molecules. Balance the hydrogens by adding hydrogen ions. Balance the charges by adding electrons.

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:

Note: You have now seen a cross-section of the sort of equations which you could be asked to work out. Now you need to practice so that you can do this reasonably quickly and very accurately! Aim to get an averagely complicated example done in about 3 minutes. If you want a few more examples, and the opportunity to practice with answers available, you might be interested in looking in chapter 1 of my book on Chemistry Calculations.

Reactions done under alkaline conditions Working out half-equations for reactions in alkaline solution is decidedly more tricky than those above. You are less likely to be asked to do this at this level (UK A level and its equivalents), and for that reason I've covered these on a separate page (link below). It would be worthwhile checking your syllabus and past papers before you start worrying about these!

OXIDATION STATES (OXIDATION NUMBERS)


This page explains what oxidation states (oxidation numbers) are and how to calculate them and make use of them. Oxidation states are straightforward to work out and to use, but it is quite difficult to define what they are in any quick way. Explaining what oxidation states (oxidation numbers) are Oxidation states simplify the whole process of working out what is being oxidised and what is being reduced in redox reactions. However, for the purposes of this introduction, it would be helpful if you knew about:

oxidation and reduction in terms of electron transfer electron-half-equations

Note: If you aren't sure about either of these things, you might want to look at the pages on redox definitionsand electron-half-equations. It would probably be best to read on and come back to these links if you feel you need to.

We are going to look at some examples from vanadium chemistry. If you don't know anything about vanadium, it doesn't matter in the slightest. Vanadium forms a number of different ions - for example, V2+ and V3+. If you think about how these might be produced from vanadium metal, the 2+ ion will be formed by oxidising the metal by removing two electrons:

The vanadium is now said to be in an oxidation state of +2. Removal of another electron gives the V3+ ion:

The vanadium now has an oxidation state of +3. Removal of another electron gives a more unusual looking ion, VO2+.

The vanadium is now in an oxidation state of +4. Notice that the oxidation state isn't simply counting the charge on the ion (that was true for the first two cases but not for this one). The positive oxidation state is counting the total number of electrons which have had to be removed - starting from the element. It is also possible to remove a fifth electron to give another ion (easily confused with the one before!). The oxidation state of the vanadium is now +5.

Every time you oxidise the vanadium by removing another electron from it, its oxidation state increases by 1. Fairly obviously, if you start adding electrons again the oxidation state will fall. You could eventually get back to the element vanadium which would have an oxidation state of zero. What if you kept on adding electrons to the element? You can't actually do that with vanadium, but you can with an element like sulphur.

The sulphur has an oxidation state of -2. Summary Oxidation state shows the total number of electrons which have been removed from an element (a positive oxidation state) or added to an element (a negative oxidation state) to get to its present state.

Oxidation involves an increase in oxidation state Reduction involves a decrease in oxidation state Recognising this simple pattern is the single most important thing about the concept of oxidation states. If you know how the oxidation state of an element changes during a reaction, you can instantly tell whether it is being oxidised or reduced without having to work in terms of electron-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!

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:

element Group 1 metals Group 2 metals Oxygen Hydrogen Fluorine Chlorine

usual oxidation state always +1 always +2 usually -2 usually +1 always -1 usually -1

exceptions

except in peroxides and F2O (see below) except in metal hydrides where it is -1 (see below)

except in compounds with O or F (see below)

The reasons for the exceptions Hydrogen in the metal hydrides Metal hydrides include compounds like sodium hydride, NaH. In this, the hydrogen is present as a hydride ion, H-. The oxidation state of a simple ion like hydride is equal to the charge on the ion - in this case, -1. Alternatively, you can think of it that the sum of the oxidation states in a neutral compound is zero. Since Group 1 metals always have an oxidation state of +1 in their compounds, it follows that the hydrogen must have an

oxidation state of -1 (+1 -1 = 0). Oxygen in peroxides Peroxides include hydrogen peroxide, H2O2. This is an electrically neutral compound and so the sum of the oxidation states of the hydrogen and oxygen must be zero. Since each hydrogen has an oxidation state of +1, each oxygen must have an oxidation state of -1 to balance it. Oxygen in F2O The problem here is that oxygen isn't the most electronegative element. The fluorine is more electronegative and has an oxidation state of -1. In this case, the oxygen has an oxidation state of +2. Chlorine in compounds with fluorine or oxygen There are so many different oxidation states that chlorine can have in these, that it is safer to simply remember that the chlorine doesn't have an oxidation state of -1 in them, and work out its actual oxidation state when you need it. You will find an example of this below. Warning! Don't get too bogged down in these exceptions. In most of the cases you will come across, they don't apply! Examples of working out oxidation states What is the oxidation state of chromium in Cr2+? That's easy! For a simple ion like this, the oxidation state is the charge on the ion - in other words: +2 (Don't forget the + sign.) What is the oxidation state of chromium in CrCl3? This is a neutral compound so the sum of the oxidation states is zero. Chlorine has an oxidation state of -1. If the oxidation state of chromium is n: n + 3(-1) = 0 n = +3 (Again, don't forget the + sign!) What is the oxidation state of chromium in Cr(H2O)63+? This is an ion and so the sum of the oxidation states is equal to the charge on the ion. There is a short-cut for working out oxidation states in complex ions like this where the metal atom is surrounded by electrically neutral molecules like water or ammonia. The sum of the oxidation states in the attached neutral molecule must be zero. That means that you can ignore them when you do the sum. This would be essentially the same as an unattached chromium ion, Cr3+. The

oxidation state is +3. What is the oxidation state of chromium in the dichromate ion, Cr2O72-? The oxidation state of the oxygen is -2, and the sum of the oxidation states is equal to the charge on the ion. Don't forget that there are 2 chromium atoms present. 2n + 7(-2) = -2 n = +6
Warning: Because these are simple sums it is tempting to try to do them in your head. If it matters (like in an exam) write them down using as many steps as you need so that there is no chance of making careless mistakes. Your examiners aren't going to be impressed by your mental arithmetic - all they want is the right answer! If you want some more examples to practice on, you will find them in most text books, including my chemistry calculations book.

What is the oxidation state of copper in CuSO4? Unfortunately, it isn't always possible to work out oxidation states by a simple use of the rules above. The problem in this case is that the compound contains two elements (the copper and the sulphur) whose oxidation states can both change. The only way around this is to know some simple chemistry! There are two ways you might approach it. (There might be others as well, but I can't think of them at the moment!)

You might recognise this as an ionic compound containing copper ions and sulphate ions, SO42-. To make an electrically neutral compound, the copper must be present as a 2+ ion. The oxidation state is therefore +2. You might recognise the formula as being copper(II) sulphate. The "(II)" in the name tells you that the oxidation state is 2 (see below). You will know that it is +2 because you know that metals form positive ions, and the oxidation state will simply be the charge on the ion.

Using oxidation states In naming compounds You will have come across names like iron(II) sulphate and iron(III) chloride. The (II) and (III) are the oxidation states of the iron in the two compounds: +2 and +3 respectively. That tells you that they contain Fe2+ and Fe3+ions. This can also be extended to the negative ion. Iron(II) sulphate is FeSO4. There is also a compound FeSO3 with the old name of iron(II) sulphite. The modern names reflect the oxidation states of the sulphur in the two

compounds. The sulphate ion is SO42-. The oxidation state of the sulphur is +6 (work it out!). The ion is more properly called the sulphate(VI) ion. The sulphite ion is SO32-. The oxidation state of the sulphur is +4 (work that out as well!). This ion is more properly called the sulphate(IV) ion. The ate ending simply shows that the sulphur is in a negative ion. So FeSO4 is properly called iron(II) sulphate(VI), and FeSO3 is iron(II) sulphate(IV). In fact, because of the easy confusion between these names, the old names sulphate and sulphite are normally still used in introductory chemistry courses.
Note: Even these aren't the full name! The oxygens in the negative ions should also be identified. FeSO 4 is properly called iron(II) tetraoxosulphate(VI). It all gets a bit out of hand for everyday use for common ions.

Using oxidation states to identify what's been oxidised and what's been reduced This is easily the most common use of oxidation states. Remember: Oxidation involves an increase in oxidation state Reduction involves a decrease in oxidation state In each of the following examples, we have to decide whether the reaction involves redox, and if so what has been oxidised and what reduced. Example 1: This is the reaction between magnesium and hydrochloric acid or hydrogen chloride gas:

Have the oxidation states of anything changed? Yes they have - you have two elements which are in compounds on one side of the equation and as uncombined elements on the other. Check all the oxidation states to be sure:.

The magnesium's oxidation state has increased - it has been oxidised. The hydrogen's oxidation state has fallen it has been reduced. The chlorine is in the same oxidation state on both sides of the equation - it hasn't been oxidised or reduced.

Example 2: The reaction between sodium hydroxide and hydrochloric acid is:

Checking all the oxidation states:

Nothing has changed. This isn't a redox reaction. Example 3: This is a sneaky one! The reaction between chlorine and cold dilute sodium hydroxide solution is:

Obviously the chlorine has changed oxidation state because it has ended up in compounds starting from the original element. Checking all the oxidation states shows:

The chlorine is the only thing to have changed oxidation state. Has it been oxidised or reduced? Yes! Both! One atom has been reduced because its oxidation state has fallen. The other has been oxidised. This is a good example of a disproportionation reaction. A disproportionation reaction is one in which a single substance is both oxidised and reduced. Using oxidation states to work out reacting proportions This is sometimes useful where you have to work out reacting proportions for use in titration reactions where you don't have enough information to work out the complete ionic equation. Remember that each time an oxidation state changes by one unit, one electron has been transferred. If one substance's oxidation state in a reaction falls by 2, that means that it has gained 2 electrons. Something else in the reaction must be losing those electrons. Any oxidation state fall by one substance must be

accompanied by an equal oxidation state increase by something else. This example is based on information in an old AQA A' level question. Ions containing cerium in the +4 oxidation state are oxidising agents. (They are more complicated than just Ce4+.) They can oxidise ions containing molybdenum from the +2 to the +6 oxidation state (from Mo2+ to MoO42-). In the process the cerium is reduced to the +3 oxidation state (Ce3+). What are the reacting proportions? The oxidation state of the molybdenum is increasing by 4. That means that the oxidation state of the cerium must fall by 4 to compensate. But the oxidation state of the cerium in each of its ions only falls from +4 to +3 - a fall of 1. So there must obviously be 4 cerium ions involved for each molybdenum ion. The reacting proportions are 4 cerium-containing ions to 1 molybdenum ion.

PERIOD 3 (Na to Ar) MENU

Physical and atomic properties of the elements . . . Discusses the trends in first ionisation energy, atomic radius, electronegativity, conductivity, melting point and boiling point as you go across the period from sodium to argon. Reactions of the Period 3 elements . . . The reactions of the elements from sodium to argon with water, oxygen and chlorine. Physical properties of the oxides . . . Discusses the link between the physical properties of the oxides of the elements from sodium to chlorine and their structure and bonding. Acid-base behaviour of the oxides . . . Reactions of the oxides with water, and their acid-base behaviour. Chlorides of the Period 3 elements. . . The physical properties of the chlorides of the elements from sodium to sulphur, their structure and bonding, and their reactions with water. Hydroxides of the Period 3 elements . . . The acid-base behaviour of the "hydroxides" (using the term in the widest possible sense) of the Period 3 elements.

ATOMIC AND PHYSICAL PROPERTIES OF THE PERIOD 3 ELEMENTS


This page describes and explains the trends in atomic and physical properties of the Period 3 elements from sodium to argon. It covers ionisation energy, atomic radius, electronegativity, electrical conductivity, melting point and boiling point. These topics are covered in various places elsewhere on the site and this page simply brings everything together - with links to the original pages if you need more information about particular points.

Atomic Properties
Electronic structures In Period 3 of the Periodic Table, the 3s and 3p orbitals are filling with electrons. Just as a reminder, the shortened versions of the electronic structures for the eight elements are: Na [Ne] 3s1 Mg [Ne] 3s2 Al Si P S Cl Ar [Ne] 3s2 3px1 [Ne] 3s2 3px1 3py1 [Ne] 3s2 3px1 3py1 3pz1 [Ne] 3s2 3px2 3py1 3pz1 [Ne] 3s2 3px2 3py2 3pz1 [Ne] 3s2 3px2 3py2 3pz2

In each case, [Ne] represents the complete electronic structure of a neon atom.

Note: If you aren't happy about electronic structures, it is essential to follow this link before you go any further. Use the BACK button on your browser to return quickly to this page.

First ionisation energy 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+.

It is the energy needed to carry out this change per mole of X. The pattern of first ionisation energies across Period 3

Notice that the general trend is upwards, but this is broken by falls between magnesium and aluminium, and between phosphorus and sulphur. Explaining the pattern First ionisation energy is governed by:

the charge on the nucleus; the distance of the outer electron from the nucleus; the amount of screening by inner electrons; whether the electron is alone in an orbital or one of a pair.

Note: If you aren't certain about the reasons for any of these statements, you must go and read the page aboutionisation energies before you go any further. Use the BACK button on your browser to return to this page.

The upward trend In the whole of period 3, the outer electrons are in 3-level orbitals. These are all the same sort of distances from the nucleus, and are screened by the same electrons in the first and second levels. The major difference is the increasing number of protons in the nucleus as you go from sodium across to argon. 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. The fall at aluminium You might expect the aluminium value to be more than the magnesium value because of the extra proton. Offsetting that is the fact that aluminium's outer electron is in a 3p orbital rather than a 3s. The 3p electron is slightly more distant from the nucleus than the 3s, and partially screened by the 3s electrons as well as the inner electrons. Both of these factors offset the effect of the extra proton. The fall at sulphur As you go from phosphorus to sulphur, something extra must be offsetting the effect of the extra proton The screening is identical in phosphorus and sulphur (from the inner electrons and, to some extent, from the 3s electrons), and the electron is being removed from an identical orbital. The difference is that in the sulphur case the electron being removed is one of the 3px2 pair. The repulsion between the two electrons in the same orbital means that the electron is easier to remove than it would otherwise be. Atomic radius The trend The diagram shows how the atomic radius changes as you go across Period 3.

The figures used to construct this diagram are based on:


metallic radii for Na, Mg and Al; covalent radii for Si, P, S and Cl; the van der Waals radius for Ar because it doesn't form any strong bonds.

It is fair to compare metallic and covalent radii because they are both being measured in tightly bonded

circumstances. It isn't fair to compare these with a van der Waals radius, though. The general trend towards smaller atoms across the period is NOT broken at argon. You aren't comparing like with like. The only safe thing to do is to ignore argon in the discussion which follows.

Note: If you aren't sure about the way that atomic radii are measured, it is essential to follow this link before you go any further. Use the BACK button on your browser to return to this page.

Explaining the trend A metallic or covalent radius is going to be a measure of the distance from the nucleus to the bonding pair of electrons. If you aren't sure about that, go back and follow the last link. From sodium to chlorine, the bonding electrons are all in the 3-level, being screened by the electrons in the first and second levels. The increasing number of protons in the nucleus as you go across the period pulls the bonding electrons more tightly to it. 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 3s electrons in the cases of the elements from aluminium to chlorine 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 3s 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.

Electronegativity Electronegativity is a measure of the tendency of an atom to attract a bonding pair of electrons. The Pauling scale is the most commonly used. Fluorine (the most electronegative element) is assigned a value of 4.0, and values range down to caesium and francium which are the least electronegative at 0.7. The trend

The trend across Period 3 looks like this:

Notice that argon isn't included. Electronegativity is about the tendency of an atom to attract a bonding pair of electrons. Since argon doesn't form covalent bonds, you obviously can't assign it an electronegativity. Explaining the trend The trend is explained in exactly the same way as the trend in atomic radii. As you go across the period, the bonding electrons are always in the same level - the 3-level. They are always being screened by the same inner electrons. All that differs is the number of protons in the nucleus. As you go from sodium to chlorine, the number of protons steadily increases and so attracts the bonding pair more closely.

Note: If you want a more detailed discussion of electronegativity, follow this link to the bonding section of the site. Use the BACK button on your browser to return to this page.

Physical Properties
This section is going to look at the electrical conductivity and the melting and boiling points of the elements. To understand these, you first have to understand the structure of each of the elements. Structures of the elements The structures of the elements change as you go across the period. The first three are metallic, silicon is giant covalent, and the rest are simple molecules. Three metallic structures Sodium, magnesium and aluminium all have metallic structures. In sodium, only one electron per atom is involved in the metallic bond - the single 3s electron. In magnesium, both of its outer electrons are involved, and in aluminium all three.

Note: If you aren't sure about metallic bonding, you must follow this link before you go on. Look also at the further link to the structures of metals that you will find at the bottom of that page. Use the BACK button (or GO menu or HISTORY file) on your browser to return to this page when you are ready.

The other difference you need to be aware of is the way the atoms are packed in the metal crystal. Sodium is 8-co-ordinated - each sodium atom is touched by only 8 other atoms. Both magnesium and aluminium are 12-co-ordinated (although in slightly different ways). This is a more efficient way to pack atoms, leading to less wasted space in the metal structures and to stronger bonding in the metal.

Note: If this talk about co-ordination doesn't mean anything to you, you need to look at the page about metallic structures where it is explained in some detail. Use the BACK button on your browser to return to this page.

A giant covalent structure

Silicon has a giant covalent structure just like diamond. A tiny part of the structure looks like this:

The structure is held together by strong covalent bonds in all three dimensions. Four simple molecular structures The structures of phosphorus and sulphur vary depending on the type of phosphorus or sulphur you are talking about. For phosphorus, I am assuming the common white phosphorus. For sulphur, I am assuming one of the crystalline forms - rhombic or monoclinic sulphur.

The atoms in each of these molecules are held together by covalent bonds (apart, of course, from argon). In the liquid or solid state, the molecules are held close to each other by van der Waals dispersion forces.

Note: You will find van der Waals dispersion forces described in great detail if you follow this link Use the BACK button on your browser to return to this page.

Electrical conductivity

Sodium, magnesium and aluminium are all good conductors of electricity. Conductivity increases as you go from sodium to magnesium to aluminium.

Silicon is a semiconductor. None of the rest conduct electricity.

The three metals, of course, conduct electricity because the delocalised electrons (the "sea of electrons") are free to move throughout the solid or the liquid metal. In the silicon case, explaining how semiconductors conduct electricity is beyond the scope of A level chemistry courses. With a diamond structure, you mightn't expect it to conduct electricity, but it does! The rest don't conduct electricity because they are simple molecular substances. There are no electrons free to move around. Melting and boiling points The chart shows how the melting and boiling points of the elements change as you go across the period. The figures are plotted in kelvin rather than C to avoid having negative values.

It is best to think of these changes in terms of the types of structure that we have talked about further up the page. The metallic structures Melting and boiling points rise across the three metals because of the increasing strength of the metallic bonds. The number of electrons which each atom can contribute to the delocalised "sea of electrons" increases. The atoms also get smaller and have more protons as you go from sodium to magnesium to aluminium. The attractions and therefore the melting and boiling points increase because:

The nuclei of the atoms are getting more positively charged. The "sea" is getting more negatively charged. The "sea" is getting progressively nearer to the nuclei and so more strongly attracted.

Note: Boiling point is a better guide to the strength of the metallic bonds than melting point. Metallic bonds still exist in the liquid metals and aren't completely broken until the metal boils. I don't know why there is such a small increase in melting point as you go from magnesium to

aluminium. The boiling point of aluminium is much higher than magnesium's - as you would expect.

Silicon Silicon has high melting and boiling points because it is a giant covalent structure. You have to break strong covalent bonds before it will melt or boil. Because you are talking about a different type of bond, it isn't profitable to try to directly compare silicon's melting and boiling points with aluminium's. The 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 sizes of the melting and boiling points are governed entirely by the sizes of the molecules. Remember the structures of the molecules:

Phosphorus Phosphorus contains P4 molecules. To melt phosphorus you don't have to break any covalent bonds - just the much weaker van der Waals forces between the molecules. Sulphur Sulphur consists of S8 rings of atoms. The molecules are bigger than phosphorus molecules, and so the van der Waals attractions will be stronger, leading to a higher melting and boiling point. Chlorine Chlorine, Cl2, is a much smaller molecule with comparatively weak van der Waals attractions, and so chlorine will have a lower melting and boiling point than sulphur or phosphorus. Argon

Argon molecules are just single argon atoms, Ar. The scope for van der Waals attractions between these is very limited and so the melting and boiling points of argon are lower again.

CHEMICAL REACTIONS OF THE PERIOD 3 ELEMENTS


This page describes the reactions of the Period 3 elements from sodium to argon with water, oxygen and chlorine.

Reactions with water


Sodium Sodium has a very exothermic reaction with cold water producing hydrogen and a colourless solution of sodium hydroxide.

Magnesium Magnesium has a very slight reaction with cold water, but burns in steam. A very clean coil of magnesium dropped into cold water eventually gets covered in small bubbles of hydrogen which float it to the surface. Magnesium hydroxide is formed as a very thin layer on the magnesium and this tends to stop the reaction.

Magnesium burns in steam with its typical white flame to produce white magnesium oxide and hydrogen.

Note: If you are heating the magnesium in a glass tube, the magnesium also reacts with the glass. That leaves dark grey products (including silicon and perhaps boron from the glass) as well as the white magnesium oxide. Notice also that the oxide is produced on heating in steam. Hydroxides are only ever produced using liquid water.

Aluminium Aluminium powder heated in steam produces hydrogen and aluminium oxide. The reaction is relatively slow because of the existing strong aluminium oxide layer on the metal, and the build-up of even more oxide during the reaction.

Silicon There is a fair amount of disagreement in the books and on the web about what silicon does with water or steam. The truth seems to depend on the precise form of silicon you are using. The common shiny grey lumps of silicon with a rather metal-like appearance are fairly unreactive. Most sources suggest that this form of silicon will react with steam at red heat to produce silicon dioxide and hydrogen.

But it is also possible to make much more reactive forms of silicon which will react with cold water to give the same products.

Note: These more reactive forms are produced as powders. Cotton and Wilkinson's Advanced Inorganic Chemistry (third edition - page 316) suggests that the reactivity of one of these could be due to a very high surface area, or perhaps because the silicon exists in a graphite-like structure. A correspondent from the silicon industry tells me that when silicon is cut into slices, the silicon dust formed reacts with water at room temperature - producing hydrogen and getting very hot. He says "The silicon is cut in a glycol slurry [. . .] The powdered Si is protected somewhat from moisture in the glycol slurry, but when we clean the slurry in aqueous solutions the reaction with water takes off." This is probably the effect of the high surface area of the dust produced, combined with the fact that you are exposing uncontaminated silicon to the water. One source suggests that the lack of reactivity of silicon is due to a layer of silicon dioxide on its surface. If you expose a new surface by cutting the silicon, that layer won't, of course, exist.

Phosphorus and sulphur These have no reaction with water. Chlorine Chlorine dissolves in water to some extent to give a green solution. A reversible reaction takes place to produce a mixture of hydrochloric acid and chloric(I) acid (hypochlorous acid).

Note: You may also find the chloric(I) acid written as HClO. The form I have used more accurately reflects the way the atoms are joined up. It doesn't matter which you use.

In the presence of sunlight, the chloric(I) acid slowly decomposes to produce more hydrochloric acid, releasing oxygen gas, and you may come across an equation showing the overall change:

Argon There is no reaction between argon and water.

Reactions with oxygen


Sodium Sodium burns in oxygen with an orange flame to produce a white solid mixture of sodium oxide and sodium peroxide. For the simple oxide:

For the peroxide:

Magnesium Magnesium burns in oxygen with an intense white flame to give white solid magnesium oxide.

Note: If magnesium is burns in air rather than in pure oxygen, it also reacts with the nitrogen in the air. You get a mixture of magnesium oxide and magnesium nitride formed.

Aluminium Aluminium will burn in oxygen if it is powdered, otherwise the strong oxide layer on the aluminium tends to inhibit the reaction. If you sprinkle aluminium powder into a Bunsen flame, you get white sparkles. White aluminium oxide is formed.

Silicon Silicon will burn in oxygen if heated strongly enough. Silicon dioxide is produced.

Note: There is disagreement between various web or textbook sources about the temperature needed to ignite the silicon, varying from 400C to well over 1000C. In fact, there isn't a "right" answer to this. It depends on what sort of silicon you are talking about and how finely divided it is. For example, one of the amorphous (noncrystalline powder) forms of silicon even catches fire spontaneously in air at room temperature. Other forms need higher temperatures and a richer oxygen supply.

Phosphorus White phosphorus catches fire spontaneously in air, burning with a white flame and producing clouds of white smoke - a mixture of phosphorus(III) oxide and phosphorus(V) oxide. The proportions of these depend on the amount of oxygen available. In an excess of oxygen, the product will be almost entirely phosphorus(V) oxide. For the phosphorus(III) oxide:

For the phosphorus(V) oxide:

Note: You may come across these oxides written as P2O3 and P2O5. Don't use these forms! They are as logical as writing, say, ethene as CH2 and ethane as CH3.

Sulphur Sulphur burns in air or oxygen on gentle heating with a pale blue flame. It produces colourless sulphur dioxide gas.

Note: Sulphur dioxide can, of course, be converted further into sulphur trioxide in the presence of oxygen, but it needs the presence of a catalyst and fairly carefully controlled conditions. If you are interested in this, see the page on the Contact Process. This isn't particularly relevant to the current topic, but if you do feel the urge to follow this link, use the BACK button on your browser to return to this page.

Chlorine and argon Despite having several oxides, chlorine won't react directly with oxygen. Argon doesn't react either.

Reactions with chlorine


Sodium Sodium burns in chlorine with a bright orange flame. White solid sodium chloride is produced.

Magnesium Magnesium burns with its usual intense white flame to give white magnesium chloride.

Aluminium Aluminium is often reacted with chlorine by passing dry chlorine over aluminium foil heated in a long tube. The aluminium burns in the stream of chlorine to produce very pale yellow aluminium chloride. This sublimes (turns straight from solid to vapour and back again) and collects further down the tube where it is cooler.

Note: You may find versions of this equation showing the aluminium chloride as Al2Cl6. In fact, this exists in the vapour at temperatures not too far above the sublimation temperature - not in the solid. The structure of aluminium chloride is discussed on the page about Period 3 chlorides. If you follow this link, use the BACK button on your browser to return to this page.

Silicon If chlorine is passed over silicon powder heated in a tube, it reacts to produce silicon tetrachloride. This is a colourless liquid which vaporises and can be condensed further along the apparatus.

Phosphorus White phosphorus burns in chlorine to produce a mixture of two chlorides, phosphorus(III) chloride and phosphorus(V) chloride (phosphorus trichloride and phosphorus pentachloride). Phosphorus(III) chloride is a colourless fuming liquid.

Phosphorus(V) chloride is an off-white (going towards yellow) solid.

Note: These equations are often given starting from P rather than P4. It depends which form of phosphorus you are talking about. If you are talking about white phosphorus (as I am here), P4 is the correct version. If you are talking about red phosphorus, then P is correct. Red phosphorus has a different (polymeric) structure, and P 4 would be wrong for it. In my experience, red phosphorus is less commonly used in labs at this level (it isn't as excitingly reactive as white phosphorus!) - which is why I am concentrating on the white form.

Sulphur If a stream of chlorine is passed over some heated sulphur, it reacts to form an orange, evil-smelling liquid, disulphur dichloride, S2Cl2.

Chlorine and argon It obviously doesn't make sense to talk about chlorine reacting with itself, and argon doesn't react with chlorine.

PHYSICAL PROPERTIES OF THE PERIOD 3 OXIDES


This page explains the relationship between the physical properties of the oxides of Period 3 elements (sodium to chlorine) and their structures. Argon is obviously omitted because it doesn't form an oxide.

A quick summary of the trends


The oxides The oxides we'll be looking at are: Na2O MgO Al2O3 SiO2 P4O10 SO3 Cl2O7

P4O6

SO2

Cl2O

Those oxides in the top row are known as the highest oxides of the various elements. These are the oxides where the Period 3 elements are in their highest oxidation states. In these oxides, all the outer electrons in the Period 3 element are being involved in the bonding - from just the one with sodium, to all seven of chlorine's outer electrons.

Note: If you aren't sure about oxidation states (oxidation numbers) you will find them discussed in if you follow this link. Use the BACK button on your browser to return quickly to this page later.

The structures The trend in structure is from the metallic oxides containing giant structures of ions on the left of the period via a giant covalent oxide (silicon dioxide) in the middle to molecular oxides on the right. Melting and boiling points The giant structures (the metal oxides and silicon dioxide) will have high melting and boiling points because a lot of energy is needed to break the strong bonds (ionic or covalent) operating in three dimensions. The oxides of phosphorus, sulphur and chlorine consist of individual molecules - some small and simple; others polymeric. The attractive forces between these molecules will be van der Waals dispersion and dipole-dipole interactions. These vary in size depending on the size, shape and polarity of the various molecules - but will always be much

weaker than the ionic or covalent bonds you need to break in a giant structure. These oxides tend to be gases, liquids or low melting point solids. Electrical conductivity None of these oxides has any free or mobile electrons. That means that none of them will conduct electricity when they are solid. The ionic oxides can, however, undergo electrolysis when they are molten. They can conduct electricity because of the movement of the ions towards the electrodes and the discharge of the ions when they get there.

Warning: The rest of this page contains quite a lot of detail about the structures of the various oxides. Don't lose sight of the overall trends in the period when you are looking at all this detail.

The metallic oxides


The structures Sodium, magnesium and aluminium oxides consist of giant structures containing metal ions and oxide ions. Magnesium oxide has a structure just like sodium chloride. The other two have more complicated arrangements of the ions beyond the scope of syllabuses at this level (UK A level or its equivalents).

Note: You can find the structure of sodium chloride (the same as the magnesium oxide structure) by following this link. Use the BACK button on your browser to return quickly to this page later.

Melting and boiling points There are strong attractions between the ions in each of these oxides and these attractions need a lot of heat energy to break. These oxides therefore have high melting and boiling points.

Problems! I intended at this point to quote values for each of the oxides, hoping to show that the melting and boiling points increase as the charges on the positive ion increase from 1+ in sodium to 3+ in aluminium. You would expect that the greater the charge, the greater the attractions. Unfortunately, the oxide with the highest

melting and boiling point is magnesium oxide, not aluminium oxide! So that theory bit the dust! The reason for this probably lies in the increase in electronegativity as you go from sodium to magnesium to aluminium. That would mean that the electronegativity difference between the metal and the oxygen is decreasing. The smaller difference means that the bond won't be so purely ionic. It is also likely that molten aluminium oxide contains complex ions containing both aluminium and oxygen rather than simple aluminium and oxide ions. All this means, of course, that you aren't really comparing like with like - so wouldn't necessarily expect a neat trend. The other problems I came across lie with sodium oxide. Most sources say that this sublimes (turns straight from solid to vapour) at 1275C. However, the usually reliable Webelements gives a melting point of 1132C followed by a decomposition temperature (before boiling) of 1950C. Other sources talk about it decomposing (to sodium and sodium peroxide) above 400C. I have no idea what the truth of this is - although I suspect that the Webelements melting point value is probably for a pressure above atmospheric pressure (although it doesn't say so).

Electrical conductivity None of these conducts electricity in the solid state, but electrolysis is possible if they are molten. They conduct electricity because of the movement and discharge of the ions present. The only important example of this is in the electrolysis of aluminium oxide in the manufacture of aluminium. Whether you can electrolyse molten sodium oxide depends, of course, on whether it actually melts instead of subliming or decomposing under ordinary circumstances. If it sublimes, you won't get any liquid to electrolyse! Magnesium and aluminium oxides have melting points far too high to be able to electrolyse them in a simple lab.

Note: You will find full details of the electrolysis of aluminium oxide during the extraction of aluminium if you follow this link. Use the BACK button on your browser to return to this page later.

Silicon dioxide (silicon(IV) oxide)


The structure The electronegativity of the elements increases as you go across the period, and by the time you get to silicon, there isn't enough electronegativity difference between the silicon and the oxygen to form an ionic bond. Silicon dioxide is a giant covalent structure.

Note: If you aren't happy about electronegativity you will find it explained if you follow this link. Use the BACK button on your browser to return quickly to this page later.

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 in 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. If you need help in drawing this structure you will find a suggestion by following this link

Use the BACK button on your browser to return quickly to this page later.

Melting and boiling points Silicon dioxide has a high melting point - varying depending on what the particular structure is (remember that the structure given is only one of three possible structures), but they are all around 1700C. Very strong siliconoxygen covalent bonds have to be broken throughout the structure before melting occurs. Silicon dioxide boils at 2230C. Because you are talking about a different form of bonding, it doesn't make sense to try to compare these values directly with the metallic oxides. What you can safely say is that because the metallic oxides and silicon dioxide have giant structures, the melting and boiling points are all high. Electrical conductivity Silicon dioxide doesn't have any mobile electrons or ions - so it doesn't conduct electricity either as a solid or a liquid.

The molecular oxides


Phosphorus, sulphur and chlorine all form oxides which consist of molecules. Some of these molecules are fairly simple - others are polymeric. We are just going to look at some of the simple ones. Melting and boiling points of these oxides will be much lower than those of the metal oxides or silicon dioxide. The intermolecular forces holding one molecule to its neighbours will be van der Waals dispersion forces or dipoledipole interactions. The strength of these will vary depending on the size of the molecules. None of these oxides conducts electricity either as solids or as liquids. None of them contains ions or free electrons. The phosphorus oxides Phosphorus has two common oxides, phosphorus(III) oxide, P4O6, and phosphorus(V) oxide, P4O10. Phosphorus(III) oxide Phosphorus(III) oxide is a white solid, melting at 24C and boiling at 173C. The structure of its molecule is best worked out starting from a P4 molecule which is a little tetrahedron.

Pull this apart so that you can see the bonds . . .

. . . and then replace the bonds by new bonds linking the phosphorus atoms via oxygen atoms. These will be in a V-shape (rather like in water), but you probably wouldn't be penalised if you drew them on a straight line between the phosphorus atoms in an exam.

The phosphorus is using only three of its outer electrons (the 3 unpaired p electrons) to form bonds with the oxygens. Phosphorus(V) oxide Phosphorus(V) oxide is also a white solid, subliming (turning straight from solid to vapour) at 300C. In this case, the phosphorus uses all five of its outer electrons in the bonding. Solid phosphorus(V) oxide exists in several different forms - some of them polymeric. We are going to concentrate on a simple molecular form, and this is also present in the vapour. This is most easily drawn starting from P4O6. The other four oxygens are attached to the four phosphorus atoms via double bonds.

Note: If you look carefully, the shape of this molecule looks very much like the way we usually draw the repeating unit in the diamond giant structure. Don't confuse the two, though! The P4O10 molecule stops here. This isn't a little bit of a giant structure - it's all there is. In diamond, of course, the structure just continues almost endlessly in three dimensions.

The sulphur oxides Sulphur has two common oxides, sulphur dioxide (sulphur(IV) oxide), SO2, and sulphur trioxide (sulphur(VI) oxide), SO3. Sulphur dioxide Sulphur dioxide is a colourless gas at room temperature with an easily recognised choking smell. It consists of simple SO2 molecules.

The sulphur uses 4 of its outer electrons to form the double bonds with the oxygen, leaving the other two as a lone pair on the sulphur. The bent shape of SO2 is due to this lone pair. Sulphur trioxide Pure sulphur trioxide is a white solid with a low melting and boiling point. It reacts very rapidly with water vapour in the air to form sulphuric acid. That means that if you make some in the lab, you tend to see it as a white sludge

which fumes dramatically in moist air (forming a fog of sulphuric acid droplets). Gaseous sulphur trioxide consists of simple SO3 molecules in which all six of the sulphur's outer electrons are involved in the bonding.

There are various forms of solid sulphur trioxide. The simplest one is a trimer, S3O9, where three SO3 molecules are joined up and arranged in a ring.

There are also other polymeric forms in which the SO3 molecules join together in long chains. For example:

Note: It is difficult to draw this convincingly. In fact, on each sulphur atom, one of the double bonded oxygens is coming out of the diagram towards you, and the other one is going back in away from you.

The fact that the simple molecules join up in this way to make bigger structures is what makes the sulphur trioxide a solid rather than a gas.

Note: It is pretty unlikely that you will need these solid structures for the purposes of UK A level or its equivalents. If in doubt, check your syllabus and past papers. If you don't have either of these follow this link to find out how to get them.

The chlorine oxides Chlorine forms several oxides. Here we are just looking at two of them (the only ones mentioned by any of the UK syllabuses) - chlorine(I) oxide, Cl2O, and chlorine(VII) oxide, Cl2O7. Chlorine(I) oxide Chlorine(I) oxide is a yellowish-red gas at room temperature. It consists of simple small molecules.

There's nothing in the least surprising about this molecule and it's physical properties are just what you would expect for a molecule this size. Chlorine(VII) oxide In chlorine(VII) oxide, the chlorine uses all of its seven outer electrons in bonds with oxygen. This produces a much bigger molecule, and so you would expect its melting point and boiling point to be higher than chlorine(I) oxide. Chlorine(VII) oxide is a colourless oily liquid at room temperature. In the diagram, for simplicity I have drawn a standard structural formula. In fact, the shape is tetrahedral around both chlorines, and V-shaped around the central oxygen.

ACID-BASE BEHAVIOUR OF THE PERIOD 3 OXIDES


This page looks at the reactions of the oxides of Period 3 elements (sodium to chlorine) with water, and with acids or bases where relevant. Argon is obviously omitted because it doesn't form an oxide.

A quick summary of the trend

The oxides The oxides we'll be looking at are: Na2O MgO Al2O3 SiO2 P4O10 SO3 Cl2O7

P4O6

SO2

Cl2O

Note: If you haven't already been there, you might be interested in looking at the page about the structures and physical properties of the Period 3 oxides as a useful introduction before you go any further. Use the BACK button on your browser to return quickly to this page later if you choose to follow this link.

The trend in acid-base behaviour The trend in acid-base behaviour is shown in various reactions, but as a simple summary:

The trend is from strongly basic oxides on the left-hand side to strongly acidic ones on the right, via an amphoteric oxide (aluminium oxide) in the middle. An amphoteric oxide is one which shows both acidic and basic properties.

For this simple trend, you have to be looking only at the highest oxides of the individual elements. Those are the ones on the top row above, and are where the element is in its highest possible oxidation state. The pattern isn't so simple if you include the other oxides as well. For the non-metal oxides, their acidity is usually thought of in terms of the acidic solutions formed when they react with water - for example, sulphur trioxide reacting to give sulphuric acid. They will, however, all react with bases such as sodium hydroxide to form salts such as sodium sulphate. These reactions are all explored in detail on the rest of this page.

Warning: The rest of this page contains quite a lot of detail about the various oxides. Don't lose sight of the overall trend in the period with respect to the highest oxides when you are looking at all this detail. It is essential to know what your syllabus says about this topic, and to explore past papers and mark schemes otherwise you are going to end up bogged down in a mass of detail that you don't actually need to know about. If you are working towards a UK-based exam (A level or its equivalent) and haven't got any of these things follow

this link before you go any further to find out how to get them.

Chemistry of the individual oxides


Sodium oxide Sodium oxide is a simple strongly basic oxide. It is basic because it contains the oxide ion, O2-, which is a very strong base with a high tendency to combine with hydrogen ions. Reaction with water Sodium oxide reacts exothermically with cold water to produce sodium hydroxide solution. Depending on its concentration, this will have a pH around 14.

Reaction with acids As a strong base, sodium oxide also reacts with acids. For example, it would react with dilute hydrochloric acid to produce sodium chloride solution.

Magnesium oxide Magnesium oxide is again a simple basic oxide, because it also contains oxide ions. However, it isn't as strongly basic as sodium oxide because the oxide ions aren't so free. In the sodium oxide case, the solid is held together by attractions between 1+ and 2- ions. In the magnesium oxide case, the attractions are between 2+ and 2-. It takes more energy to break these. Even allowing for other factors (like the energy released when the positive ions form attractions with water in the solution formed), the net effect of this is that reactions involving magnesium oxide will always be less exothermic than those of sodium oxide. Reaction with water If you shake some white magnesium oxide powder with water, nothing seems to happen - it doesn't look as if it reacts. However, if you test the pH of the liquid, you find that it is somewhere around pH 9 - showing that it is slightly alkaline. There must have been some slight reaction with the water to produce hydroxide ions in solution. Some magnesium hydroxide is formed in the reaction, but this is almost insoluble - and so not many hydroxide ions actually get into solution.

Reaction with acids Magnesium oxide reacts with acids as you would expect any simple metal oxide to react. For example, it reacts

with warm dilute hydrochloric acid to give magnesium chloride solution.

Aluminium oxide Describing the properties of aluminium oxide can be confusing because it exists in a number of different forms. One of those forms is very unreactive. It is known chemically as alpha-Al2O3 and is produced at high temperatures. In what follows we are assuming one of the more reactive forms. Aluminium oxide is amphoteric. It has reactions as both a base and an acid. Reaction with water Aluminium oxide doesn't react in a simple way with water in the sense that sodium oxide and magnesium oxide do, and doesn't dissolve in it. Although it still contains oxide ions, they are held too strongly in the solid lattice to react with the water.

Note: Some forms of aluminium oxide do, however, absorb water very effectively. I haven't been able to establish whether this absorption just involves things like hydrogen bonds or whether an actual chemical reaction to produce some sort of hydroxide occurs. If you have any firm information on this, could you contact me via the address on the about this site page.

Reaction with acids Aluminium oxide contains oxide ions and so reacts with acids in the same way as sodium or magnesium oxides. That means, for example, that aluminium oxide will react with hot dilute hydrochloric acid to give aluminium chloride solution.

In this (and similar reactions with other acids), aluminium oxide is showing the basic side of its amphoteric nature. Reaction with bases Aluminium oxide has also got an acidic side to its nature, and it shows this by reacting with bases such as sodium hydroxide solution. Various aluminates are formed - compounds where the aluminium is found in the negative ion. This is possible because aluminium has the ability to form covalent bonds with oxygen. In the case of sodium, there is too much electronegativity difference between sodium and oxygen to form anything

other than an ionic bond. But electronegativity increases as you go across the period - and the electronegativity difference between aluminium and oxygen is smaller. That allows the formation of covalent bonds between the two.

Note: If you aren't happy about electronegativity you will find it explained if you follow this link. Use the BACK button on your browser to return quickly to this page later.

With hot, concentrated sodium hydroxide solution, aluminium oxide reacts to give a colourless solution of sodium tetrahydroxoaluminate.

Note: You may find all sorts of other formulae given for the product from this reaction. These range from NaAlO2 (which is a dehydrated form of the one in the equation) to Na 3Al(OH)6 (which is a different product altogether). What you actually get will depend on things like the temperature and the concentration of the sodium hydroxide solution. In any case, the truth is almost certainly a lot more complicated than any of these. This is a case where it is a good idea to find out what your examiners quote in their support material or mark schemes, and stick with that. If necessary, get this sort of information from your examiners (if you are doing a UK-based course) by following the links on the syllabuses page.

Silicon dioxide (silicon(IV) oxide) By the time you get to silicon as you go across the period, electronegativity has increased so much that there is no longer enough electronegativity difference between silicon and oxygen to form ionic bonds. Silicon dioxide has no basic properties - it doesn't contain oxide ions and it doesn't react with acids. Instead, it is very weakly acidic, reacting with strong bases. Reaction with water Silicon dioxide doesn't react with water, because of the difficulty of breaking up the giant covalent structure. Reaction with bases

Silicon dioxide reacts with sodium hydroxide solution, but only if it is hot and concentrated. A colourless solution of sodium silicate is formed.

You may also be familiar with one of the reactions happening in the Blast Furnace extraction of iron - in which calcium oxide (from the limestone which is one of the raw materials) reacts with silicon dioxide to produce a liquid slag, calcium silicate. This is also an example of the acidic silicon dioxide reacting with a base.

Important! For the remainder of the oxides, we are mainly going to be considering the results of reacting them with water to give solutions of various acids. When we talk about the acidity of the oxides increasing as you go from, say, phosphorus(V) oxide to sulphur trioxide to chlorine(VII) oxide, what we are normally talking about is the increasing strengths of the acids formed when they react with water. The phosphorus oxides We are going to be looking at two phosphorus oxides, phosphorus(III) oxide, P4O6, and phosphorus(V) oxide, P4O10. Phosphorus(III) oxide Phosphorus(III) oxide reacts with cold water to give a solution of the weak acid, H3PO3 - known variously as phosphorous acid, orthophosphorous acid or phosphonic acid. Its reaction with hot water is much more complicated.

Note: Notice the "-ous" ending in the first two names. That's not a spelling mistake - it's for real! It is used to distinguish it from phosphoric acid which is quite different (see below). The names of the phosphorus-containing acids are a bit of a nightmare! (In fact, as far as I'm concerned, the phosphorus acids in general have always been and continue to be a complete nightmare!) Don't get too worried about these names at this level. Just be sure that you can write the formulae if you need to - and be grateful that you don't need to know all that much else about them!

The pure un-ionised acid has the structure:

The hydrogens aren't released as ions until you add water to the acid, and even then not many are released because phosphorous acid is only a weak acid. Phosphorous acid has a pKa of 2.00 which makes it stronger than common organic acids like ethanoic acid (pKa= 4.76).

Note: If you should know about pKa, but aren't very confident, you could follow this link - but it is likely to take you a long time. All you really need to know for this topic is that the lower the pKa value, the stronger the acid.

It is pretty unlikely that you would ever react phosphorus(III) oxide directly with a base, but you might need to know what happens if you react the phosphorous acid formed with a base. In phosphorous acid, the two hydrogen atoms in the -OH groups are acidic, but the other one isn't. That means that you can get two possible reactions with, for example, sodium hydroxide solution depending on the proportions used.

In the first case, only one of the acidic hydrogens has reacted with the hydroxide ions from the base. In the second case (using twice as much sodium hydroxide), both have reacted. If you were to react phosphorus(III) oxide directly with sodium hydroxide solution rather than making the acid first, you would end up with the same possible salts.

Note: Check your syllabus, past papers and mark schemes before you get too bogged down in this! Follow this link to find out how to get hold of them if you haven't already got them (UK-based syllabuses only).

Phosphorus(V) oxide Phosphorus(V) oxide reacts violently with water to give a solution containing a mixture of acids, the nature of which depends on the conditions. We usually just consider one of these, phosphoric(V) acid, H3PO4 - also known just as phosphoric acid or as orthophosphoric acid.

This time the pure un-ionised acid has the structure:

Phosphoric(V) acid is also a weak acid with a pKa of 2.15. That makes it fractionally weaker than phosphorous acid. Solutions of both of these acids of concentrations around 1 mol dm-3 will have a pH of about 1. Once again, you are unlikely ever to react this oxide with a base, but you may well be expected to know how phosphoric(V) acid reacts with something like sodium hydroxide solution. If you look back at the structure, you will see that it has three -OH groups, and each of these has an acidic hydrogen atom. You can get a reaction with sodium hydroxide in three stages, with one after another of these hydrogens reacting with the hydroxide ions.

Again, if you were to react phosphorus(V) oxide directly with sodium hydroxide solution rather than making the acid first, you would end up with the same possible salts. This is getting ridiculous, and so I will only give one example out of the possible equations:

Note: If you get a question in an exam which just asks you to write an equation for the reaction of sodium hydroxide with phosphoric(V) acid, which equation should you write? It shouldn't really matter - all of them are perfectly valid. In each case, it just depends on the proportions of the two reagents you are using. If you really want to be certain, check past papers and mark schemes. I found one question about the reaction between sodium oxide and phosphoric(V) acid where the mark scheme accepted any of the possible equations -

which is what I would expect. (I know I haven't given you that particular set of equations, but they aren't difficult to work out as long as you understand the principle, and I can't possibly give every single acid-base equation. This already long page would go on for ever, and everybody would give up in despair well before the end! That's why you are trying to understand chemistry rather than learn it parrot-fashion.) Please don't waste time learning equations - or at least, not until you know and understand all the rest of the chemistry that you need to know and understand! Any one equation stands a very small chance of coming up in an exam, even if it is on your particular syllabus. Life is too short to waste time learning equations. Know how to work them out if you need to.

The sulphur oxides We are going to be looking at sulphur dioxide, SO2, and sulphur trioxide, SO3. Sulphur dioxide Sulphur dioxide is fairly soluble in water, reacting with it to give a solution of sulphurous acid (sulphuric(IV) acid), H2SO3. This only exists in solution, and any attempt to isolate it just causes sulphur dioxide to be given off again.

The un-ionised acid has the structure:

Sulphurous acid is also a weak acid with a pKa of around 1.8 - very slightly stronger than the two phosphoruscontaining acids above. A reasonably concentrated solution of sulphurous acid will again have a pH of about 1.

Note: There is some variability in the pKa value quoted for sulphurous acid by various sources - ranging from 1.77 to 1.92. I have no way of knowing which of these is right.

Sulphur dioxide will also react directly with bases such as sodium hydroxide solution. If sulphur dioxide is bubbled through sodium hydroxide solution, sodium sulphite solution is formed first followed by sodium hydrogensulphite solution when the sulphur dioxide is in excess.

Note: Sodium sulphite is also called sodium sulphate(IV). Sodium hydrogensulphite is also sodium hydrogensulphate(IV) or sodium bisulphite. Notice that the equations for these reactions are different from the phosphorus examples. In this case, we are reacting the oxide directly with the sodium hydroxide, because that's the way we are most likely to do it. If you want to, it isn't difficult to work out equations for the reactions between sodium hydroxide and sulphurous acid in exactly the same way as we did with the phosphorus acids.

Another important reaction of sulphur dioxide is with the base calcium oxide to form calcium sulphite (calcium sulphate(IV)). This is at the heart of one of the methods of removing sulphur dioxide from flue gases in power stations.

Sulphur trioxide Sulphur trioxide reacts violently with water to produce a fog of concentrated sulphuric acid droplets.

Note: If you know about the Contact Process for the manufacture of sulphuric acid, you will know that the sulphur trioxide is always converted into sulphuric acid by a round-about process to avoid the problem of the sulphuric acid fog. You will find details of the Contact Process elsewhere on this site if you are interested, but it isn't relevant to the current topic.

Pure un-ionised sulphuric acid has the structure:

Sulphuric acid is a strong acid, and solutions will typically have pH's of around 0. The acid reacts with water to give a hydroxonium ion (a hydrogen ion in solution, if you like) and a hydrogensulphate ion. This reaction is virtually 100% complete.

The second hydrogen is more difficult to remove. In fact the hydrogensulphate ion is a relatively weak acid similar in strength to the acids we have already discussed on this page. This time you get an equilibrium:

You probably won't need this for the purposes of UK A level (or its equivalents), but it is useful if you understand the reason that sulphuric acid is a stronger acid than sulphurous acid. You can apply the same reasoning to other acids that you find on this page as well. When a hydrogen ion is lost from one of the -OH groups, the negative charge left on the oxygen is spread out (delocalised) over the ion by interacting with the doubly-bonded oxygens. It follows that the more of these you have, the more delocalisation you can get - and the more delocalisation, the more stable the ion becomes. The more stable the ion, the less likely it is to recombine with a hydrogen ion and revert to the un-ionised acid. Sulphurous acid only has one doubly-bonded oxygen, whereas sulphuric acid has two - that makes for a much more effective delocalisation, a much more stable ion, and so for a stronger acid.

Note: This is just like the delocalisation which occurs in the ethanoate ion formed when ethanoic acid is behaving as a weak acid. You will find this described in some detail on a page about organic acids. Use the BACK button on your browser if you choose to follow this link.

Sulphuric acid, of course, has all the reactions of a strong acid that you are familiar with from introductory chemistry courses. For example, the normal reaction with sodium hydroxide solution is to form sodium sulphate solution - in which both of the acidic hydrogens react with hydroxide ions.

In principle, you can also get sodium hydrogensulphate solution by using half as much sodium hydroxide and just reacting with one of the two acidic hydrogens in the acid. In practice, I personally have never ever done it - I can't at the moment see much point! Sulphur trioxide itself will also react directly with bases to form sulphates. For example, it will react with calcium oxide to form calcium sulphate. This is just like the reaction with sulphur dioxide described above.

The chlorine oxides Chlorine forms several oxides, but the only two mentioned by any of the UK A level syllabuses are chlorine(VII) oxide, Cl2O7, and chlorine(I)oxide, Cl2O. Chlorine(VII) oxide is also known as dichlorine heptoxide, and chlorine(I) oxide as dichlorine monoxide. Chlorine(VII) oxide Chlorine(VII) oxide is the highest oxide of chlorine - the chlorine is in its maximum oxidation state of +7. It continues the trend of the highest oxides of the Period 3 elements towards being stronger acids. Chlorine(VII) oxide reacts with water to give the very strong acid, chloric(VII) acid - also known as perchloric acid. The pH of typical solutions will, like sulphuric acid, be around 0.

Un-ionised chloric(VII) acid has the structure:

When the chlorate(VII) ion (perchlorate ion) forms by loss of a hydrogen ion (when it reacts with water, for example), the charge can be delocalised over every oxygen atom in the ion. That makes it very stable, and means that chloric(VII) acid is very strong. Chloric(VII) acid reacts with sodium hydroxide solution to form a solution of sodium chlorate(VII).

Chlorine(VII) oxide itself also reacts with sodium hydroxide solution to give the same product.

Chlorine(I) oxide Chlorine(I) oxide is far less acidic than chlorine(VII) oxide. It reacts with water to some extent to give chloric(I) acid, HOCl - also known as hypochlorous acid.

Note: You may also find the chloric(I) acid written as HClO. The form I have used more accurately reflects the way the atoms are joined up.

The structure of chloric(I) acid is exactly as shown by its formula, HOCl. It has no doubly-bonded oxygens, and no way of delocalising the charge over the negative ion formed by loss of the hydrogen. That means that the negative ion formed isn't very stable, and readily reclaims its hydrogen to revert to the acid. Chloric(I) acid is very weak (pKa = 7.43). Chloric(I) acid reacts with sodium hydroxide solution to give a solution of sodium chlorate(I) (sodium hypochlorite).

Chlorine(I) oxide also reacts directly with sodium hydroxide to give the same product.

PROPERTIES OF THE PERIOD 3 CHLORIDES


This page looks at the structures of the chlorides of the Period 3 elements (sodium to sulphur), their physical properties and their reactions with water. Chlorine and argon are omitted - chlorine because it is meaningless to talk about "chlorine chloride", and argon because it doesn't form a chloride.

A quick summary of the trends


The chlorides The chlorides we'll be looking at are: NaCl MgCl2 AlCl3 SiCl4 PCl5 S2Cl2

PCl3

There are three chlorides of sulphur, but the only one mentioned by any of the UK-based syllabuses (A level or its equivalents) is S2Cl2. As you will see later, aluminium chloride exists in some circumstances as a dimer, Al2Cl6. The structures Sodium chloride and magnesium chloride are ionic and consist of giant ionic lattices at room temperature Aluminium chloride and phosphorus(V) chloride are tricky! They change their structure from ionic to covalent when the solid turns to a liquid or vapour. There is much more about this later on this page. The others are simple covalent molecules. Melting and boiling points Sodium and magnesium chlorides are solids with high melting and boiling points because of the large amount of heat which is needed to break the strong ionic attractions. The rest are liquids or low melting point solids. Leaving aside the aluminium chloride and phosphorus(V) chloride cases where the situation is quite complicated, the attractions in the others will be much weaker intermolecular forces such as van der Waals dispersion forces. These vary depending on the size and shape of the molecule, but will always be far weaker than ionic bonds.

Note: Follow this link if you aren't sure about intermolecular attractions such as van der Waals dispersion forces. Use the BACK button on your browser to return quickly to this page later.

Electrical conductivity Sodium and magnesium chlorides are ionic and so will undergo electrolysis when they are molten. Electricity is carried by the movement of the ions and their discharge at the electrodes. In the aluminium chloride and phosphorus(V) chloride cases, the solid doesn't conduct electricity because the ions aren't free to move. In the liquid (where it exists - both of these sublime at ordinary pressures), they have converted into a covalent form, and so don't conduct either. The rest of the chlorides don't conduct electricity either solid or molten because they don't have any ions or any

mobile electrons. Reactions with water As an approximation, the simple ionic chlorides (sodium and magnesium chloride) just dissolve in water. Tho other chlorides all react with water in a variety of ways described below for each individual chloride. The reaction with water is known as hydrolysis.

Warning: The rest of this page contains quite a lot of detail about the various chlorides, covering material from all the UK A level (or its equivalent) syllabuses. It is very unlikely that you will need all of this, and it is quite possible that your examiners will allow (or even expect) simplifications in some cases. It is essential to know what your examiners expect. You should obviously check your syllabus, but you also need to look at past papers and mark schemes so that you know what your examiners are actually asking. If you are working towards a UK-based exam and haven't got any of these things follow this link before you go any further to find out how to get them. It would also be useful to look at books written specifically for your actual syllabus. These will have been checked by your examiners, and they can hardly argue with anything you find in them. Look at the text book suggestions page to find some of the available books.

The individual chlorides


Sodium chloride, NaCl Sodium chloride is a simple ionic compound consisting of a giant array of sodium and chloride ions. A small representative bit of a sodium chloride lattice looks like this:

This is normally drawn in an exploded form as:

The strong attractions between the positive and negative ions need a lot of heat energy to break, and so sodium chloride has high melting and boiling points. It doesn't conduct electricity in the solid state because it hasn't any mobile electrons and the ions aren't free to move. However, when it melts it undergoes electrolysis. Sodium chloride simply dissolves in water to give a neutral solution.

Note: You will find the structure and physical properties of sodium chloride dealt with in a bit more detail (including an explanation of how to draw the last diagram) by following this link. Use the BACK button on your browser to return quickly to this page later.

Magnesium chloride, MgCl2

Magnesium chloride is also ionic, but with a more complicated arrangement of the ions to allow for having twice as many chloride ions as magnesium ions. This structure isn't needed for UK A level purposes. Again, lots of heat energy is needed to overcome the attractions between the ions, and so the melting and boiling points are again high.

Note: There is a problem here, though! You would expect the attractions between magnesium ions and chloride ions to be greater than those between sodium and chloride ions due to the extra charge on the magnesium. However, magnesium chloride melts at a lower temperature than sodium chloride, and the boiling points are almost identical (to within one degree). I have no idea what the explanation for this is, unless magnesium chloride isn't as ionic as we usually pretend. If anyone has any reliable information on this could they contact me via the address on the about this site page.

Solid magnesium chloride is a non-conductor of electricity because the ions aren't free to move. However, it undergoes electrolysis when the ions become free on melting. Magnesium chloride dissolves in water to give a faintly acidic solution (pH = approximately 6).

Note: This is one point when you need to know exactly what your examiners want you to say about this by looking at your syllabus, past papers and mark schemes. Some examiners simply say that magnesium chloride just dissolves in water. However, that wouldn't account for the slightly lowered pH. On the other hand, there is no point in learning a complicated bit of chemistry if all you need is a simplification. Be aware that it is a simplification, though.

When magnesium ions are broken off the solid lattice and go into solution, there is enough attraction between the 2+ ions and the water molecules to get co-ordinate (dative covalent) bonds formed between the magnesium ions and lone pairs on surrounding water molecules. Hexaaquamagnesium ions are formed, [Mg(H2O)6]2+.

Note: You will find the bonding in ions of this sort discussed with reference to the corresponding aluminium ion on the page about co-ordinate (dative covalent) bonding. The magnesium case is exactly the same.

Use the BACK button on your browser to return quickly to this page later.

Ions of this sort are acidic - the degree of acidity depending on how much the electrons in the water molecules are pulled towards the metal at the centre of the ion. The hydrogens are made rather more positive than they would otherwise be, and more easily pulled off by a base. In the magnesium case, the amount of distortion is quite small, and only a small proportion of the hydrogen atoms are removed by a base - in this case, by water molecules in the solution.

Note: The reason for the colour-coding is to try to avoid confusion between the water molecules attached to the ion and those in the solution.

The presence of the hydroxonium ions in the solution causes it to be acidic. The fact that there aren't many of them formed (the position of equilibrium lies well to the left), means that the solution is only weakly acidic. You may also find the last equation in a simplified form:

Hydrogen ions in solution are hydroxonium ions. If you use this form, it is essential to include the state symbols.

Note: You will find lots more about the acidity of hexaaqua ions by following this link. Use the BACK button on your browser to return quickly to this page later.

Aluminium chloride, AlCl3 Electronegativity increases as you go across the period and, by the time you get to aluminium, there isn't enough electronegativity difference between aluminium and chlorine for there to be a simple ionic bond.

Aluminium chloride is complicated by the way its structure changes as temperature increases. At room temperature, the aluminium in aluminium chloride is 6-coordinated. That means that each aluminium is surrounded by 6 chlorines. The structure is an ionic lattice - although with a lot of covalent character. At ordinary atmospheric pressure, aluminium chloride sublimes (turns straight from solid to vapour) at about 180C. If the pressure is raised to just over 2 atmospheres, it melts instead at a temperature of 192C. Both of these temperatures, of course, are completely wrong for an ionic compound - they are much too low. They suggest comparatively weak attractions between molecules - not strong attractions between ions. The coordination of the aluminium changes at these temperatures. It becomes 4-coordinated - each aluminium now being surrounded by 4 chlorines rather than 6. What happens is that the original lattice has converted into Al2Cl6 molecules. If you have read the page on coordinate bonding mentioned above, you will have seen that the structure of this is:

This conversion means, of course, that you have completely lost any ionic character - which is why the aluminium chloride vaporises or melts (depending on the pressure). There is an equilibrium between these dimers and simple AlCl3 molecules. As the temperature increases further, the position of equilibrium shifts more and more to the right.

Summary

At room temperature, solid aluminium chloride has an ionic lattice with a lot of covalent character. At temperatures around 180 - 190C (depending on the pressure), aluminium chloride coverts to a molecular form, Al2Cl6. This causes it to melt or vaporise because there are now only comparatively weak intermolecular attractions. As the temperature increases a bit more, it increasingly breaks up into simple AlCl3 molecules.

Solid aluminium chloride doesn't conduct electricity at room temperature because the ions aren't free to move. Molten aluminium chloride (only possible at increased pressures) doesn't conduct electricity because there aren't any ions any more.

Note: One very reliable source says that although solid aluminium chloride has zero conductivity at room

temperature, it conducts just below the melting point. I haven't at the moment been able to confirm this - neither do I have any idea why it might happen.

The reaction of aluminium chloride with water is dramatic. If you drop water onto solid aluminium chloride, you get a violent reaction producing clouds of steamy fumes of hydrogen chloride gas. If you add solid aluminium chloride to an excess of water, it still splutters, but instead of hydrogen chloride gas being given off, you get an acidic solution formed. A solution of aluminium chloride of ordinary concentrations (around 1 mol dm-3, for example) will have a pH around 2 - 3. More concentrated solutions will go lower than this. The aluminium chloride reacts with the water rather than just dissolving in it. In the first instance, hexaaquaaluminium ions are formed together with chloride ions.

You will see that this is very similar to the magnesium chloride equation given above - the only real difference is the charge on the ion. That extra charge pulls electrons from the water molecules quite strongly towards the aluminium. That makes the hydrogens more positive and so easier to remove from the ion. In other words, this ion is much more acidic than in the corresponding magnesium case. These equilibria (whichever you choose to write) lie further to the right, and so the solution formed is more acidic there are more hydroxonium ions in it.

or, more simply:

We haven't so far accounted for the burst of hydrogen chloride formed if there isn't much water present. All that happens is that because of the heat produced in the reaction and the concentration of the solution formed, hydrogen ions and chloride ions in the mixture combine together as hydrogen chloride molecules and are given off as a gas. With a large excess of water, the temperature never gets high enough for that to happen - the ions just stay in solution. Silicon tetrachloride, SiCl4 Silicon tetrachloride is a simple no-messing-about covalent chloride. There isn't enough electronegativity difference between the silicon and the chlorine for the two to form ionic bonds. Silicon tetrachloride is a colourless liquid at room temperature which fumes in moist air. The only attractions

between the molecules are van der Waals dispersion forces. It doesn't conduct electricity because of the lack of ions or mobile electrons. It fumes in moist air because it reacts with water in the air to produce hydrogen chloride. If you add water to silicon tetrachloride, there is a violent reaction to produce silicon dioxide and fumes of hydrogen chloride. In a large excess of water, the hydrogen chloride will, of course, dissolve to give a strongly acidic solution containing hydrochloric acid.

The phosphorus chlorides There are two phosphorus chlorides - phosphorus(III) chloride, PCl3, and phosphorus(V) chloride, PCl5. Phosphorus(III) chloride (phosphorus trichloride), PCl3 This is another simple covalent chloride - again a fuming liquid at room temperature. It is a liquid because there are only van der Waals dispersion forces and dipole-dipole attractions between the molecules.

Note: The phosphorus(III) chloride molecule has a permanent dipole, which is why dipole-dipole attractions are possible. There is a discussion about polar molecules and polar bonds on the page aboutelectronegativity. The phosphorus(III) chloride case is rather similar to CHCl 3 (discussed on that page), except that there is a lone pair of electrons at the top of the molecule rather than a hydrogen atom. Use the BACK button on your browser to return quickly to this page later.

It doesn't conduct electricity because of the lack of ions or mobile electrons. Phosphorus(III) chloride reacts violently with water. You get phosphorous acid, H3PO3, and fumes of hydrogen chloride (or a solution containing hydrochloric acid if lots of water is used).

Note: Phosphorous acid is also known as orthophosphorous acid or as phosphonic acid. Notice the "-ous" ending in the first two names. That's not a spelling mistake - it's for real! It is used to distinguish it from phosphoric acid which is quite different (see below).

Phosphorus(V) chloride (phosphorus pentachloride), PCl5 Unfortunately, phosphorus(V) chloride is structurally more complicated. Phosphorus(V) chloride is a white solid which sublimes at 163C. The higher the temperature goes above that, the more the phosphorus(V) chloride dissociates (splits up reversibly) to give phosphorus(III) chloride and chlorine.

Solid phosphorus(V) chloride contains ions - which is why it is a solid at room temperature. The formation of the ions involves two molecules of PCl5. A chloride ion transfers from one of the original molecules to the other, leaving a positive ion, [PCl 4]+, and a negative ion, [PCl6]-.

At 163C, the phosphorus(V) chloride converts to a simple molecular form containing PCl5 molecules. Because there are only van der Waals dispersion forces between these, it then vaporises. Solid phosphorus(V) chloride doesn't conduct electricity because the ions aren't free to move.

Note: Phosphorus(V) chloride does, however, undergo electrolysis in a suitable solvent which it doesn't react with. For example, Inorganic Chemistry by Heslop and Robinson quotes it as conducting electricity in solution in methyl nitrite, CH3ONO.

Phosphorus(V) chloride has a violent reaction with water producing fumes of hydrogen chloride. As with the other covalent chlorides, if there is enough water present, these will dissolve to give a solution containing hydrochloric acid. The reaction happens in two stages. In the first, with cold water, phosphorus oxychloride, POCl3, is produced along with HCl.

If the water is boiling, the phosphorus(V) chloride reacts further to give phosphoric(V) acid and more HCl. Phosphoric(V) acid is also known just as phosphoric acid or as orthophosphoric acid.

The overall equation in boiling water is just a combination of these:

Note: I'm not entirely happy about the conditions for these reactions. Several sources mention the need for boiling water for the second half of the reaction, but some don't quote any temperature. All of the safety data sheets available on the web talk about phosphorus oxychloride reacting strongly with water without any suggestion that it needs to be heated.

Disulphur dichloride, S2Cl2 Disulphur dichloride is just one of three sulphur chlorides, but is the only one mentioned by any of the UK A level syllabuses. This is possibly because it is the one which is formed when chlorine reacts with hot sulphur. Disulphur dichloride is a simple covalent liquid - orange and smelly! The shape is surprisingly difficult to draw convincingly! The atoms are all joined up in a line - but twisted:

The reason for drawing the shape is to give a hint about what sort of intermolecular attractions are possible. There is no plane of symmetry in the molecule and that means that it will have an overall permanent dipole. The liquid will have van der Waals dispersion forces and dipole-dipole attractions. There are no ions in disulphur dichloride and no mobile electrons - so it never conducts electricity. Disulphur dichloride reacts slowly with water to produce a complex mixture of things including hydrochloric acid, sulphur, hydrogen sulphide and various sulphur-containing acids and anions (negative ions). There is no way that you can write a single equation for this - and one would never be expected in an exam.

PROPERTIES OF THE PERIOD 3 "HYDROXIDES"


This page looks briefly at how the chemistry of the "hydroxides" of the Period 3 elements from sodium to chlorine varies as you cross the period. I am taking the word "hydroxide" to include anything which contains either a hydroxide ion or an -OH group covalently bound to the element in question. You wouldn't usually think of some of the compounds on this page as hydroxides at all.

A quick summary of the trends


Sodium and magnesium hydroxides These contain hydroxide ions, and are simple basic hydroxides. Aluminium hydroxide Aluminium hydroxide, like aluminium oxide, is amphoteric - it has both basic and acidic properties.

Note: I have no real idea of how to describe aluminium hydroxide on the ionic-covalent spectrum. Part of the problem is that there are several forms of aluminium hydroxide. For example, when it is prepared by adding 3+ ammonia solution to a solution containing hexaaquaaluminium ions, [Al(H2O)6] , it is probably first formed as the covalently bound Al(H2O)3(OH)3. But this will lose water and rearrange itself on standing - and I don't know what exactly is formed. The chemistry textbooks that I have to hand aren't too clear about the structure of aluminium hydroxide as far as the degree of covalent character is concerened, and a web search (until I got totally bored with it!) didn't throw up any reliable chemistry sites which discussed it. Several geology sites describe the mineral gibbsite (a naturally occurring form of aluminium hydroxide) in terms of ions, but whether it actually contains ions or whether this is just a simplification as a convenient way of talking about and drawing a complicated structure, I don't know.

The other "hydroxides" In all of these have -OH groups covalently bound to the atom from period 3. These compounds are all acidic ranging from the very weakly acidic silicic acids (one of which is shown below) to the very strong sulphuric or chloric(VII) acids.

There are other acids (also containing -OH groups) formed by these elements, but these are the ones where the Period 3 element is in its highest oxidation state.

Adding some detail


Sodium and magnesium hydroxides These are both basic because they contain hydroxide ions - a strong base. Both react with acids to form salts. For example, with dilute hydrochloric acid, you get colourless solutions of sodium chloride or magnesium chloride.

Aluminium hydroxide Aluminium hydroxide is amphoteric. Like sodium or magnesium hydroxides, it will react with acids. This is showing the basic side of its nature. With dilute hydrochloric acid, a colourless solution of aluminium chloride is formed.

But aluminium hydroxide also has an acidic side to its nature. It will react with sodium hydroxide solution to give a colourless solution of sodium tetrahydroxoaluminate.

Note: You may find all sorts of other formulae given for the product from this reaction. These range from NaAlO2 (which is a dehydrated form of the one in the equation) to Na 3Al(OH)6 (which is a different product altogether). What you actually get will depend on things like the temperature and the concentration of the sodium hydroxide solution. In any case, the truth is almost certainly a lot more complicated than any of these. (That's equally true of the previous equation involving the acid.) This is a case where it is a good idea to find out what your examiners quote in their support material or mark schemes, and stick with that. If necessary, get this sort of information from your Exam Board by following the links on the syllabuses page.

The other "hydroxides" A quick reminder of what we are talking about here:

None of these contains hydroxide ions. In each case the -OH group is covalently bound to the Period 3 element, and in each case it is possible for the hydrogens on these -OH groups to be removed by a base. In other words, all of these compounds are acidic. But they vary considerably in strength:

Orthosilicic acid is very weak indeed. Phosphoric(V) acid is a weak acid - although somewhat stronger than simple organic acids like ethanoic

acid. Sulphuric acid and chloric(VII) acids are both very strong acids.

The main factor in determining the strength of the acid is how stable the anion (the negative ion) is once the hydrogen has been removed. This in turn depends on how much the negative charge can be spread around the rest of the ion. If the negative charge stays entirely on the oxygen atom left behind from the -OH group, it will be very attractive to hydrogen ions. The lost hydrogen ion will be easily recaptured and the acid will be weak. On the other hand, if the charge can be spread out (delocalised) over the whole of the ion, it will be so "dilute" that it won't attract the hydrogen back very easily. The acid will then be strong. Wherever possible, the negative charge is delocalised by interacting with doubly-bonded oxygens. For example, in chloric(VII) acid, the ion produced is the chlorate(VII) ion (also known as the perchlorate ion), ClO4-. The structure of the ion doesn't stay like this:

Instead, the negative charge is delocalised over the whole ion, and all four chlorine-oxygen bonds are identical.

Note: This is just like the delocalisation which occurs in the ethanoate ion formed when ethanoic acid is behaving as a weak acid (except on a larger scale). You will find this described in some detail on a page aboutorganic acids. Use the BACK button on your browser if you choose to follow this link.

When sulphuric acid loses a hydrogen ion to form the hydrogensulphate ion, HSO4-, the charge can be spread

over three oxygens (the original one with the negative charge, and the two sulphur-oxygen double bonds. That's still an effective delocalisation, and sulphuric acid is almost as strong as chloric(VII) acid.

Note: Sulphuric acid can, of course, lose a second hydrogen ion as well from the other -OH group and form sulphate ions. However, that is a bit more difficult. If you lose that second hydrogen, you can use all four oxygens to delocalise the charge - but now you have to delocalise two negative charges rather than just one. The hydrogensulphate ion isn't a strong acid. It's strength is similar to phosphoric(V) acid.

Phosphoric(V) acid is much weaker than sulphuric acid because it only has one phosphorus-oxygen double bond which it can use to help delocalise the charge on the ion formed by losing one hydrogen ion - so the charge on that ion is delocalised less effectively. In orthosilicic acid, there aren't any silicon-oxygen double bonds to delocalise the charge. That means the ion formed by loss of a hydrogen ion isn't at all stable, and easily recovers its hydrogen.

Note: If you want some reactions of these acids with bases, you will find them on the page about the reactions of the Period 3 oxides. You will also find information there about some other phosphorus, sulphur and chlorinecontaining acids, all of which are formed when the relevant oxides react with water.

PERIODIC TABLE GROUP 1 MENU

Atomic and physical properties . . . Discusses trends in atomic radius, ionisation energy, electronegativity, melting and boiling points, and densities of the Group 1 elements. Reactions with water . . . Looks at the trends in the reactions between the Group 1 elements and water. Reactions with oxygen and chlorine. . . Looks at the reactions of the Group 1 elements with oxygen, including the formation of peroxides and superoxides. The reactions of the various oxides with water and acids. Also a brief look at the reactions between the metals and chlorine. Some Group 1 compounds . . . Some properties and reactions of the nitrates, carbonates, hydrogencarbonates and hydrides of the Group 1 elements - limited to what is required by various UK A level syllabuses.

Flame tests . . . A brief introduction to flame tests for Group 1 (and other) metal ions.

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

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

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 the corresponding 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 the corresponding 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!

Note: If your maths is reasonable, it isn't too difficult to work out what the density of sodium should be if you know lithium's density and have figures for relative atomic mass and atomic radius - and similarly all the rest of the way down the Group. I'm not going to waste time and space doing it, because I can't conceive of any situation when you would ever be asked to do it! If you want to try it, you will find that the density is inversely proportional to the volumes of the atoms and directly proportional to their masses.

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

-222

Na

-184

-196

Rb

-195

Cs

-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

1st IE

hydr. enthalpy

total

Li

+161

+519

-519

+161

Na

+109

+494

-406

+197

+90

+418

-322

+186

Rb

+86

+402

-301

+187

Cs

+79

+376

-276

+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:

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

1st IE

total

Li

+161

+519

+680

Na

+109

+494

+603

+90

+418

+508

Rb

+86

+402

+488

Cs

+79

+376

+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.

Note: If you are a UK A level student, you will almost certainly find that your examiners will only expect you to explain this in terms of the fall in ionisation energy as you go down the Group. In other words, they simplify things by overlooking the contribution from atomisation energy. Stick with what your examiners expect - don't make life difficult for yourself! I'm trying to be as rigorous as I can because a sizeable part of my audience is working in systems outside the UK or beyond A level.

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?

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

-236

RbO2

-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!

SOME 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 are delocalised. This is a rather more complicated version of the bonding you might have come across in benzene or in ions like ethanoate. For the purposes of this topic, you don't need to understand how this bonding has come about.

Note: If you are interested, you could follow these links to benzene or to organic acids. Either of these links is

likely to involve you in a fairly time-consuming detour!

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.

FLAME TESTS
This page describes how to do a flame test for a range of metal ions, and briefly describes how the flame colour arises. Flame tests are used to identify the presence of a relatively small number of metal ions in a compound. Not all metal ions give flame colours. For Group 1 compounds, flame tests are usually by far the easiest way of identifying which metal you have got. For other metals, there are usually other easy methods which are more reliable - but the flame test can give a useful hint as to where to look.

Carrying out a flame test


Practical details Clean a platinum or nichrome (a nickel-chromium alloy) wire by dipping it into concentrated hydrochloric acid and then holding it in a hot Bunsen flame. Repeat this until the wire doesn't produce any colour in the flame.

Note: There will, in fact, always be a trace of orange in the flame if you use nichrome. You soon learn to ignore this. Platinum is much better to use, but is much, much more expensive. If you have a particularly dirty bit of nichrome wire, you can just chop the end off. You don't do that with platinum! Dilute hydrochloric acid can be used instead of concentrated acid for safety reasons, but doesn't always give such intense flame colours.

When the wire is clean, moisten it again with some of the acid and then dip it into a small amount of the solid you are testing so that some sticks to the wire. Place the wire back in the flame again. If the flame colour is weak, it is often worthwhile to dip the wire back in the acid again and put it back into the flame as if you were cleaning it. You often get a very short but intense flash of colour by doing that. 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

red

Na

strong persistent orange

lilac (pink)

Rb

red (reddish-violet)

Cs

blue? violet? (see below)

Ca

orange-red

Sr

red

Ba

pale green

Cu

blue-green (often with white flashes)

Pb

greyish-white

What do you do if you have a red flame colour for an unknown compound and don't know which of the various reds it is? Get samples of known lithium, strontium (etc) compounds and repeat the flame test, comparing the colours produced by one of the known compounds and the unknown compound side by side until you have a good match.

Note: I don't have confidence in the caesium flame colour. There is total disagreement about this on the web and

in the books I have looked at, and I have never seen this flame colour myself. However, I have received a helpful email from a student who says: "At my school we did some flame testing experiments, and . . . caesium is actually either blue or violet, depending on the way you look at it. I think it looks more violet than blue, but it sort of changes each time you do it." (Kara Gates, March 2006). If you thought chemistry was clear-cut, you are sadly mistaken!

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. However, it might jump back in two (or more) stages. For example, first to the 5 level and then back to the 2 level. Each of these jumps involves a specific amount of energy being released as light energy, and each corresponds to a particular colour. As a result of all these jumps, a spectrum of coloured lines will be produced. The colour you see will be a combination of all these individual colours. The exact sizes of the possible jumps in energy terms vary from one metal ion to another. That means that each different ion will have a different pattern of spectral lines, and so a different flame colour.

PERIODIC TABLE GROUP 2 MENU

Atomic and physical properties . . . Discusses trends in atomic radius, ionisation energy, electronegativity and melting point of the Group 2 elements. Reactions with water . . .

Looks at the trends in the reactions between the Group 2 elements and water. Reactions with oxygen . . . Looks at the trends in the reactions between the Group 2 elements and oxygen. The solubility of the hydroxides, sulphates and carbonates. . . Describes the patterns in the solubilities of the hydroxides, sulphates and carbonates of the Group 2 elements. The thermal stability of the nitrates and carbonates . . . Describes and explains patterns in the effect of heat on the nitrates and carbonates of the Group 2 elements. Some atypical properties of beryllium compounds . . . Describes and explains some cases where beryllium compounds don't fit the patterns shown by the rest of the Group.

ATOMIC AND PHYSICAL PROPERTIES OF THE GROUP 2 ELEMENTS


This page explores the trends in some atomic and physical properties of the Group 2 elements - beryllium, magnesium, calcium, strontium and barium. You will find separate sections below covering the trends in atomic radius, first ionisation energy, electronegativity and 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

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

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 not exactly 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 face-centred cubic; barium is bodycentred 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-coordinated. 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 magnesiumcalcium 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 book Inorganic 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. REACTIONS OF THE GROUP 2 ELEMENTS WITH WATER This page looks at the reactions of the Group 2 elements - beryllium, magnesium, calcium, strontium and barium - with water (or steam). It uses these reactions to explore the trend in reactivity in Group 2. The Facts Beryllium Beryllium has no reaction with water or steam even at red heat. Magnesium Magnesium burns in steam to produce 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:

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 can do this with the reactions with water (or steam), and you might like to follow this link if you haven't already been there. Trying to account for a non-existent trend in the reactions with oxygen is just silly!

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

Beryllium: I can't find a reference anywhere (text books or internet) to the colour of the flame that beryllium burns with. My best guess would be the same sort of silvery sparkles that magnesium or aluminium powder burn with if they are scattered into a flame - but I don't know that for sure. Magnesium, of course, burns with a typical intense white flame. Calcium is quite reluctant to start burning, but then bursts dramatically into flame, burning with an intense white flame with a tinge of red at the end. Strontium: I have 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 a very reactive metal like potassium in air. The potassium melts at a low temperature and almost instantly turns into a pool of molten potassium oxide. The activation energy is so low that the reaction happens very quickly at quite a low temperature. There is often no trace of flame. It can be fairly boring! Magnesium, on the other hand, has to be heated to quite a high temperature before it will start to react. The activation energy is much higher. There are also problems with surface coatings. It is then so hot that it produces the typical intense white flame. It would obviously be totally misleading to say that magnesium is more reactive than potassium on the evidence of the bright flame. You haven't had to heat them by the same amount to get the reactions happening.

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 N3- ions. All of these processes absorb energy. This energy has to be recovered from somewhere to give an overall exothermic reaction - if the energy can't be recovered, the overall change will be endothermic and won't happen.

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. SOLUBILITY OF THE HYDROXIDES, SULPHATES AND CARBONATES OF THE GROUP 2 ELEMENTS IN WATER This page looks at the solubility in water of the hydroxides, sulphates and carbonates of the Group 2 elements beryllium, magnesium, calcium, strontium and barium. Although it describes the trends, there isn't any attempt to explain them on this page - for reasons discussed later. You will find that there aren't any figures given for any of the solubilities. There are major discrepancies between the figures given by two common UK A level Data Books (Nuffield Advanced Science Book of Data, andChemistry Data Book by Stark and Wallace). There are also important inconsistencies within the books (one set of figures doesn't agree with those which can be calculated from another set). I haven't been able to find data which I am sure is correct, and therefore prefer not to give any. The Facts Solubility of the hydroxides

The hydroxides become more soluble as you go down the Group.

This is a trend which holds for the whole Group, and applies whichever set of data you choose. Some examples may help you to remember the trend: Magnesium hydroxide appears to be insoluble in water. However, if you shake it with water, filter it and test the pH

of the solution, you find that it is slightly alkaline. This shows that there are more hydroxide ions in the solution than there were in the original water. Some magnesium hydroxide must have dissolved. Calcium hydroxide solution is used as "lime water". 1 litre of pure water 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 dm3 at room temperature. Solubility of the sulphates

The sulphates become less soluble as you go down the Group.

The simple trend is true provided you include hydrated beryllium sulphate in it, but not if the beryllium sulphate is anhydrous. The Nuffield Data Book quotes anyhydrous beryllium sulphate, BeSO4, as insoluble (I haven't been able to confirm this from any other source), whereas the hydrated form, BeSO4.4H2O is soluble. (The Data Books agree on this - giving a figure of about 39 g dissolving in 100 g of water at room temperature.) Figures for magnesium sulphate and calcium sulphate also vary depending on whether the salt is hydrated or not, but nothing like so dramatically. Two common examples may help you to remember the trend: You are probably familiar with the reaction between magnesium and dilute sulphuric acid to give lots of hydrogen and a colourless solution of magnesium sulphate. Notice that you get a solution, not a precipitate. The magnesium sulphate is obviously soluble. You may also remember that barium sulphate is formed as a white precipitate during the test for sulphate ions in solution. The ready formation of a precipitate shows that the barium sulphate must be pretty insoluble. In fact, 1 litre of water will only dissolve about 2 mg of barium sulphate at room temperature. Solubility of the carbonates

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

None of the carbonates is anything more than very sparingly soluble. Magnesium carbonate (the most soluble one I have data for) is soluble to the extent of about 0.02 g per 100 g of water at room temperature. I can't find any data for beryllium carbonate, but it tends to react with water and so that might confuse the trend. The trend to lower solubility is, however, broken at the bottom of the Group. Barium carbonate is slightly more soluble than strontium sulphate. There are no simple examples which might help you to remember the carbonate trend. What - no explanations? Before I started to write this page, I thought I understood the trends in solubility patterns including the

explanations for them. The more I have dug around to try to find reliable data, and the more time I have spent thinking about it, the less I'm sure that it is possible to come up with any simple explanation of the solubility patterns.

Note: If you are interested in the reasons why I am unwilling to give the usual over-simplified explanations, I have described some of the problems as I see them on a separate page. That page also includes an attempt at a better explanation. Unless your syllabus specifically asks for explanations of these trends, you would be better off ignoring this follow-up page!

EXPLANATIONS FOR THE TRENDS IN SOLUBILITY OF SOME GROUP 2 COMPOUNDS


This page looks at the usual explanations for the solubility patterns in the hydroxides, sulphates and carbonates of Group 2. It goes on to look at my misgivings about these. Don't expect this page to be easy - it is probably best avoided unless your syllabus specifically asks for these explanations!

Warning: If you have come straight to this page via a search engine, you should be aware that it is a follow-up to another page describing the solubility patterns. You should read that page before going on with this one.

The usual explanations


Enthalpy changes during the process The usual explanation is in terms of the enthalpy changes which occur when an ionic compound dissolves in water. Energy has to be supplied to break up the lattice of ions, and energy is released when these ions form bonds of one sort or another with water molecules.

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

As you go down a Group, the energy needed to break up the lattice falls as the positive ions get bigger. The bigger the ions, the more distance there is between them, and the weaker the forces holding them together. Again as the positive ions get bigger, the energy released as the ions bond to water molecules (their hydration enthalpies) falls as well. Bigger ions aren't so strongly attracted to the water molecules. Since both of these important enthalpy terms fall as you go down the Group, what matters in deciding whether the change becomes more endothermic or more exothermic overall is how fast they fall relative to each other. A sample calculation to make this clear Here is an example of the sort of calculations you might do to work out the enthalpy change of solution for sodium chloride and potassium chloride.

Note: You may wonder why I haven't used an example directly relevant to the hydroxides, sulphates or carbonates of Group 2. I tried doing this with figures (taken from a variety of sources) for carbonates, but got answers for enthalpy changes of solution which didn't bear the slightest resemblance to the actual values quoted in a data book. I have no idea why! Doing the sums with Group 1 chlorides does give (fairly) accurate results, and the methods and discussion applies equally well to the Group 2 compounds we are interested in. The data used comes from Chemistry Data Book by Stark and Wallace. The Nuffield Data Book doesn't have any hydration enthalpy values.

In this case, we are defining lattice enthalpy as the heat needed to convert 1 mole of crystal in its standard state into separate gaseous ions - an endothermic change.

Note: Lattice enthalpy is more usually defined in the opposite direction - as the heat evolved when the lattice is formed from its gaseous ions. Defining it in the endothermic direction makes the argument a bit easier in this particular case. This endothermic change involving the break up of the lattice is properly called "lattice dissociation enthalpy", although the simpler term "lattice enthalpy" is still often used for it.

You can see that the enthalpy of solution changes from NaCl to KCl because the lattice enthalpy and hydration enthalpy of the positive ion fall by different amounts.

If the hydration enthalpy falls faster than the lattice enthalpy (as in this case), the net effect is that the overall change becomes more endothermic (or less exothermic in other possible cases where the total enthalpy change turns out to be negative). If the lattice enthalpy falls faster than the hydration enthalpy, the opposite happens - the change will become less endothermic (or more exothermic).

What controls the relative rate of fall of the two terms? It turns out that the main factor is the size of the negative ion.

For small negative ions like hydroxide, the lattice enthalpy falls faster than the hydration enthalpy of the positive ions. That means that the enthalpy of solution will become less positive (or more negative). For large negative ions like sulphate or carbonate, the hydration enthalpy of the positive ions falls faster than the lattice enthalpy. In this case, the enthalpy of solution will become more positive (or less negative).

Why the difference? Lattice enthalpy is governed by several factors - including the distance between the negative and positive ions. Where you have a big negative ion, this inter-ionic distance is largely controlled by the size of that negative ion. Changes in the size of the positive ion don't make as great a percentage difference to the interionic distance as they would if the negative ion was small. With sulphates, for example, the percentage increase in the inter-ionic distance as you go from magnesium to calcium sulphate isn't as great as it would be with a smaller negative ion like hydroxide. Since the percentage increase in inter-ionic distance isn't very great, the change in the lattice enthalpy won't be very great either. The relationship between enthalpy of solution and solubility The assumption is made that the more endothermic (or less exothermic) the enthalpy of solution is, the less soluble the compound. So sulphates and carbonates become less soluble as you go down the Group; hydroxides become more soluble.

Problems with the usual explanations

Problems with the data Unfortunately, the enthalpy of solution values for the Group 1 chlorides as calculated above don't agree with the values given in the same Data Book:

Calculated value (kJ mol-1)

Data book value (kJ mol-1)

NaCl

+1

+3.9

KCl

+15

+17.2

RbCl

+10

+16.7

CsCl

+5

+17.9

The discrepancies are enough to disrupt any pattern (such as there is!). The reasons for the discrepancies lie in the way the numbers are calculated. These are very small numbers worked out from much larger ones. Small uncertainties in those large numbers will cause large swings in the answers. For example, if each of the numbers in the calculations we did earlier on this page was out by just 5 kJ, each answer could vary by +/- 15 kJ - completely disrupting the patterns! You can't therefore reliably use the data available to calculate the trends you want with sufficient accuracy to make sense. Problems correlating enthalpy data with the facts The solubilities of the Group 1 chlorides (in moles of solute saturating 100 g of water at 298 K) compared with their enthalpies of solution are:

Enthalpy of solution (kJ mol-1)

Solubility (mol /100 g of water)

NaCl

+3.9

0.615

KCl

+17.2

0.481

RbCl

+16.7

0.781

CsCl

+17.9

1.13

There is no obvious relationship connecting the relative movements of these solubility values with the enthalpy of solution figures. It would be quite untrue to say that the more endothermic the change, the less soluble the compound! Problems in relating the sign of the enthalpy change to solubility The table above illustrates this problem, but it gets worse! Taking the sign of enthalpy of solution at face value, you get some bizarre results. The enthalpy of solution figures for the Group 2 carbonates are: (source: Chemistry Data Book by Stark and Wallace; values in kJ mol-1)

MgCO3

-25.3

CaCO3

-12.3

SrCO3

-3.4

BaCO3

+4.2

Sodium chloride and the other Group 1 chlorides dissolve despite the fact that their enthalpies of solution are positive, and yet magnesium carbonate (and most of the other Group 2 carbonates) are very sparingly soluble, but have exothermic enthalpies of solution. You might have expected exactly the opposite to happen. However if you ignore the comparison with the Group 1 chlorides, you could argue that the figures get progressively less exothermic, and at barium carbonate become endothermic. That would seem to support the decrease in solubility as you go down the Group quite nicely. Unfortunately, if you look at the solubility data, the trend is broken at the bottom of the Group. Barium carbonate is more soluble than strontium carbonate! Although figures from my two data sources differ in detail, they agree on this.

Clearly, trying to correlate solubility simply with the enthalpy change of solution doesn't work.

Introducing entropy changes


The basic explanation To get around the problem of many compounds dissolving freely in water despite the fact that their enthalpies of solution are endothermic you have to introduce the concept of entropy change. The only way of making sense of entropy without getting bogged down in some serious maths is to think of it as a measure of the amount of disorder in a system. Entropy is given the symbol S. If a system becomes more disordered, then its entropy increases. In order to see whether a change is possible or not, you have to think about a combination of the enthalpy change and the entropy change. These can be combined mathematically to give an important term known asfree energy change.

As an approximation, for a reaction to happen, the free energy change must be negative. What happens if the enthalpy change is positive - as for example when sodium chloride dissolves in water (+3.9 kJ mol-1, using the values in one of the tables above)? As long as the entropy change is positive enough, it is possible to get a negative value for free energy change. In the sodium chloride case, you don't have to have very much increase in entropy to outweigh the small enthalpy change of +3.9 kJ mol-1.

So . . . does the entropy increase when sodium chloride dissolve in water? Yes, it does! Originally, the sodium and chloride ions were arranged in a very tidy way in the crystal lattice - their entropy was low. (Remember that entropy is a measure of disorder.) When you dissolve the crystal in water, the entropy increases as the ions and water molecules become completely jumbled up - they become much more disordered than they were originally.

This is where the explanation usually stops, but to stop at this point is very misleading because it won't explain all the facts! What's wrong with this explanation? The most obvious thing that's wrong is that it won't explain why some compounds (like magnesium carbonate, and most of the other Group 2 carbonates) don't dissolve in water even though their enthalpies of solution are mainly negative. In these cases, the entropy of the system must fall when the compounds dissolve in water - in other words, the solution in water is more ordered than the original crystal and water!

This happens because the water molecules become more ordered when the compound dissolves in them. Instead of milling around pretty much at random, they become attracted to the ions present and arranged around them. This is particularly effective if the ions are small and highly charged - and so the effect is greatest for the positive ions at the top of the Group, and gets less as you go down. It is also much more important in Group 2 than in Group 1 where the ions only carry one positive charge. That means that you have two entropy effects to consider. There is the increase in disorder as the crystal lattice breaks up, but a corresponding increase in order in the water - which varies depending on the sizes and charges of the ions present. What a nightmare! To explain this properly, you need to think about the way lattice enthalpy changes as you go down the Group, the way that hydration enthalpies change, and the way that entropy changes. The way those changes happen will vary from one type of compound to another. Can we explain everything now? No - at least not easily! For example, although it might be possible to account for the lack of pattern in the solubilities of the Group 1 chlorides (and also the bromides) by a mathematical application of these effects, trying to do it in general terms defeats me completely! You could, however, make a reasonable suggestion as to why the solubility trend in the carbonates is broken at barium. (Don't expect the explanation to be instantly understandable though!) Remember that the solubility of the carbonates falls as you go down Group 2, apart from an increase as you go

from strontium to barium carbonate. The general fall is because hydration enthalpies are falling faster than lattice enthalpies. Remember that where you have a big negative ion, its size dominates the inter-ionic distance and so doesn't allow the lattice enthalpy to change much. This gives the enthalpy of solution values we've already looked at (values in kJ mol-1):

MgCO3

-25.3

CaCO3

-12.3

SrCO3

-3.4

BaCO3

+4.2

But the entropy change will also be varying as you go down the Group. At the top, where you have small 2+ ions, the overall entropy change in the system must be negative - the system as a whole becomes more ordered when the compound dissolves because of the way the water molecules become organised around the positive ions. That negative entropy change is going to be enough to wipe out the effect of the exothermic enthalpy of solution.

Note: You need to think about the effect of the positiveness and negativeness of these values on the size of the free energy change. Use the equation further up this page.

Towards the bottom of the Group, this effect changes. The bigger ions have less organising effect on the water molecules. The entropy change is becoming less negative (or perhaps even at this stage, positive). That's going to tend to make the compounds more soluble. The overall effect is a complex balance between the way the enthalpy of solution varies and the way the entropy change of solution alters. At barium carbonate, the effect of increasing entropy must be enough to make it more soluble than strontium carbonate.

A personal comment: I see it as quite wrong that this should be discussed at this level (UK A level or its equivalent). The usual explanations are over-simplistic and potentially misleading - especially for anyone who might go on to do Chemistry at degree level. Fortunately most syllabuses have enough sense not to ask for these explanations! Science progresses by offering theories which have to explain all the facts. Where a fact won't fit a theory, the theory has to be modified, or even discarded. What we seem to be doing here is presenting students with an

inadequate theory and then ignoring all the facts which don't fit it. I see that as quite dangerous. It would be much better not to discuss this at all at this level, rather than to give students a false view of the way science works.

SOLUBILITY OF THE HYDROXIDES, SULPHATES AND CARBONATES OF THE GROUP 2 ELEMENTS IN WATER
This page looks at the solubility in water of the hydroxides, sulphates and carbonates of the Group 2 elements beryllium, magnesium, calcium, strontium and barium. Although it describes the trends, there isn't any attempt to explain them on this page - for reasons discussed later. You will find that there aren't any figures given for any of the solubilities. There are major discrepancies between the figures given by two common UK A level Data Books (Nuffield Advanced Science Book of Data, andChemistry Data Book by Stark and Wallace). There are also important inconsistencies within the books (one set of figures doesn't agree with those which can be calculated from another set). I haven't been able to find data which I am sure is correct, and therefore prefer not to give any.

The Facts
Solubility of the hydroxides

The hydroxides become more soluble as you go down the Group.

This is a trend which holds for the whole Group, and applies whichever set of data you choose. Some examples may help you to remember the trend: Magnesium hydroxide appears to be insoluble in water. However, if you shake it with water, filter it and test the pH of the solution, you find that it is slightly alkaline. This shows that there are more hydroxide ions in the solution than there were in the original water. Some magnesium hydroxide must have dissolved. Calcium hydroxide solution is used as "lime water". 1 litre of pure water 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 dm3 at room temperature. Solubility of the sulphates

The sulphates become less soluble as you go down the Group.

The simple trend is true provided you include hydrated beryllium sulphate in it, but not if the beryllium sulphate is

anhydrous. The Nuffield Data Book quotes anyhydrous beryllium sulphate, BeSO4, as insoluble (I haven't been able to confirm this from any other source), whereas the hydrated form, BeSO4.4H2O is soluble. (The Data Books agree on this - giving a figure of about 39 g dissolving in 100 g of water at room temperature.) Figures for magnesium sulphate and calcium sulphate also vary depending on whether the salt is hydrated or not, but nothing like so dramatically. Two common examples may help you to remember the trend: You are probably familiar with the reaction between magnesium and dilute sulphuric acid to give lots of hydrogen and a colourless solution of magnesium sulphate. Notice that you get a solution, not a precipitate. The magnesium sulphate is obviously soluble. You may also remember that barium sulphate is formed as a white precipitate during the test for sulphate ions in solution. The ready formation of a precipitate shows that the barium sulphate must be pretty insoluble. In fact, 1 litre of water will only dissolve about 2 mg of barium sulphate at room temperature. Solubility of the carbonates

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

None of the carbonates is anything more than very sparingly soluble. Magnesium carbonate (the most soluble one I have data for) is soluble to the extent of about 0.02 g per 100 g of water at room temperature. I can't find any data for beryllium carbonate, but it tends to react with water and so that might confuse the trend. The trend to lower solubility is, however, broken at the bottom of the Group. Barium carbonate is slightly more soluble than strontium sulphate. There are no simple examples which might help you to remember the carbonate trend.

What - no explanations?
Before I started to write this page, I thought I understood the trends in solubility patterns including the explanations for them. The more I have dug around to try to find reliable data, and the more time I have spent thinking about it, the less I'm sure that it is possible to come up with any simple explanation of the solubility patterns.

Note: If you are interested in the reasons why I am unwilling to give the usual over-simplified explanations, I have described some of the problems as I see them on a separate page. That page also includes an attempt at a better explanation. Unless your syllabus specifically asks for explanations of these trends, you would be better off ignoring this follow-up page!

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.

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.

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 yoursyllabus 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 are delocalised. This is a rather more complicated version of the bonding you might have come across in benzene or in ions like ethanoate. For the purposes of this topic, you don't need to understand how this bonding has come about.

Note: If you are interested, you could follow these links to benzene or to organic acids. Either of these links is likely to involve you in a fairly time-consuming detour!

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

+117

CaCO3

+178

SrCO3

+235

BaCO3

+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 -1 its gaseous ions. In that case, the lattice enthalpy for magnesium oxide would be -3889 kJ mol . 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+

0.065

Ca2+

0.099

O2-

0.140

CO32-

The lattice enthalpies fall at different rates because of the different sizes of the two negative ions - oxide and carbonate. The oxide ion is relatively small for a negative ion (0.140 nm), whereas the carbonate ion is large (no figure available). In the oxides, when you go from magnesium oxide to calcium oxide, for example, the inter-ionic distance increases from 0.205 nm (0.140 + 0.065) to 0.239 nm (0.140 + 0.099) - an increase of about 17%. In the carbonates, the inter-ionic distance is dominated by the much larger carbonate ion. Although the inter-ionic distance will increase by the same amount as you go from magnesium carbonate to calcium carbonate, as a percentage of the total distance the increase will be much less. Some made-up figures show this clearly. I can't find a value for the radius of a carbonate ion, and so can't use real figures. For the sake of argument, suppose that the carbonate ion radius was 0.3 nm. The inter-ionic distances in the two cases we are talking about would increase from 0.365 nm to 0.399 nm - an increase of only about 9%. 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.

SOME BERYLLIUM CHEMISTRY UNTYPICAL OF GROUP 2


This page describes and explains three examples from beryllium chemistry where it behaves differently from the rest of Group 2.

Beryllium chloride is covalent


The facts

Physical properties Beryllium chloride, BeCl2, melts at 405C and boils at 520C. That compares with 714C and 1412C for magnesium chloride. Notice how much dramatically lower the boiling point of beryllium chloride is compared with magnesium chloride. The much higher boiling point of magnesium chloride is what you might expect from the strong forces between the positive and negative ions present. Because its boiling point is much lower, it follows that beryllium chloride can't contain ions - it must be covalent. On the other hand, the melting point is quite high for such a small covalent molecule. There must be something more complicated going on! Reaction with water Beryllium chloride reacts vigorously and exothermically with water with the evolution of acidic, steamy hydrogen chloride gas. This is typical of covalent chlorides. In the first instance, it reacts to give hydrated beryllium ions, [Be(H2O)4]2+, and chloride ions. But the hydrated beryllium ions (called tetraaquaberyllium ions) are quite strongly acidic. The small beryllium ion at the centre attracts the electrons in the bonds towards itself, and that makes the hydrogen atoms in the water even more positive than they usually are. If the solution is hot and concentrated (as it is likely to be if you add water to solid beryllium chloride - a very exothermic reaction), chloride ions can remove one or more of these hydrogen ions to produce hydrogen chloride gas. All the other ionic chlorides in Group 2 dissolve in water without any obvious reaction.

Note: There is actually a very small amount of reaction between anhydrous magnesium chloride and water, although you wouldn't notice it. Magnesium chloride isn't quite as purely ionic as we sometimes pretend! Follow this link if you are interested in exploring the naming of complex ions. . . . or this one for detailed explanations of why complex ions similar to the beryllium one are acidic.

The structure of beryllium chloride As a gas . . . Beryllium chloride, BeCl2, is a linear molecule with all three atoms in a straight line. Showing only the outer electrons:

Beryllium chloride is known as an electron-deficient compound because it has the two empty orbitals at the bonding level. As a solid . . . If it had this same simple structure as a solid, you would expect the melting point to be much lower than it actually is. It is a very small molecule, and so the intermolecular attractions would be expected to be fairly weak.

Note: If you aren't sure about intermolecular forces (Van der Waals forces) you could follow this link to find out more about them. Use the BACK button on your browser to return to this page.

In the solid, the BeCl2 molecules polymerise to make long chains. They do this by forming coordinate bonds (dative covalent bonds) between lone pairs on chlorine atoms and adjacent beryllium atoms. The diagram shows a simple dimer - the start of the polymerisation process.

Note: This is exactly the same as the coordinate bonding in aluminium chloride. If you aren't happy about coordinate (dative covalent) bonding, it would pay you to look at this page before you go on. Use the BACK button on your browser to return to this page.

You can see from the last diagram that the beryllium atoms are still electron deficient. The process continues. The next diagram shows the coordinate bonds in the conventional way using arrows. The arrow goes from the atom which is supplying the pair of electrons to the atom with the empty orbital.

Make sure that you fully understand how this diagram relates to the dimer shown in the previous diagram. Why isn't beryllium chloride ionic? Beryllium has quite a high electronegativity compared with the rest of the Group. That means that it attracts a bonding pair of electrons towards itself more strongly than magnesium and the rest do. In order for an ionic bond to form, the beryllium has to let go of its electrons. It is too electronegative to do that.

Note: The trends in electronegativity in Group 2 are discussed on another page. That page looks at the way the electons are arranged in the beryllium-chlorine bond compared with the magnesium-chlorine bond. Use the BACK button on your browser to return to this page - or come back via the Group 2 menu.

Beryllium forms 4-coordinated complex ions


Some simple background Although beryllium doesn't normally form simple ions, Be2+, it does form ions in solution. In these, the beryllium ion becomes attached to four water molecules to give a complex ion with the formula [Be(H2O)4]2+. The ion is said to be 4-coordinated, or to have a coordination number of 4, because there are four water molecules arranged around the central beryllium. Many hydrated metal ions are 6-coordinated. For example, magnesium ions in solution exist as [Mg(H2O)6]2+. The water molecules in these ions are attached to the central metal ion via coordinate bonds (dative covalent bonds). One of the lone pairs on each water molecule is used to form a bond with an empty orbital in the metal ion. Each time one of these bonds is formed, energy is released, and the ion becomes more stable. It would seem logical for the metal ion to form as many bonds like this as it possibly can.

Note: If you aren't happy about coordinate bonding you must follow this link before you go on. You will find the bonding in hydrated metal ions discussed in some detail on that page. Use the BACK button on your browser to return to this page.

Why does beryllium only attach four water molecules? The hydration of beryllium The problem is that there has to be somewhere that the lone pairs on the water molecules can attach to. Beryllium has the electronic structure 1s22s2. It is helpful to draw this as an "electrons-in-boxes" diagram:

Note: If you aren't happy about orbitals you really ought to follow this link before you go on. You may want to explore further in that part of the site as well. Unless you understand exactly what this electrons-in-boxes diagram is about, you won't be able to make sense of what is coming up next.

When beryllium forms a 2+ ion it loses the 2 electrons in the 2s orbital. That leaves the 2-level completely empty. The 2-level orbitals reorganise themselves (hybridise) to make four equal orbitals, each of which can accept a lone pair of electrons from a water molecule. In the next diagram the 1s electrons have been left out. They aren't relevant to the bonding.

Each water molecule, of course, has two lone pairs of electrons. Only one of them is shown to avoid cluttering the diagram. Notice that once four water molecules have bonded in this way, there isn't any more space available at the bonding level. All the empty orbitals from the original beryllium ion are being used. The water molecules arrange themselves to get as far apart as possible - which is pointing towards the corners of a tetrahedron. The ion therefore has a tetrahedral shape. The hydration of magnesium You might think that magnesium would behave just the same, but at the 3-level there are 3d orbitals available as well as 3s and 3p. When the magnesium ion is formed, it leaves empty 3s, 3p and 3d orbitals. When that ion is hydrated, it uses the 3s orbital, all three of the 3p orbitals and two of the 3d orbitals. These are reorganised to leave a total of six empty orbitals which are then used for bonding.

Why does magnesium stop at attaching six waters? Why doesn't it use the remaining 3d orbitals as well? You can't physically fit more than six water molecules around the magnesium - they take up too much room. What about the other ions in Group 2? As the ions get bigger, there is less tendency for them to form proper coordinate bonds with water molecules. The ions become so big that they aren't sufficiently attractive to the lone pairs on the water molecules to form formal bonds - instead the water molecules tend to cluster more loosely around the positive ions. Where they do form coordinate bonds with the water, however, they will be 6-coordinated just like the magnesium.

Beryllium hydroxide is amphoteric


Amphoteric means that it can react with both acids and bases to form salts. The other Group 2 hydroxides The other hydroxides of the Group 2 metals are all basic. They react with acids to form salts. For example:

Calcium hydroxide reacts with dilute hydrochloric acid to give calcium chloride and water. Beryllium hydroxide Beryllium hydroxide reacts with acids, forming solutions of beryllium salts. For example:

But it also reacts with bases such as sodium hydroxide solution. Beryllium hydroxide reacts with the sodium hydroxide to give a colourless solution of sodium tetrahydroxoberyllate.

This contains the complex ion, [Be(OH)4]2-. The name describes this ion. Tetra means four; hydroxo refers to the OH groups; beryllate shows that the beryllium is present in a negative ion. The "ate" ending always shows that the

ion is negative. A simple explanation of what is happening You need to think about where the beryllium hydroxide came from in the first place. It would probably have been made by adding sodium hydroxide solution to a solution of a beryllium salt like beryllium sulphate. Remember that beryllium ions in solution exist as the hydrated ion, [Be(H2O)4]2+. The beryllium has such a strongly polarising effect on the water molecules that hydrogen ions are very easily removed from them. The sodium hydroxide solution contains hydroxide ions which are powerful bases. If you add just the right amount of sodium hydroxide solution, you get a precipitate of what is normally called "beryllium hydroxide" - but which is a shade more complicated than that!

The product (other than water) is a neutral complex, and it is covalently bonded. All that has happened to the original complex ion is that two hydrogen ions have been removed from the water molecules. You get a precipitate of the neutral complex because of the lack of charge on it. There isn't enough attraction between this neutral complex and water molecules to bring it into solution. What happens if you add an acid to this? The hydrogen ions that were originally removed are simply replaced. The precipitate dissolves as the original hydrated beryllium ion is re-formed.

What happens if you add a base? Adding more hydroxide ions to the neutral complex pulls more hydrogen ions off the water molecules to give the tetrahydroxoberyllate ion:

The beryllium hydroxide dissolves because the neutral complex is converted into an ion which will be sufficiently attracted to water molecules. Why doesn't this happen with, for example, calcium hydroxide? Calcium hydroxide is truly ionic - and contains simple hydroxide ions, OH-. These react with hydrogen ions from an acid to form water - and so the hydroxide reacts with acids. However, there isn't any equivalent to the neutral complex. Adding more hydroxide ions from a base has no effect because they haven't got anything to react with.

Note: This has been simplified to bring it into line with the sort of treatment you will meet for the acid-base behaviour of transition metal hydroxides. In particular, the structure of beryllium hydroxide is probably even more complicated than has been suggested above!

PERIODIC TABLE GROUP 4 MENU

The trend from non-metal to metal . . . Discusses the trend from non-metallic to metallic behaviour as you go down Group 4. Describes the structures and physical properties of the elements, and some important atomic properties. Oxidation state trends . . . Looks at the increasing tendency towards an oxidation state of +2 as you go down the Group. Group 4 chlorides . . . Describes and explains the differences between the chlorides of carbon, silicon and lead as required by UK A level chemistry syllabuses. Group 4 oxides . . . Describes and explains the differences between the dioxides of carbon and silicon, and looks at the trend in acid-base behaviour of the oxides as you go down the Group. Some chemistry of aqueous lead(II) ions . . . Describes the reactions of lead(II) ions in solution with hydroxide ions, chloride ions, iodide ions and sulphate ions.

THE TREND FROM NON-METAL TO METAL IN THE GROUP 4 ELEMENTS


This page explores the trend from non-metallic to metallic behaviour in the Group 4 elements - carbon (C), silicon (Si), germanium (Ge), tin (Sn) and lead (Pb). It describes how this trend is shown in the structures and physical properties of the elements, and finally makes a not entirely successful attempt to explain the trend.

Structures and Physical Properties

Structures of the elements The trend from non-metal to metal as you go down the Group is clearly seen in the structures of the elements themselves. Carbon at the top of the Group has giant covalent structures in its two most familiar allotropes - diamond and graphite.

Allotropes: Two or more forms of the same element in the same physical state. The structures of diamond and graphite are explored in more detail on a page about giant covalent structuresin another part of this site. It would probably be worth your while to read this page before you go any further. Use the BACK button on your browser to return quickly to this page.

Diamond has a three-dimensional structure of carbon atoms each joined covalently to 4 other atoms. The diagram shows a small part of that structure.

Exactly this same structure is found in silicon and germanium and in one of the allotropes of tin - "grey tin" or "alpha-tin". The common allotrope of tin ("white tin" or "beta-tin") is metallic and has its atoms held together by metallic bonds. The structure is a distorted close-packed arrangement. In close-packing, each atom is surrounded by 12 nearneighbours. By the time you get to lead, the atoms are arranged in a straightforward 12-co-ordinated metallic structure.

Note: If you aren't sure about metallic bonding or metallic structures, you should follow these links before you go

any further. The first link will actually lead you to the second one if you want to explore both of these topics. Use the BACK button on your browser to return to this page.

There is therefore a clear trend from the typical covalency found in non-metals to the metallic bonding in metals, with the change-over obvious in the two entirely different structures found in tin. Physical properties of the elements Melting points and boiling points If you look at the trends in melting and boiling points as you go down Group 4, it is very difficult to make any sensible comments about the shift from covalent to metallic bonding. The trends reflect the increasing weakness of the covalent or metallic bonds as the atoms get bigger and the bonds get longer.

The low value for tin's melting point compared with lead is presumably due to tin forming a distorted 12-coordinated structure rather than a pure one. The tin values in the chart refer to metallic white tin.

Note: The data in this chart comes from the University of Sheffield's excellent Webelements site. There is an awful lot of variability in the data depending on where you get it from. I have to admit to choosing this set because it shows simple, largely unbroken patterns!

Brittleness There is a much clearer non-metal / metal difference shown if you look at the brittleness of the elements. Carbon as diamond is, of course, very hard - reflecting the strength of the covalent bonds. However, if you hit it

with a hammer, it shatters. Once you apply enough energy to break the existing carbon-carbon bonds, that's it! Silicon, germanium and grey tin (all with the same structure as diamond) are also brittle solids. However, white tin and lead have metallic structures. The atoms can roll over each other without any permanent disruption of the metallic bonds - leading to typical metallic properties like being malleable and ductile. Lead in particular is a fairly soft metal. Electrical conductivity Carbon as diamond doesn't conduct electricity. In diamond the electrons are all tightly bound and not free to move.

Note: In graphite, each atom donates one electron to a delocalised system of electrons which takes in the whole of its layer. These electrons are free to move around, and so graphite conducts electricity - but this is a special case. If you are interested, the bonding in graphite is like a vastly extended version of the bonding in benzene. Each 2 carbon atom undergoes sp 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.

Unlike diamond (which doesn't conduct electricity), silicon, germanium and grey tin are semiconductors.

Semiconductors: The theory of semiconductors lies outside A level chemistry, but briefly . . . When lots of atoms come together to make a giant structure, their atomic orbitals merge to produce a huge number of molecular orbitals, which arrange themselves in bands of increasing energy. One of these is often described as a valence band. The molecular orbitals in this band hold the electrons which make up the normal covalent (or metallic) bonding. The other band is called the conductance band. This usually has a higher energy than the valence band, and in something like diamond or silicon at absolute zero, the conductance band is empty of electrons. However, as electrons gain thermal energy as the temperature increases, some electrons may jump from the valence band into the conductance band - especially if the gap between the two is small. Once they are in the conductance band, they are delocalised from their original atoms and are free to move and conduct electricity. In diamond, the energy gap between the valence band and conductance band is too high for this to happen. In silicon, the band gap is small enough for electrons to jump, and so silicon is a semiconductor. If you are interested in this, you might like to try a Google search on silicon semiconductors band theory (or similar).

White tin and lead are normal metallic conductors of electricity. There is therefore a clear trend from the typically non-metallic conductivity behaviour of carbon as diamond, and the typically metallic behaviour of white tin and lead.

Trying to explain the trends


The main characteristic of metals is that they form positive ions. What we need to do is to look at the factors which increase the likelihood of positive ions being formed as you go down Group 4. Electronegativity Electronegativity is a measure of the tendency of an atom to attract a bonding pair of electrons. It is usually measured on the Pauling scale, where the most electronegative element (fluorine) is given an electronegativity of 4. The lower the electronegativity of an atom, the less strongly the atom attracts a bonding pair of electrons. That means that this atom will tend to lose the electron pair towards whatever else it is attached to. The atom we are interested in will therefore tend to carry either a partial positive charge or form a positive ion. Metallic behaviour is usually associated with a low electronegativity.

Note: If you aren't sure about electronegativity you really ought to read about it before you go any further. Use the BACK button on your browser to return quickly to this page.

So what happens to electronegativity in Group 4? Does it decrease as you go down the Group, suggesting a trend towards metallic behaviour?

Well! It certainly falls from carbon to silicon, but from there on it is a complete mess! There therefore seems to be no relationship between the non-metal to metal trend and the electronegativity values. Assuming the electronegativity values are correct, I am completely at a loss to understand this!

Note: The data in this chart again comes from the University of Sheffield's Webelements site. Again, there is an awful lot of variability in the data depending on where you get it from. But in no case that I have found is there any trend to lower electronegativities as you go down the Group. Older data sources give a fall from carbon (2.5) to silicon (1.8), but then give all the other elements in the Group the same value (all 1.8). If anyone reading this has a simple explanation for the lack of correlation between the trend to metallic behaviour and the electronegativity values, could you please contact me via the address on the about this sitepage.

Ionisation energies If you are thinking about the formation of positive ions, the obvious place to start looking is how ionisation energies change as you go down Group 4. Ionisation energies are defined as the energy needed to carry out each of the following changes. They are quoted in kJ mol-1. First ionisation energy:

Second ionisation energy:

. . . and so on.

Note: If you aren't sure about ionisation energies it would pay you to follow this link before you go any further. Use the BACK button on your browser to return quickly to this page.

None of the Group 4 elements form 1+ ions, so looking at the first ionisation energy alone isn't very helpful. Some of the elements do, however, form 2+ and (to some extent) 4+ ions. The first chart shows how the total ionisation energy needed to form the 2+ ions varies as you go down the Group. The values are all in kJ mol-1.

You can see that the ionisation energies tend to fall as you go down the Group - although there is a slightincrease at lead. The main trend is because:

The atoms are getting bigger because of the extra layers of electrons. The further the outer electrons are from the nucleus, the less they are attracted - and so the easier they are to remove. The outer electrons are screened from the full effect of the nucleus by the increasing number of inner electrons. These two effects outweigh the effect of the increasing nuclear charge.

Note: The reason for the oddity at lead is discussed in some more detail on a page about the oxidation statesshown by elements in Group 4. It isn't particularly important to the present discussion.

If you look at the amount of ionisation energy needed to form 4+ ions, the pattern is similar, but not entirely clear cut. Again, the values are all in kJ mol-1.

Note: The increase in total ionisation energy at lead is even more obvious in the case of the possible formation of 4+ ions. This is important when it comes to looking at the preferred oxidation states of lead.

What is clear looking at these two charts is that you have to put in large amounts of ionisation energy to form 2+ ions, and huge amounts to form 4+ ions. However, in each case there is a fall in ionisation energy as you go down the Group which makes it more likely that tin and lead could form positive ions - however, there is no indication from these figures that they are likely to form positive ions. The ionisation energies of carbon at the top of the Group are so huge that there is no possibility of it forming simple positive ions.

Note: Even for tin and lead, you have to put in huge amounts of energy to form either 2+ or 4+ ions. So why do they form ions at all? You have to remember that there are lots of other energy terms involved in the formation of an ionic compound apart from ionisation energy. Some of these give out large amounts of energy - for example, lattice enthalpy if you are forming an ionic solid, or hydration enthalpy if you are forming a solution. You will need to read about BornHaber cycles in order to understand this fully, and you might want to explore the energetics section of Chemguide,

or my chemistry calculations book.

S-ar putea să vă placă și