Sie sind auf Seite 1von 137

Chemguide: Support for CIE A level Chemistry Learning outcome 1(a)

This statement deals with the definitions of relative isotopic mass, relative atomic mass, relative molecular mass and relative formula mass. Before you go on, you should find and read the statement in your copy of the syllabus. Relative isotopic mass Isotopes are atoms of the same element (and so with the same number of protons and electrons), but with different masses due to having different numbers of neutrons. So, for example, chlorine has two different sorts of atoms each of which has 17 protons and 17 electrons. However, one of these atoms has 18 neutrons whereas the other has 20. All the chemical properties of the atoms are the same because they have the same numbers of protons and electrons, but they have different masses. The mass of an atom is incredibly small, and it doesn't make sense to use standard units like grams to measure it. Instead masses are measured on a scale based on the mass of an atom of the 12C isotope. On the 12C scale, the 12C isotope is given a mass of exactly 12 units, and the masses of all other isotopes are measured on the same scale. For example, an atom of 24Mg is twice as heavy as an atom of 12C, and so is given a relative isotopic mass of 24. An atom of the 1H isotope is only 1/12 of the mass of the 12C isotope and so is given a relative isotopic mass of 1.
Note: Actually, measured accurately, the hydrogen-1 isotope has a relative isotopic mass of about 1.008. At A level, for simplicity, the hydrogen value is just rounded off. That's true of quite a lot of other atoms as well.

There are two equivalent ways in which you can define relative isotopic mass. They mean exactly the same thing. Either: The relative isotopic mass of an isotope is the mass of the isotope on a scale on which a carbon-12 atom has a mass of exactly 12 units. Or: The relative isotopic mass of an isotope is the mass of the isotope relative to 1/12 of the mass of a carbon-12 atom. Learn whichever of these you have been taught. Relative atomic mass, Ar Similarly, relative atomic mass can be defined in two exactly equivalent ways: Either: The relative atomic mass of an element is the weighted average of the masses of its isotopes on a scale on which a carbon-12 atom has a mass of exactly 12 units. Or: The relative atomic mass of an element is the weighted average of the masses of its isotopes relative to 1/12 of the mass of a carbon-12 atom.

Again, learn whichever of these you have been taught. A "weighted average" allows for the fact that there won't be equal amounts of the various isotopes. This example should make that clear: In chlorine, there are 3 atoms of chlorine-35 for every 1 of chlorine-37. Suppose you had 4 typical atoms of chlorine. The total mass of these would be (3 x 35) + (1 x 37) = 142 The average mass of these 4 atoms would be 142 / 4 = 35.5. 35.5 is the relative atomic mass of chlorine. Notice the effect of the "weighted" average. A simple average of 35 and 37 is, of course, 36. Our answer of 35.5 allows for the fact that there are more of the lighter isotope of chlorine - and so the "weighted" average ought to be closer to that. You can always find the relative atomic mass of an element from a Periodic Table. But take care to choose the right number! Look at the key for the table, but, in any case, the relative atomic mass will always be the larger number given. Look at the Periodic Table provided by CIE for exam use, which you will find towards the end of the syllabus. Relative molecular mass, Mr You have to be careful with this term, because it should only be applied to substances which actually exist as molecules. A molecule consists of a fixed number of atoms joined together by covalent bonds. (Actually, in the case of the noble gases, a molecule can also be a single atom.) You shouldn't use the term for things, like sodium chloride, which are ionically bonded. Working out the relative molecular mass You work out the relative molecular mass of a substance by adding up the relative atomic masses of the atoms it consists of. So, for example, to work out the relative molecular mass of water, H2O, you add the relative atomic masses of two hydrogens and one oxygen. Mr of H2O = (2 x 1) + 16 = 18 To work out the relative molecular mass of CHCl 3: Mr of CHCl3 = 12 + 1 + (3 x 35.5) = 119.5 Defining the relative molecular mass Either: The relative molecular mass of a substance is the weighted average of the masses of the molecules on a scale on which a carbon-12 atom has a mass of exactly 12 units. Or: The relative molecular mass of a substance is the weighted average of the masses of the molecules relative to 1/12 of the mass of a carbon-12 atom.
Note: You may find that many sources miss out the bit about weighted averages, but this should be included unless you are thinking about the mass of a particular molecule with a particular combination of isotopes of the various atoms. For example, taking the example of CHCl 3 above: There is no single molecule of CHCl 3 which has a mass of 119.5. The problem is that an average sample of these molecules will contain isotopes of both chlorine-35 and chlorine-37. That means that

individual molecules can have the following masses: 12 + 1 + (3 x 35) = 118 12 + 1 + (2 x 35) + 37 = 120 12 + 1 + 35 + (2 x 37) = 122 12 + 1 + (3 x 37) = 124 The weighted average takes account of the proportions of each of these molecules in an average sample of the substance. Don't get too worried about all this! It is far more likely that you will have to work out a relative molecular mass by adding up the relative atomic masses than that you will have to define it.

Relative formula mass, Mr Notice that relative formula mass is given exactly the same symbol, M r, as relative molecular mass. In fact, relative formula mass is a much more useful term than relative molecular mass because it includes everything, whatever the bonding. It works just as well for ionic substances as for covalent substances. I strongly recommend that you use this term all the time, and only talk about relative molecular mass if you are specifically asked about it in an exam. Working out the relative formula mass Write down the formula, and then add up all the relative atomic masses of the atoms it contains. Example 1 The relative formula mass of NaCl = 23 + 35.5 = 58.5 Example 2 The relative formula mass of copper(II) sulfate crystals, CuSO 4.5H2O: Mr of CuSO4.5H2O = 63.5 + 32 + (4 x 16) + 5 x [(2 x 1) + 16] = 249.5
Note: The relative atomic mass of copper is often quoted as 64. I am using 63.5 here because that is the figure that comes from the CIE Periodic Table.

Be careful with things which contain water of crystallisation like the copper(II) sulfate crystals in this example. Add the water up first and then multiply it by 5 (or whatever other number you need). If you try to do it as hydrogen and oxygen separately, you stand a good chance of getting it wrong. Students usually remember to multiply the 2 hydrogens by 5, but forget to multiply the oxygen by 5. If you add the water up as a whole, that can't happen. Defining the relative formula mass I find it hard to imagine an exam question in which you were asked to define relative formula mass rather then just work it out, but just in case . . . Either: The relative formula mass of a substance is the weighted average of the masses of the formula units on a scale on which a carbon-12 atom has a mass of exactly 12 units. Or:

The relative formula mass of a substance is the weighted average of the masses of the formula units relative to 1/12 of the mass of a carbon-12 atom.

The "formula unit" is just the formula as you have written it.

Learning outcome 1(b) This statement deals with the mole concept and the Avogadro Constant. Working with moles I am going to start by explaining how you use moles rather than the reasoning behind them. Using them is easy, and you will have to be really comfortable about doing it, otherwise you won't be able to do most chemistry sums. Once you are confident about this, we can worry about the logic behind them. Working out the mass of a mole of a substance Easy! Work out the relative formula mass, and then add the unit "grams". Example 1: The mass of 1 mole of water, H2O The relative formula mass of water, H2O = (2 x 1) + 16 =18 1 mole of water, H2O, weighs 18 g. Example 2: The mass of 1 mole of sodium carbonate crystals, Na2CO3.10H2O The relative formula mass of sodium carbonate crystals, Na 2CO3.10H2O = (2 x 23) + 12 + (3 x 16) + 10 x [(2 x 1) + 16] = 286 1 mole of sodium carbonate crystals, Na 2CO3.10H2O, weighs 286 g The importance of quoting the formula Whenever you talk about a mole of something you must quote the formula. For example, if you were talking about a mole of sodium carbonate, you could mean a mole of anhydrous sodium carbonate, Na2CO3, or a mole of the crystals, Na2CO3.10H2O. The mass of 1 mole of the anhydrous compound is 106 g; the mass of a mole of the crystals is 286 g. Or if you were talking about the mass of a mole of oxygen, you might be talking about a mole of oxygen atoms, O, or a mole of oxygen molecules, O 2. Always write down the formula you are referring to. Converting between the mass of a substance and the number of moles You need to practise this until you can do it without really thinking about it. The formula is this:

So, if 1 mole of a substance weighs 40 g and you have 10 g of it, you have 10/40 moles - in other words you have 0.25 moles. Incidentally, the abbreviation for moles which you would normally use in calculations is mol.

If you had to work out other things, you could obviously rearrange this equation to give: and

If you are new to this, and your maths isn't very good, I can easily imagine that at this point you are feeling that this is a bit scary! There are suddenly a whole lot of equations which all look similar and confusing. Do you need to learn them all? No - in fact, you shouldn't even try! It would probably pay you to learn the first one, and then make sure that you could rearrange it to get the other two. Just practise doing it until you are sure there isn't a problem. But, in truth, you can do all of this perfectly well without learning any equations at all. You just need a bit of common sense. If you are good at maths, this next bit isn't really designed for you. Finding the number of moles using common sense Suppose the mass of 1 mole of a substance was 40 g, and you have 80 g of it. How many moles of it have you got? It is fairly obvious that you have twice as much as 1 mole - in other words, 2 moles. Suppose you have 400 g of the same substance. How many moles is this? It is obviously 10 times as much, and so you have 10 moles. Now let's make it more difficult. Suppose you have 47.2 g - how many moles is that? It isn't obvious this time, except that if you think about it, you have a bit more than 1 mole, but not as much as 2 moles. So what can you do with the numbers you have got (40 and 47.2) to get an answer like this? In the first two examples, what did you do with the numbers to get the answers 2 and 10? In each case you divided the mass by the mass of 1 mole (40 g in this case). So do the same thing again here. 47.2 divided by 40 = 1.18. Is that a sensible answer - a bit more than 1, but not as much as 2? Yes, it is. If you look at what you have done, you are actually using the formula above, but without deliberately slotting numbers into it. If you aren't very comfortable with maths, and have some awkward numbers to play with, replace them by easy numbers, decide what you would do with those, and then do the same thing with the more difficult ones. So what if you had to work it out for a mass like 2.4 g? Obviously, that's a lot less than 40 g, and so the answer is going to be a lot less than 1 mole. If you had an easy number like 80 g, you would divide the the 80 by the 40. Do the same thing here - divide 2.4 by 40.

That gives 0.06. Is that the sort of answer you want? Well, it is certainly a lot less than 1, so that is OK. Anything else you did with the numbers 2.4 and 40 would produce silly answers. Finding the mass using common sense Suppose you know the number of moles and the mass of 1 mole. How would you find the mass of substance that you have got? Well, you could use one of the equations above, but actually it is extremely easy - you really don't need an equation! If you have 2 moles of a substance where 1 mole weighs 100 g, what mass of the substance do you have? It is obvious that you have twice as much as the mass of 1 mole - 200 g. You really don't need an equation to solve this. But make it more difficult. If 1 mole weighs 100 g, what is the mass of 0.15 moles? With the simple number (2), how did you find the mass of the 2 moles? You multiplied the 100 by the 2. Do the same thing here - multiply the 100 by 0.15 to give 15 g. Is this a sensible answer? Well, suppose you were trying to work out how much 0.1 mole weighed. 0.1 of a mole is 1/10 of a mole, which would be 10 g in this case. 0.15 of a mole will weigh a bit more than that, and that's what you have got. Finding the mass of 1 mole using common sense Suppose 0.3 moles weighs 30 g. What is the mass of 1 mole? Again, you could use one of the equations above, but let's try it with a bit of common sense. These sums always look more difficult if one of your numbers is less than 1. So replace it with a simpler number - like, say, 2. If 2 moles weighed 30 g, what would 1 mole weigh? Obviously, you would just divide the 30 g by 2. If that's what you would do with a simple number, you are going to do exactly the same with a more awkward one. Divide 30 g by 0.3 to find that the mass of 1 mole is 100 g. It doesn't matter how you work these sums out as long as you always get them right. Practise and practise until you always succeed. Don't worry about using real examples (unless you need the practice in working out the mass of 1 mole of things) - just make the numbers up. If you are going to use the equations, only use the first one, and rearrange it if you need to. If you try to learn all three equations, you will probably get confused. Each time, check that your rearrangement is right by comparing it with the versions higher up the page. If you want to do it by a common sense method, check your answers (or get

someone else to check them) by slotting the numbers into one of the equations. Don't go on until you are certain that you can do these accurately every time. You won't be able to do any other mole calculations unless you can do these. Whichever way you do them, you should be able to get the right answer almost without thinking about it. What is the point of moles? Defining a mole Let's start with the hard bit, and define a mole. A mole of substance is the amount of that substance that contains the same number of stated elementary units as there are atoms in 12 g of carbon-12. "Stated elementary units" could be any of the following examples: chlorine atoms, Cl chlorine molecules, Cl2 ammonia molecules, NH3 sodium chloride formula units, NaCl and anything else that you care to write the formula of. 1 mole of any of these contains exactly the same number of the formula units you have written - the same as the number of atoms in 12 g of the carbon-12 isotope. That number is called the Avogadro Constant, and is given the symbol L. The Avogadro Constant is a huge number (approximately 6.02 x 10 23). It has units of mol-1, which you would read as "per mole". In other words, there are 6.02 x 1023 of whatever formula units you are talking about per mole of that formula. So, in 18 g of water, H2O, there are 6.02 x 1023 water molecules. In 58.5 g of sodium chloride, NaCl, there are 6.02 x 10 23 NaCl formula units. In 2 g of hydrogen, H2, there are 6.02 x 1023 hydrogen molecules. And so on, and so on . . . So what? Take a simple equation like: How do you read this? It could just be simply "Hydrogen and oxygen react to make water." Or it could be "2 molecules of hydrogen and 1 molecule of oxygen react to make 2 molecules of water." Moles enable you to do something much more useful. Notice that you need twice as many molecules of hydrogen as you do of oxygen. If you had 1 mole of hydrogen and 1 mole of oxygen, then they would contain the same number

of molecules - the Avogadro Constant. So if you wanted to be sure that you were reacting twice as many molecules of hydrogen as you had of oxygen, you would need twice as many moles. That means that you can read the equation in a new way. The equation above could say "2 moles of hydrogen react with 1 mole of oxygen to give 2 moles of water." And that's why moles are so important. For calculation work, you can interpret equations in terms of the number of moles of everything present - and you can then go on to turn moles into actual masses of substances reacting. For example, iron and sulfur react on heating to give iron(II) sulfide. You can read that as "1 mole of iron, Fe, reacts with 1 mole of sulfur, S, to give 1 mole of iron(II) sulfide, FeS." And by looking up the relative atomic masses, you find out that 56 g of iron will react with 32 g of sulfur to give 88 g of iron(II) sulfide. That lets you know what proportion to mix real iron and real sulfur together in, and how much product to expect. That's why moles are so important - they let you link equations to the actual quantities of materials you are using. Learning outcomes 1(c) and 1(d)
These statements deal with mass spectra and their interpretation.

AS-:

MASS SPECTROMETRY MENU


How the mass spectrometer works . . . An explanation of how a mass spectrum is produced The mass spectra of elements . . . How the mass spectrum of an element can be used to find its relative atomic mass. The mass spectra of organic compounds
The formation of fragmentation patterns . . . What the mass spectra of organic compounds look like, how the patterns are formed, and the sort of information you can get from them. The molecular ion (M+) peak . . . How the molecular ion peak can be used to find the relative formula mass of the compound and its molecular formula. The M+1 peak . . . How the M+1 peak arises, and its use in finding the number of carbon atoms a compound contains. Organic compounds containing halogen atoms . . . How the M+2 (and possibly M+4) peak arises, and its use in showing the presence of chlorine or bromine in a compound.

A2-:

Understanding Chemistry
MASS SPECTROMETRY MENU
How the mass spectrometer works . . . An explanation of how a mass spectrum is produced The mass spectra of elements . . . How the mass spectrum of an element can be used to find its relative atomic mass.

The mass spectra of organic compounds


The formation of fragmentation patterns . . . What the mass spectra of organic compounds look like, how the patterns are formed, and the sort of information you can get from them. The molecular ion (M+) peak . . . How the molecular ion peak can be used to find the relative formula mass of the compound and its molecular formula. The M+1 peak . . . How the M+1 peak arises, and its use in finding the number of carbon atoms a compound contains. Organic compounds containing halogen atoms . . . How the M+2 (and possibly M+4) peak arises, and its use in showing the presence of chlorine or bromine in a compound.

Learning outcome 1(e)


This statement deals with the meaning of the words "empirical formula" and "molecular formula". Before you go on, you should find and read the statement in your copy of the syllabus. Molecular formula This can obviously only apply to substances which consist of molecules substances with a fixed number of atoms joined together by covalent bonds. (Or in the case of the noble gases, a molecule can also be a single atom.) The term "molecular formula" can't be applied to ionic substances like sodium chloride. The molecular formula of the hydrocarbon ethane is C 2H6, for example, and the molecular formula of hydrogen peroxide is H 2O2. The molecular formula gives a precise count of the numbers of each sort of atom which make up the molecule. Empirical formula By contrast, the empirical formula tells you the simplest ratio of the various atoms present in a substance. For example, in ethane (above) the ratio of the number of

carbon to hydrogen atoms is 1:3. The empirical formula is CH 3. For hydrogen peroxide, where the simplest ratio is 1:1, the empirical formula is HO. The empirical formulae for most molecular substances are virtually never used, unless they happen to be the same as the molecular formula - as in H2O, for example. So what is the point of them? The word "empirical" means "derived from observation or experiment" - so an empirical formula is one which you can find by doing experiments, as you will see in Learning Outcome 1(f). In other words, you can do an experiment which will tell you that the empirical formula of a particular hydrocarbon (not ethane) is CH 2, for example. That in itself isn't very helpful. The hydrocarbon could be C 2H4, C3H6, C4H8, and so on and so on - anything with a carbon to hydrogen ratio of 1:2. To find out the correct molecular formula from the empirical formula, you would have to do further experiments. You will find out how to do that in Outcome 1(f) as well. The empirical formula is just a stage on the way to finding out the molecular formula of something. The empirical formula and ionic compounds For ionic compounds, like sodium chloride, the formula quoted isalmost always the empirical formula. In an ionic compound, there are no fixed numbers of ions - it depends on how big the crystal is. So the formula of sodium chloride is simply given as NaCl, showing the 1:1 ratio. The formula of sodium oxide is Na2O, showing a 2:1 ratio.
Note: You may come across a few ionic compounds such as sodium peroxide, Na2O2, or mercury(I) chloride, Hg2Cl2, where the formula normally used isn't the empirical formula. There are good reasons for this which would be unnecessarily confusing to discuss now. Almost all ionic compounds use the empirical formula.

In a real crystal of sodium chloride or sodium oxide, there will be some huge variable number of positive and negative ions. The formula we write just tells us what the ratio is.

Learning outcome 1(f)


This statement deals with how you work out empirical formulae and molecular formulae from experimental data. It expects you to be able to do it in two ways - from data about the masses or percentages of the various elements combined together, or from data obtained from combustion experiments. You also have to be able to work out a molecular formula from an empirical formula.

Before you go on, you should find and read the statement in your copy of the syllabus. Finding empirical formulae from masses or percentages If you have a formula like, say, CH4, you can read this as saying that 1 mole of carbon atoms are combined with 4 moles of hydrogen atoms. In a different example, if you could work out that phosphorus and oxygen atoms combined together in the ratio of 2 moles of phosphorus atoms to 3 moles of oxygen atoms, then you would know that the empirical formula was P2O3. You don't, of course know anything about the molecular formula. All you know is that the ratio is 2:3. The molecular formula could equally well be P 4O6 or P6O9 or whatever. You can find mole ratios from data involving either the masses or percentages of the combining atoms. Finding empirical formulae from mass data Suppose you found that 0.46 g of sodium formed 0.78 g of sodium sulfide. That means that 0.46 g of sodium combines with (0.78 - 0.46) g = 0.32 g of sulfur. It is clearest if you set your answer out as a simple table: Na mass number of moles of atoms 0.46 g 0.46/23 = 0.02 ratio 2 S 0.32 g 0.32/32 = 0.01 1

That would tell you that the empirical formula was Na 2S. Finding empirical formulae from percentage data You might have been given the last example in a different form. You could have been told that the compound contained 59.0% of Na and 41.0% of S by mass. That's not a problem. If you had 100 g of the compound, then the masses of sodium and sulfur would be 59.0 g and 41.0 g respectively. So use those figures in a sum like the last one. Try it, and make sure you get the same answer as before.
Note: You may find that this time it isn't as easy to spot the ratio.

If it isn't immediately obvious, try dividing through by the smallest number. That will almost invariably help.

Because this topic is covered in my chemistry calculations book (see pages 31 to 33), I can't go any further than this with online help. For reasons that I have explained on another page, all I can do in these cases is to refer you to the book. Finding empirical formulae from combustion data This isn't covered explicitly in my book, and so I can give a complete explanation here. If you burn a compound containing carbon and hydrogen in an excess of air or oxygen, carbon dioxide and water are formed. If you can trap the water and carbon dioxide separately, you can find out what mass of each is formed. From that, you can work out the empirical formula. How you do this varies slightly depending on whether the compound contains anything else as well as the carbon and hydrogen. We'll look at the two cases separately. If the compound only contains carbon and hydrogen A compound which only contains carbon and hydrogen is called a hydrocarbon. Look for this word in the question. You can only use this shorter method if you know that there is nothing else present. Here's an example: When 0.78 g of a hydrocarbon was burned in excess air, 2.64 g of carbon dioxide and 0.54 g of water were formed. Find the empirical formula of the hydrocarbon. The important things to notice is that every mole of CO 2 contains 1 mole of carbon atoms. Every mole of H2O contains 2 moles of hydrogen atoms. 1 mole of CO2 weighs 44 g. No of moles of CO2 = 2.64/44 = 0.06 Therefore, no of moles of carbon atoms, C = 0.06 1 mole of H2O weighs 18 g. No of moles of H2O = 0.54/18 = 0.03 Each mole of H2O contains 2 moles of hydrogen atoms. Therefore, no of moles of hydrogen atoms, H = 0.06 The ratio of the number of moles of C : H is 1 : 1. The empirical formula is CH. It has taken several lines to write this down, but it is a simple calculation.
Note: If you are writing down a complete calculation rather than filling in answers in a structured exam paper, it is important to include lots of words. Notice how many words there are in the simple calculation above compared with the amount of numbers. You have to

include enough words to make it absolutely clear what you are doing

If the compound contains other things as well as carbon and hydrogen This also applies to cases where you haven't been told that the compound is a hydrocarbon. It might be a hydrocarbon, but you aren't sure. The most likely cases you will come across will probably contain oxygen as well as carbon and hydrogen. In these cases, you have to put in extra steps in order to find out how much oxygen (or whatever) is present. The sequence is: Use the carbon dioxide figure to work out the number of moles of carbon atoms. From this, work out the mass of carbon present. Use the water figure to work out the number of moles of hydrogen atoms. From this, work out the mass of hydrogen present. Add up the carbon and hydrogen masses, and compare them with the original mass of compound. If there is no difference, then you have a hydrocarbon, and the problem disappears. Just look at the ratio of the numbers of moles of carbon and hydrogen atoms. If there is a difference, assume the difference is due to oxygen unless you are told differently. Work out the mass of oxygen present. Find the number of moles of oxygen atoms present. From the mole figures for carbon, hydrogen and oxygen, work out the empirical formula. This all seems a bit long-winded, but it is simple enough if you follow the stages through in an example. The question: 0.23 g of a compound containing carbon, hydrogen and possibly oxygen was burned in an excess of air. 0.44 g of carbon dioxide and 0.27 g of water were formed. Work out the empirical formula of the compound. No of moles of CO2 = 0.44/44 = 0.01 Therefore, no of moles of carbon atoms, C = 0.01 and mass of carbon = 0.01 x 12 g = 0.12 g No of moles of H2O = 0.27/18 = 0.015 Each mole of water contains 2 moles of hydrogen atoms. Therefore, no of moles of hydrogen atoms, H = 2 x 0.015 = 0.03 and mass of hydrogen = 0.03 x 1 g = 0.03 g Total mass of carbon and hydrogen in compound = 0.12 + 0.03 g = 0.15 g Since you started with 0.23 g of compound, there is missing mass which must be oxygen. Mass of oxygen = 0.23 - 0.15 g = 0.08 g

No of moles of oxygen atoms, O = 0.08/16 = 0.005 The ratio of number of moles is C 0.01 : H 0.03 : O 0.005 Dividing through by the smallest number to give a simple ratio gives: C2:H6:O1 The empirical formula is C2H6O. Converting empirical formulae into molecular formulae This is really simple! You can do it if you are told either the relative formula mass of the compound or the mass of 1 mole (which is just the relative formula mass expressed in grams). Example 1 Let's follow on with the example above which resulted in an empirical formula of CH. Suppose you knew that the relative formula mass was 78. If you add up the relative formula mass of the empirical formula, CH, it comes to 12 + 1 = 13. The true molecular formula must be some multiple of this - so how many times does 13 go into 78? Dividing 78 by 13 gives 6, and so the molecular formula must be 6 times bigger than CH - in other words C6H6. Example 2 And let's also follow on with the other example above which resulted in an empirical formula of C2H6O. Suppose you knew that the mass of 1 mole was 46 g. 1 mole of the empirical formula, C2H6O, would have a mass of (2 x 12) + (6 x 1) + 16 g = 46 g. That's the same as the mass of 1 mole that you are given. Therefore the molecular formula must be the same as the empirical formula C2H6O.

CHAPTER 2-: ATOMIC STRUCTURE

Learning outcome 2(a)


This statement deals with the definitions of relative charges and masses of protons, neutrons and electrons.

A SIMPLE VIEW OF ATOMIC STRUCTURE


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

The sub-atomic particles


Protons, neutrons and electrons. relative mass proton 1 neutron 1 electron 1/1836 relative charge +1 0 -1

Beyond A'level: Protons and neutrons don't in fact haveexactly the same mass - neither of them has a mass of exactly 1 on the carbon-12 scale (the scale on which the relative masses of atoms are measured). On the carbon-12 scale, a proton has a mass of 1.0073, and a neutron a mass of 1.0087.

The behaviour of protons, neutrons and electrons in electric fields What happens if a beam of each of these particles is passed between two electrically charged plates - one positive and one negative? Opposites will attract. Protons are positively charged and so would be deflected on a curving path towards the negative plate. Electrons are negatively charged and so would be deflected on a curving path towards the positive plate. Neutrons don't have a charge, and so would continue on in a straight line. Exactly what happens depends on whether the beams of particles enter the electric field with the various particles having the same speeds or the same energies If the particles have the same energy If beams of the three sorts of particles, all with the same energy, are passed between two electrically charged plates: Protons are deflected on a curved path towards the negative plate. Electrons are deflected on a curved path towards the positive plate. The amount of deflection is exactly the same in the electron beam as the proton beam if the energies are the same - but, of course, it is in the opposite direction. Neutrons continue in a straight line.

If the electric field was strong enough, then the electron and proton beams might curve enough to hit their respective plates. If the particles have the same speeds If beams of the three sorts of particles, all with the same speed, are passed between two electrically charged plates: Protons are deflected on a curved path towards the negative plate. Electrons are deflected on a curved path towards the positive plate. If the electrons and protons are travelling with the same speed, then the lighter electrons are deflected far more strongly than the heavier protons. Neutrons continue in a straight line.

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

The nucleus
The nucleus is at the centre of the atom and contains the protons and neutrons. Protons and neutrons are collectively

known asnucleons. Virtually all the mass of the atom is concentrated in the nucleus, because the electrons weigh so little. Working out the numbers of protons and neutrons No of protons = ATOMIC NUMBER of the atom The atomic number is also given the more descriptive name ofproton number. No of protons + no of neutrons = MASS NUMBER of the atom The mass number is also called the nucleon number. This information can be given simply in the form:

How many protons and neutrons has this atom got? The atomic number counts the number of protons (9); the mass number counts protons + neutrons (19). If there are 9 protons, there must be 10 neutrons for the total to add up to 19. The atomic number is tied to the position of the element in the Periodic Table and therefore the number of protons defines what sort of element you are talking about. So if an atom has 8 protons (atomic number = 8), it must be oxygen. If an atom has 12 protons (atomic number = 12), it must be magnesium. Similarly, every chlorine atom (atomic number = 17) has 17 protons; every uranium atom (atomic number = 92) has 92 protons. Isotopes The number of neutrons in an atom can vary within small limits. For example, there are three kinds of carbon atom 12C, 13C and 14C. They all have the same number of protons, but the number of neutrons varies. protons neutrons mass number carbon-12 6 6 12 carbon-13 6 7 13 carbon-14 6 8 14 These different atoms of carbon are called isotopes. The fact that they have varying numbers of neutrons makes no difference whatsoever to the chemical reactions of the carbon. Isotopes are atoms which have the same atomic number but different mass numbers. They have the same number of protons

but different numbers of neutrons.

The electrons
Working out the number of electrons Atoms are electrically neutral, and the positiveness of the protons is balanced by the negativeness of the electrons. It follows that in a neutral atom: no of electrons = no of protons So, if an oxygen atom (atomic number = 8) has 8 protons, it must also have 8 electrons; if a chlorine atom (atomic number = 17) has 17 protons, it must also have 17 electrons. The arrangement of the electrons The electrons are found at considerable distances from the nucleus in a series of levels called energy levels. Each energy level can only hold a certain number of electrons. The first level (nearest the nucleus) will only hold 2 electrons, the second holds 8, and the third also seems to be full when it has 8 electrons. At GCSE you stop there because the pattern gets more complicated after that. These levels can be thought of as getting progressively further from the nucleus. Electrons will always go into the lowest possible energy level (nearest the nucleus) - provided there is space. To work out the electronic arrangement of an atom Look up the atomic number in the Periodic Table - making sure that you choose the right number if two numbers are given. The atomic number will always be the smaller one. This tells you the number of protons, and hence the number of electrons. Arrange the electrons in levels, always filling up an inner level before you go to an outer one. e.g. to find the electronic arrangement in chlorine The Periodic Table gives you the atomic number of 17. Therefore there are 17 protons and 17 electrons. The arrangement of the electrons will be 2, 8, 7 (i.e. 2 in the first level, 8 in the second, and 7 in the third). The electronic arrangements of the first 20 elements

After this the pattern alters as you enter the transition series in the Periodic Table. Two important generalisations If you look at the patterns in this table: The number of electrons in the outer level is the same as the group number. (Except with helium which has only 2 electrons. The noble gases are also usually called group 0 - not group 8.) This pattern extends throughout the Periodic Table for the main groups (i.e. not including the transition elements). So if you know that barium is in group 2, it has 2 electrons in its outer level; iodine (group 7) has 7 electrons in its outer level; lead (group 4) has 4 electrons in its outer level. Noble gases have full outer levels. This generalisation will need modifying for A'level purposes. Dots-and-crosses diagrams In any introductory chemistry course you will have come across the electronic structures of hydrogen and carbon, for example, drawn as:

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

The circles show energy levels - representing increasing

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

Thinking of the arrangement of the electrons in this way makes a useful bridge to the A'level view.

Learning outcome 2(b) This statement deals with what happens to protons, neutrons and electrons in electric fields. Before you go on, you should find and read the statement in your copy of the syllabus. You will this discussed near the beginning of the page which exploresa simple view of atomic structure. On that page, you will find that I have given two different versions of what happens depending on whether the beams of different particles have the same energy or the same speed. At the time of writing (July 2010), it isn't clear which of these CIE want. There was only one multiple choice question on this topic up to that point, and it gave diagrams suggesting (but without saying so) that the beams were of equal energy. However, I have discussed this with someone at CIE, who thought it was more appropriate to use the curves for equal speeds. I will update this as soon as I am more certain which version they want. Learning outcome 2(c)
This statement deals with the way that mass and charge are arranged in the atom. Before you go on, you should find and read the statement in your copy of the syllabus. This is completely trivial! The mass of the atom is concentrated in the nucleus because that is where you find the protons and neutrons. The electrons are found at some distance from the nucleus, and contribute very little to its overall mass.

Because of the protons, the nucleus is positively charged. The electrons surrounding the nucleus are, of course, negatively charged.

Learning outcome 2(d)


This statement expects you to be able to work out how many protons, neutrons and electrons there are in an atom or ion from basic information. Before you go on, you should find and read the statement in your copy of the syllabus. First read the page a simple view of atomic structure, paying particular attention to the terms atomic number, proton number, mass number and nucleon number. It is important that you can recognise all of these, even though, in the CIE syllabus, only proton number and nucleon number are used. The important thing to notice is that the proton number (atomic number) counts the number of protons in the nucleus. In a neutral atom that also tells you the number of electrons. That isn't as simple in an ion, though. We will look at a couple of examples of ions later on. The nucleon number (mass number) counts the total number of protons and neutrons. You can work out the number of neutrons present by subtracting the proton number from the nucleon number. Some examples: Fe: proton number 26; nucleon number 56 There are 26 protons (from the proton number). This is a neutral atom (not an ion) and so there are 26 electrons. There are a total of 56 protons and neutrons (from the nucleon number), and so there must be 30 neutrons (56 - 26). S2-: proton number 16; nucleon number 32 This is an ion - take care! There are 16 protons (from the proton number). If it was a neutral atom, there would be 16 electrons. But this is an ion with 2 negative charges. That's because it has 2 extra electrons. There are 18 electrons. There are a total of 32 protons and neutrons (from the nucleon number), and so there must be 16 neutrons (32 - 16). Notice that the fact that you have an ion rather than a neutral atom makes no difference whatsoever to the number of protons or neutrons. Al3+: proton number 13; nucleon number 27 Another ion - take care with the electrons.

There are 13 protons (from the proton number). If it was a neutral atom, there would be 13 electrons. However, this is a 3+ ion. That's because it has lost 3 electrons. There are 10 electrons. And you can work out the number of neutrons from the nucleon number. In this case there are 14 neutrons (27 - 13).
Note: This all assumes that you know from earlier work what an ion is. If you are starting A level without any chemistry background, an ion is an electrically charged atom or group of atoms. Negative ions are negative because they have one or more extra electrons. Positive ions are positive because they have lost one or more electrons. The ion is negative or positive because there are no longer equal numbers of protons and electrons.

Learning outcome 2(e)


This statement is in two parts. The first part is essentially a repeat of learning outcome 2(d), and I am not going to add anything extra about that. The second part deals with isotopes. Before you go on, you should find and read the statement in your copy of the syllabus. Isotopes You will find a short piece about isotopes about a third of the way down the page a simple view of atomic structure, and that is really all you need. If you are asked to define isotopes you should include the facts that: They are atoms of the same element, having the same proton (or atomic) number but with different nucleon (or mass) numbers because they have different numbers of neutrons.

Learning outcomes 2(f) and 2(g) These statements concern s, p and d orbitals. I am treating the two statements together because I don't actually think it make sense to split them into two parts. As long as you read and understand the pages I am pointing you to below, you will know and understand everything you need for these two statements. Before you go on, you should find and read the statements in your copy of the syllabus. You will find everything you need to know on the page about atomic orbitals. Near the top of that page, you will find a link to another page about the difference between the words "orbit" and "orbital". Follow that link! It is really important that you understand the difference between

these words, and then permanently forget any ideas that you ever had about electrons orbiting a nucleus!

ATOMIC ORBITALS
This page explains what atomic orbitals are in a way that makes them understandable for introductory courses such as UK A level and its equivalents. It explores s and p orbitals in some detail, including their shapes and energies. d orbitals are described only in terms of their energy, and f orbitals only get a passing mention.

What is an atomic orbital?


Orbitals and orbits When a planet moves around the sun, you can plot a definite path for it which is called an orbit. A simple view of the atom looks similar and you may have pictured the electrons as orbiting around the nucleus. The truth is different, and electrons in fact inhabit regions of space known as orbitals. Orbits and orbitals sound similar, but they have quite different meanings. It is essential that you understand the difference between them. The impossibility of drawing orbits for electrons To plot a path for something you need to know exactly where the object is and be able to work out exactly where it's going to be an instant later. You can't do this for electrons. The Heisenberg Uncertainty Principle says - loosely - that you can't know with certainty both where an electron is and where it's going next. (What it actually says is that it is impossible to define with absolute precision, at the same time, both the position and the momentum of an electron.) That makes it impossible to plot an orbit for an electron around a nucleus. Is this a big problem? No. If something is impossible, you have to accept it and find a way around it.
Note: Over the years I have had a steady drip of questions from students in which it is obvious that they still think of electrons as orbiting around a nucleus - which is completely wrong! I have added a page about why the idea of orbits is wrong to try to avoid having to say the same thing over and over again!

Hydrogen's electron - the 1s orbital

Note: In this diagram (and the orbital diagrams that follow), the nucleus is shown very much larger than it really is. This is just for clarity.

Suppose you had a single hydrogen atom and at a particular instant plotted the position of the one electron. Soon afterwards, you do the same thing, and find that it is in a new position. You have no idea how it got from the first place to the second. You keep on doing this over and over again, and gradually build up a sort of 3D map of the places that the electron is likely to be found. In the hydrogen case, the electron can be found anywhere within a spherical space surrounding the nucleus. The diagram shows across-section through this spherical space. 95% of the time (or any other percentage you choose), the electron will be found within a fairly easily defined region of space quite close to the nucleus. Such a region of space is called an orbital.You can think of an orbital as being the region of space in which the electron lives.
Note: If you wanted to be absolutely 100% sure of where the electron is, you would have to draw an orbital the size of the Universe!

What is the electron doing in the orbital? We don't know, we can't know, and so we just ignore the problem! All you can say is that if an electron is in a particular orbital it will have a particular definable energy. Each orbital has a name. The orbital occupied by the hydrogen electron is called a 1s orbital. The "1" represents the fact that the orbital is in the energy level closest to the nucleus. The "s" tells you about the shape of the orbital. s orbitals are spherically symmetric around the nucleus - in each case, like a hollow ball made of rather chunky material with the nucleus at its centre. The orbital on the left is a 2s orbital.This is similar to a 1s orbital except that the region where there is the greatest chance of finding the electron is further from the nucleus - this is an orbital at the second energy level. If you look carefully, you will notice that there is another region of slightly higher electron density (where the dots are thicker) nearer the nucleus. ("Electron density" is another way of talking about how likely you are to find an electron at a particular place.) 2s (and 3s, 4s, etc) electrons spend some of their time closer to the nucleus than you might expect. The effect of this is to slightly reduce the energy of electrons in s orbitals. The nearer the nucleus the electrons get, the lower their energy. 3s, 4s (etc) orbitals get progressively further from the nucleus. p orbitals Not all electrons inhabit s orbitals (in fact, very few electrons live in s orbitals). At the first energy level, the only orbital available to electrons is the 1s orbital, but at the second level, as well as a 2s orbital, there are also orbitals called 2p orbitals. A p orbital is rather like 2 identical balloons tied together at the nucleus. The diagram on the left is a cross-section through that 3-dimensional region of space. Once again, the orbital shows where there is a 95% chance of finding a particular electron.
Taking chemistry further: If you imagine a horizontal plane through the nucleus, with one lobe of the orbital above the plane and the other beneath it, there is a zero probability

of finding the electron on that plane. So how does the electron get from one lobe to the other if it can never pass through the plane of the nucleus? At this introductory level you just have to accept that it does! If you want to find out more, read about the wave nature of electrons.

Unlike an s orbital, a p orbital points in a particular direction - the one drawn points up and down the page. At any one energy level it is possible to have three absolutely equivalent p orbitals pointing mutually at right angles to each other. These are arbitrarily given the symbols px, py and pz. This is simply for convenience - what you might think of as the x, y or z direction changes constantly as the atom tumbles in space. The p orbitals at the second energy level are called 2px, 2py and 2pz. There are similar orbitals at subsequent levels - 3px, 3py, 3pz, 4px, 4py, 4pz and so on. All levels except for the first level have p orbitals. At the higher levels the lobes get more elongated, with the most likely place to find the electron more distant from the nucleus. d and f orbitals In addition to s and p orbitals, there are two other sets of orbitals which become available for electrons to inhabit at higher energy levels. At the third level, there is a set of five d orbitals (with complicated shapes and names) as well as the 3s and 3p orbitals (3px, 3py, 3pz). At the third level there are a total of nine orbitals altogether. At the fourth level, as well the 4s and 4p and 4d orbitals there are an additional seven f orbitals - 16 orbitals in all. s, p, d and f orbitals are then available at all higher energy levels as well. For the moment, you need to be aware that there are sets of five d orbitals at levels from the third level upwards, but you probably won't be expected to draw them or name them. Apart from a passing reference, you won't come across f orbitals at all.
Note: Some UK-based syllabuses will eventually want you to be able to draw, or at least recognise, the shapes of d orbitals. I am not including them now because I don't want to add confusion to what is already a difficult introductory topic.

Check your syllabus and past papers to find out what you need to know. If you are a studying a UK-based syllabus and haven't got these, follow this link to find out how to get hold of them.

Fitting electrons into orbitals


You can think of an atom as a very bizarre house (like an inverted pyramid!) - with the nucleus living on the ground floor, and then various rooms (orbitals) on the higher floors occupied by the electrons. On the first floor there is only 1 room (the 1s orbital); on the second floor there are 4 rooms (the 2s, 2p x, 2py and 2pzorbitals); on the third floor there are 9 rooms (one 3s orbital, three 3p orbitals and five 3d orbitals); and so on. But the rooms aren't very big . . . Each orbital can only hold 2 electrons. A convenient way of showing the orbitals that the electrons live in is to draw "electrons-in-boxes". "Electrons-in-boxes" Orbitals can be represented as boxes with the electrons in them shown as arrows. Often an up-arrow and a down-arrow are used to show that the electrons are in some way different.
Taking chemistry further: The need to have all electrons in an atom different comes out of quantum theory. If they live in different orbitals, that's fine - but if they are both in the same orbital there has to be some subtle distinction between them. Quantum theory allocates them a property known as "spin" - which is what the arrows are intended to suggest.

A 1s orbital holding 2 electrons would be drawn as shown on the right, but it can be written even more quickly as 1s2. This is read as "one s two" not as "one s squared". You mustn't confuse the two numbers in this notation:

The order of filling orbitals Electrons fill low energy orbitals (closer to the nucleus) before they fill higher energy ones. Where there is a

choice between orbitals of equal energy, they fill the orbitals singly as far as possible. This filling of orbitals singly where possible is known as Hund's rule. It only applies where the orbitals have exactly the same energies (as with p orbitals, for example), and helps to minimise the repulsions between electrons and so makes the atom more stable. The diagram (not to scale) summarises the energies of the orbitals up to the 4p level.

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

Learning outcome 2(h) This statement deals with writing electronic configurations for atoms and ions using s, p, d notation. Before you go on, you should find and read the statement in your copy of the syllabus. You have to be able to write the electronic configuration of an atom or one of its ions given the proton number (atomic number) and charge. You can find the proton number from a Periodic Table, of course, and you will either have to know or be told the charge. You will find how to do this for atoms explained on the page aboutelectronic structures of atoms. At the bottom of that page, there is a link to a second page which explains how to do it for ions. Don't go on to this second page until you are sure you understand the first one. The only way of being sure is to practise doing it. Use a Periodic Table to find the proton numbers, and then practise working out the electronic structure for random atoms up to krypton in the Periodic Table, checking your answers against examples given on the first page. Once you are certain that you can do it for atoms from hydrogen to krypton, try a few examples of bigger atoms in the s and p blocks. Start with iodine and barium so that you can check your answers with the text on the first page. Then try a few more - you will have to use your initiative to find a way of checking them. (Actually, they aren't hard to find from a Google search!) Once you are sure you can do this, go on and read, and then practise, the second page about the electronic structures of ions. You will have to restrict yourself to ions whose charges you know.

ELECTRONIC STRUCTURES
This page explores how you write electronic structures for atoms using s, p, and d notation. It assumes that you know about simple atomic orbitals - at least as far as the way they are named, and their relative energies. If you want to look at the electronic structures of simple monatomic ions (such as Cl-, Ca2+ and Cr3+), you will find a link at the bottom of the page.
Important! If you haven't already read the page on atomic

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

The electronic structures of atoms


Relating orbital filling to the Periodic Table

UK syllabuses for 16 - 18 year olds tend to stop at krypton when it comes to writing electronic structures, but it is possible that you could be asked for structures for elements up as far as barium. After barium you have to worry about f orbitals as well as s, p and d orbitals - and that's a problem for chemistry at a higher level. It is important that you look through past exam papers as well as your syllabus so that you can judge how hard the questions are likely to get. This page looks in detail at the elements in the shortened version of the Periodic Table above, and then shows how you could work out the structures of some bigger atoms.
Important! You must have a copy of your syllabus and copies of recent exam papers. If you are studying a UK-based syllabus and haven't got them, follow this link to find out how to get hold of them.

The first period Hydrogen has its only electron in the 1s orbital - 1s1, and at helium the first level is completely full - 1s2. The second period Now we need to start filling the second level, and hence start the second period. Lithium's electron goes into the 2s orbital because that has a lower energy than the 2p orbitals. Lithium has an electronic structure of 1s22s1. Beryllium adds a second electron to this same level - 1s22s2. Now the 2p levels start to fill. These levels all have the same energy, and so the electrons go in singly at first. 1s22s22px1 B 1s22s22px12py1 C 1s22s22px12py12pz1 N

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

The next electrons to go in will have to pair up with those already there. 1s22s22px22py12pz1 O 1s22s22px22py22pz1 F 1s22s22px22py22pz2 Ne You can see that it is going to get progressively tedious to write the full electronic structures of atoms as the number of electrons increases. There are two ways around this, and you must be familiar with both. Shortcut 1: All the various p electrons can be lumped together. For example, fluorine could be written as 1s 22s22p5, and neon as 1s22s22p6. This is what is normally done if the electrons are in an inner layer. If the electrons are in the bonding level (those on the outside of the atom), they are sometimes written in shorthand, sometimes in full. Don't worry about this. Be prepared to meet either version, but if you are asked for the electronic structure of something in an exam, write it out in full showing all the p x, py and pz orbitals in the outer level separately. For example, although we haven't yet met the electronic structure of chlorine, you could write it as 1s22s22p63s23px23py23pz1. Notice that the 2p electrons are all lumped together whereas the 3p ones are shown in full. The logic is that the 3p electrons will be involved in bonding because they are on the outside of the atom, whereas the 2p electrons are buried deep in the atom and aren't really of any interest. Shortcut 2: You can lump all the inner electrons together using, for example, the symbol [Ne]. In this context, [Ne] means the electronic structure of neon - in other words: 1s22s22px22py22pz2 You wouldn't do this with helium because it takes longer to write [He] than it does 1s2. On this basis the structure of chlorine would be written [Ne]3s23px23py23pz1. The third period At neon, all the second level orbitals are full, and so after this we have to start the third period with sodium. The pattern of filling is now exactly the same as in the previous period,

except that everything is now happening at the 3-level. For example: short version 2 2 6 2 Mg 1s 2s 2p 3s [Ne]3s2 1s22s22p63s23px23py13pz [Ne]3s23px23py13pz S 1 1 Ar 1s22s22p63s23px23py23pz
2

[Ne]3s23px23py23pz
2

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

The beginning of the fourth period At this point the 3-level orbitals aren't all full - the 3d levels haven't been used yet. But if you refer back to the energies of the orbitals, you will see that the next lowest energy orbital is the 4s - so that fills next. 1s22s22p63s23p64s1 K 1s22s22p63s23p64s2 Ca There is strong evidence for this in the similarities in the chemistry of elements like sodium (1s22s22p63s1) and potassium (1s22s22p63s23p64s1) The outer electron governs their properties and that electron is in the same sort of orbital in both of the elements. That wouldn't be true if the outer electron in potassium was 3d 1. s- and p-block elements

The elements in group 1 of the Periodic Table all have an outer electronic structure of ns1 (where n is a number between 2 and 7). All group 2 elements have an outer electronic

structure of ns2. Elements in groups 1 and 2 are described as s-block elements. Elements from group 3 across to the noble gases all have their outer electrons in p orbitals. These are then described as p-block elements. d-block elements

Remember that the 4s orbital has a lower energy than the 3d orbitals and so fills first. Once the 3d orbitals have filled up, the next electrons go into the 4p orbitals as you would expect. d-block elements are elements in which the last electron to be added to the atom is in a d orbital. The first series of these contains the elements from scandium to zinc, which at GCSE you probably called transition elements or transition metals. The terms "transition element" and "d-block element" don't quite have the same meaning, but it doesn't matter in the present context.
If you are interested: A transition element is defined as one which has partially filled d orbitals either in the element or any of its compounds. Zinc (at the right-hand end of the d-block) always has a completely full 3d level (3d10) and so doesn't count as a transition element.

d electrons are almost always described as, for example, d 5 or d8 - and not written as separate orbitals. Remember that there are five d orbitals, and that the electrons will inhabit them singly as far as possible. Up to 5 electrons will occupy orbitals on their own. After that they will have to pair up. d5 means d8 means

Notice in what follows that all the 3-level orbitals are written together, even though the 3d electrons are added to the atom after the 4s. 1s22s22p63s23p63d14s2 Sc 1s22s22p63s23p63d24s2 Ti 1s22s22p63s23p63d34s2 V 1s22s22p63s23p63d54s1 Cr Whoops! Chromium breaks the sequence. In chromium, the electrons in the 3d and 4s orbitals rearrange so that there is one electron in each orbital. It would be convenient if the sequence was tidy - but it's not! (back to being tidy 1s22s22p63s23p63d54s2 Mn again) 2 2 6 2 6 6 2 1s 2s 2p 3s 3p 3d 4s Fe 1s22s22p63s23p63d74s2 Co 1s22s22p63s23p63d84s2 Ni 1s22s22p63s23p63d104s1 (another awkward one!) Cu 1s22s22p63s23p63d104s2 Zn And at zinc the process of filling the d orbitals is complete. Filling the rest of period 4 The next orbitals to be used are the 4p, and these fill in exactly the same way as the 2p or 3p. We are back now with the p-block elements from gallium to krypton. Bromine, for example, is 1s22s22p63s23p63d104s24px24py24pz1.
Useful exercise: Work out the electronic structures of all the elements from gallium to krypton. You can check your answers by comparing them with the elements directly above them in the Periodic Table. For example, gallium will have the same sort of arrangement of its outer level electrons as boron or aluminium - except that gallium's outer electrons will be in the 4-level.

Summary Writing the electronic structure of an element from hydrogen to krypton Use the Periodic Table to find the atomic number, and hence number of electrons. Fill up orbitals in the order 1s, 2s, 2p, 3s, 3p, 4s, 3d, 4p - until you run out of electrons. The 3d is the awkward one - remember that specially. Fill p and d orbitals singly as far as possible before pairing electrons up.

Remember that chromium and copper have electronic structures which break the pattern in the first row of the d-block.

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


Note: We are deliberately excluding the d-block elements apart from the first row that we've already looked at in detail. The pattern of awkward structures isn't the same in the other rows. This is a problem for degree level.

First work out the number of outer electrons. This is quite likely all you will be asked to do anyway. The number of outer electrons is the same as the group number. (The noble gases are a bit of a problem here, because they are normally called group 0 rather then group 8. Helium has 2 outer electrons; the rest have 8.) All elements in group 3, for example, have 3 electrons in their outer level. Fit these electrons into s and p orbitals as necessary. Which level orbitals? Count the periods in the Periodic Table (not forgetting the one with H and He in it). Iodine is in group 7 and so has 7 outer electrons. It is in the fifth period and so its electrons will be in 5s and 5p orbitals. Iodine has the outer structure 5s25px25py25pz1. What about the inner electrons if you need to work them out as well? The 1, 2 and 3 levels will all be full, and so will the 4s, 4p and 4d. The 4f levels don't fill until after anything you will be asked about at A'level. Just forget about them! That gives the full structure: 1s22s22p63s23p63d104s24p64d105s25px25py25pz1. When you've finished, count all the electrons to make sure that they come to the same as the atomic number. Don't forget to make this check - it's easy to miss an orbital out when it gets this complicated. Barium is in group 2 and so has 2 outer electrons. It is in the sixth period. Barium has the outer structure 6s2. Including all the inner levels: 1s22s22p63s23p63d104s24p64d105s25p66s2. It would be easy to include 5d10 as well by mistake, but the d level always fills after the next s level - so 5d fills after 6s just as 3d fills after 4s. As long as you counted the number of electrons you could easily spot this mistake because you

would have 10 too many.

ELECTRONIC STRUCTURES OF IONS


This page explores how you write electronic structures for simple monatomic ions (ions containing only one atom) using s, p, and d notation. It assumes that you already understand how to write electronic structures for atoms.
Important! If you have come straight to this page via a search engine, you should read the page on electronic structures of atoms before you go any further.

Working out the electronic structures of ions


Ions are atoms (or groups of atoms) which carry an electric charge because they have either gained or lost one or more electrons. If an atom gains electrons it acquires a negative charge. If it loses electrons, it becomes positively charged. The electronic structure of s- and p-block ions Write the electronic structure for the neutral atom, and then add (for a negative ion) or subtract electrons (for a positive ion). To write the electronic structure for Cl -: but Cl- has one more 2 2 6 2 2 2 1 Cl 1s 2s 2p 3s 3px 3py 3pz electron 2 2 6 2 2 2 2 Cl 1s 2s 2p 3s 3px 3py 3pz To write the electronic structure for O2-: but O2- has two more O 1s22s22px22py12pz1 electrons 22 2 2 2 2 O 1s 2s 2px 2py 2pz To write the electronic structure for Na+: Na 1s22s22p63s1 but Na+ has one less electron Na+ 1s22s22p6 To write the electronic structure for Ca2+: but Ca2+ has two less Ca 1s22s22p63s23p64s2 electrons 2+ 2 2 6 2 6 Ca 1s 2s 2p 3s 3p

The electronic structure of d-block ions Here you are faced with one of the most irritating facts in A'level chemistry! You will recall that the first transition series (from scandium to zinc) is the result of the 3d orbitals being filled after the 4s orbital. However, once the electrons are established in their orbitals, the energy order changes - and in all the chemistry of the transition elements, the 4s orbital behaves as the outermost, highest energy orbital. The reversed order of the 3d and 4s orbitals only applies to building the atom up in the first place. In all other respects, the 4s electrons are always the electrons you need to think about first. You must remember this: When d-block elements form ions, the 4s electrons are lost first. Provided you remember that, working out the structure of a dblock ion is no different from working out the structure of, say, a sodium ion. To write the electronic structure for Cr3+: Cr 1s22s22p63s23p63d54s1 Cr3+ 1s22s22p63s23p63d3 The 4s electron is lost first followed by two of the 3d electrons. To write the electronic structure for Zn 2+: Zn 1s22s22p63s23p63d104s2 Zn2+ 1s22s22p63s23p63d10 This time there is no need to use any of the 3d electrons. To write the electronic structure for Fe3+: Fe 1s22s22p63s23p63d64s2 Fe3+ 1s22s22p63s23p63d5 The 4s electrons are lost first followed by one of the 3d electrons. The rule is quite simple. Take the 4s electrons off first, and then as many 3d electrons as necessary to produce the correct positive charge.
Note: You may well have the impression from GCSE that ions have to have noble gas structures. It's not true! Most (but not all) ions formed by s- and p-block elements do have noble gas structures, but if you look at the d-block ions we've used as examples, not one of them has a noble gas structure - yet they

are all perfectly valid ions. Getting away from a reliance on the concept of noble gas structures is one of the difficult mental leaps that you have to make at the beginning of A'level chemistry.

Learning outcome 2(i) This statement deals with first ionisation energy and the way it varies across and down the Periodic Table. Before you go on, you should find and read the statement in your copy of the syllabus. You will find everything that you need to know on the page about first ionisation energy. You will find this really difficult if you aren't already confident about electronic structures using s and p notation. Don't even think about looking at this page until you are happy about that.

IONISATION ENERGY
This page explains what first ionisation energy is, and then looks at the way it varies around the Periodic Table - across periods and down groups. It assumes that you know about simple atomic orbitals, and can write electronic structures for simple atoms. You will find a link at the bottom of the page to a similar description of successive ionisation energies (second, third and so on).
Important! If you aren't reasonable happy about atomic orbitals and electronic structures you should follow these links before you go any further.

Defining first ionisation energy


Definition The first ionisation energy is the energy required to remove the most loosely held electron from one mole of gaseous atoms to produce 1 mole of gaseous ions each with a charge of 1+. This is more easily seen in symbol terms. It is the energy needed to carry out this change per mole of X.
Worried about moles? Don't be! For now, just take it as a

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

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

Patterns of first ionisation energies in the Periodic Table


The first 20 elements

First ionisation energy shows periodicity. That means that it varies in a repetitive way as you move through the Periodic Table. For example, look at the pattern from Li to Ne, and then compare it with the identical pattern from Na to Ar. These variations in first ionisation energy can all be explained in terms of the structures of the atoms involved. Factors affecting the size of ionisation energy Ionisation energy is a measure of the energy needed to pull a particular electron away from the attraction of the nucleus. A high value of ionisation energy shows a high attraction between the electron and the nucleus. The size of that attraction will be governed by: The charge on the nucleus. The more protons there are in the nucleus, the more positively charged the nucleus is, and the more strongly electrons are

attracted to it. The distance of the electron from the nucleus. Attraction falls off very rapidly with distance. An electron close to the nucleus will be much more strongly attracted than one further away. The number of electrons between the outer electrons and the nucleus. Consider a sodium atom, with the electronic structure 2,8,1. (There's no reason why you can't use this notation if it's useful!) If the outer electron looks in towards the nucleus, it doesn't see the nucleus sharply. Between it and the nucleus there are the two layers of electrons in the first and second levels. The 11 protons in the sodium's nucleus have their effect cut down by the 10 inner electrons. The outer electron therefore only feels a net pull of approximately 1+ from the centre. This lessening of the pull of the nucleus by inner electrons is known as screening or shielding.
Warning! Electrons don't, of course, "look in" towards the nucleus and they don't "see" anything either! But there's no reason why you can't imagine it in these terms if it helps you to visualise what's happening. Just don't use these terms in an exam! You may get an examiner who is upset by this sort of loose language.

Whether the electron is on its own in an orbital or paired with another electron. Two electrons in the same orbital experience a bit of repulsion from each other. This offsets the attraction of the nucleus, so that paired electrons are removed rather more easily than you might expect. Explaining the pattern in the first few elements Hydrogen has an electronic structure of 1s1. It is a very small atom, and the single electron is close to the nucleus and therefore strongly attracted. There are no electrons screening it from the nucleus and so the ionisation energy is high (1310 kJ mol-1). Helium has a structure 1s2. The electron is being removed from the same orbital as in hydrogen's case. It is close to the nucleus and unscreened. The value of the ionisation energy (2370 kJ mol-1) is much higher than hydrogen, because the nucleus now has 2 protons attracting the electrons instead of 1. Lithium is 1s22s1. Its outer electron is in the second energy

level, much more distant from the nucleus. You might argue that that would be offset by the additional proton in the nucleus, but the electron doesn't feel the full pull of the nucleus - it is screened by the 1s2 electrons.

You can think of the electron as feeling a net 1+ pull from the centre (3 protons offset by the two 1s2 electrons). If you compare lithium with hydrogen (instead of with helium), the hydrogen's electron also feels a 1+ pull from the nucleus, but the distance is much greater with lithium. Lithium's first ionisation energy drops to 519 kJ mol -1 whereas hydrogen's is 1310 kJ mol-1. The patterns in periods 2 and 3 Talking through the next 17 atoms one at a time would take ages. We can do it much more neatly by explaining the main trends in these periods, and then accounting for the exceptions to these trends. The first thing to realise is that the patterns in the two periods are identical - the difference being that the ionisation energies in period 3 are all lower than those in period 2.

Explaining the general trend across periods 2 and 3 The general trend is for ionisation energies to increase across a period. In the whole of period 2, the outer electrons are in 2-level orbitals - 2s or 2p. These are all the same sort of distances from the nucleus, and are screened by the same 1s2 electrons. The major difference is the increasing number of protons in the nucleus as you go from lithium to neon. That causes greater attraction between the nucleus and the electrons and so increases the ionisation energies. In fact the increasing nuclear charge also drags the outer electrons in closer to the nucleus. That increases ionisation energies still more as you go across the period.
Note: Factors affecting atomic radius are covered on a separate page.

In period 3, the trend is exactly the same. This time, all the electrons being removed are in the third level and are screened by the 1s22s22p6 electrons. They all have the same sort of environment, but there is an increasing nuclear charge. Why the drop between groups 2 and 3 (Be-B and Mg-Al)? The explanation lies with the structures of boron and aluminium. The outer electron is removed more easily from these atoms than the general trend in their period would suggest. Be 1s22s2 1st I.E. = 900 kJ mol -1 B 1s22s22px1 1st I.E. = 799 kJ mol -1 You might expect the boron value to be more than the beryllium value because of the extra proton. Offsetting that is the fact that boron's outer electron is in a 2p orbital rather than a 2s. 2p

orbitals have a slightly higher energy than the 2s orbital, and the electron is, on average, to be found further from the nucleus. This has two effects. The increased distance results in a reduced attraction and so a reduced ionisation energy. 2 The 2p orbital is screened not only by the 1s electrons 2 but, to some extent, by the 2s electrons as well. That also reduces the pull from the nucleus and so lowers the ionisation energy. The explanation for the drop between magnesium and aluminium is the same, except that everything is happening at the 3-level rather than the 2-level. Mg 1s22s22p63s2 1st I.E. = 736 kJ mol -1 Al 1s22s22p63s23px1 1st I.E. = 577 kJ mol -1 The 3p electron in aluminium is slightly more distant from the nucleus than the 3s, and partially screened by the 3s 2 electrons as well as the inner electrons. Both of these factors offset the effect of the extra proton.
Warning! You might possibly come across a text book which describes the drop between group 2 and group 3 by saying that a full s2 orbital is in some way especially stable and that makes the electron more difficult to remove. In other words, that the fluctuation is because the group 2 value for ionisation energy is abnormally high. This is quite simply wrong! The reason for the fluctuation is because the group 3 value is lower than you might expect for the reasons we've looked at.

Why the drop between groups 5 and 6 (N-O and P-S)? Once again, you might expect the ionisation energy of the group 6 element to be higher than that of group 5 because of the extra proton. What is offsetting it this time? N 1s22s22px12py12pz1 1st I.E. = 1400 kJ mol -1 O 1s22s22px22py12pz1 1st I.E. = 1310 kJ mol -1 The screening is identical (from the 1s2 and, to some extent, from the 2s2 electrons), and the electron is being removed from an identical orbital. The difference is that in the oxygen case the electron being removed is one of the 2px2 pair. The repulsion between the two electrons in the same orbital means that the electron is easier to remove than it would otherwise be. The drop in ionisation energy at sulphur is accounted for in the same way.

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

Why is the sodium value less than that of lithium? There are 11 protons in a sodium atom but only 3 in a lithium atom, so the nuclear charge is much greater. You might have expected a much larger ionisation energy in sodium, but offsetting the nuclear charge is a greater distance from the nucleus and more screening. Li 1s22s1 1st I.E. = 519 kJ mol-1 Na 1s22s22p63s1 1st I.E. = 494 kJ mol-1 Lithium's outer electron is in the second level, and only has the 1s2electrons to screen it. The 2s1 electron feels the pull of 3 protons screened by 2 electrons - a net pull from the centre of 1+. The sodium's outer electron is in the third level, and is screened from the 11 protons in the nucleus by a total of 10 inner electrons. The 3s1 electron also feels a net pull of 1+ from the centre of the atom. In other words, the effect of the extra protons is compensated for by the effect of the extra screening electrons. The only factor left is the extra distance between the outer electron and the nucleus in sodium's case. That lowers the ionisation energy. Similar explanations hold as you go down the rest of this group or, indeed, any other group. Trends in ionisation energy in a transition series

Apart from zinc at the end, the other ionisation energies are all much the same. All of these elements have an electronic structure [Ar]3dn4s2 (or 4s1 in the cases of chromium and copper). The electron being lost always comes from the 4s orbital.
Note: Confusingly, once the orbitals have electrons in them, the 4s orbital has a higher energy than the 3d - quite the opposite of their order when the atoms are being filled with electrons. That means that it is a 4s electron which is lost from the atom when it forms an ion. It also means that the 3d orbitals are slightly closer to the nucleus than the 4s - and so offer some screening. You will find this commented on in the page about electronic structures of ions.

As you go from one atom to the next in the series, the number of protons in the nucleus increases, but so also does the number of 3d electrons. The 3d electrons have some screening effect, and the extra proton and the extra 3d electron more or less cancel each other out as far as attraction from the centre of the atom is concerned. The rise at zinc is easy to explain. Cu [Ar]3d104s1 1st I.E. = 745 kJ mol -1 Zn [Ar]3d104s2 1st I.E. = 908 kJ mol -1 In each case, the electron is coming from the same orbital, with identical screening, but the zinc has one extra proton in the nucleus and so the attraction is greater. There will be a degree of repulsion between the paired up electrons in the 4s orbital, but in this case it obviously isn't enough to outweigh the effect of the extra proton.
Note: This is actually very similar to the increase from, say, sodium to magnesium in the third period. In that case, the outer electronic structure is going from 3s1 to 3s2. Despite the pairing-up of the electrons, the ionisation energy increases because of the extra proton in the nucleus. The repulsion between the 3s electrons obviously isn't enough to outweigh this either. I don't know why the repulsion between the paired electrons matters

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

Ionisation energies and reactivity


The lower the ionisation energy, the more easily this change happens: You can explain the increase in reactivity of the Group 1 metals (Li, Na, K, Rb, Cs) as you go down the group in terms of the fall in ionisation energy. Whatever these metals react with, they have to form positive ions in the process, and so the lower the ionisation energy, the more easily those ions will form. The danger with this approach is that the formation of the positive ion is only one stage in a multi-step process. For example, you wouldn't be starting with gaseous atoms; nor would you end up with gaseous positive ions - you would end up with ions in a solid or in solution. The energy changes in these processes also vary from element to element. Ideally you need to consider the whole picture and not just one small part of it. However, the ionisation energies of the elements are going to be major contributing factors towards the activation energy of the reactions. Remember that activation energy is the minimum energy needed before a reaction will take place. The lower the activation energy, the faster the reaction will be - irrespective of what the overall energy changes in the reaction are. The fall in ionisation energy as you go down a group will lead to lower activation energies and therefore faster reactions.
Note: You will find a page discussing this in more detail in the inorganic section of this site dealing with the reactions of Group 2 metals with water.

Learning outcomes 2(j) and 2(k) These statements deal with successive ionisation energies and the information you can get from them. Before you go on, you should find and read the statement in your

copy of the syllabus. You should then read the page about successive ionisation energies. These two statements look confusingly similar. Learning outcome 2(k) seems to me to be the easier one, so we will have a quick look at that first. Learning outcome 2(k) In this case, you are likely to be given a graph or a set of data for the various ionisation energies of an element, and be asked which Group of the Periodic Table it is in. This is covered on the Chemguide page above, but here is another example. The first eight ionisation energies of an element are: 1310 3390 5320 7450 11000 13300 71000 84100 Which Group of the Periodic Table is it in? Look for the big jump. In this case, the values are increasing by a couple of thousand at a time until you get to the gap between the sixth and seventh values, where the jump is huge. So, six electrons are relatively easy to remove, but the seventh is far more difficult. That is because it is being removed from an inner level. The element is in Group 6 of the Periodic Table. If you can get hold of a copy, there is a multi-choice question using a different example (this time with the values on a graph rather than as numbers) on November 2007 Paper 1 Q4. The answer is C. Learning outcome 2(j) You will find a multi-choice question related to this statement on June 2009 Paper 1 Q3, which relates to the same data as above. The difference this time is that you are asked for the outer electronic configuration rather than the Group number. Basically, all you need to do is to work out what Group the element is in, and then work out the arrangement of the outer electrons from that. We already know that it is in Group 6. Group 6 elements will have the electronic structure ns2np4. In the question, you were told that the element was between lithium and neon in the Periodic Table, and so the outer structure would be 2s22p4. If you had to write down an answer instead of choosing one from a list, you might also have written it in a more detailed form as 2s22px22py12pz1.

SUCCESSIVE IONISATION ENERGIES


This page explains what second, third, (etc) ionisation energy means, and then looks at patterns in successive ionisation energies for selected elements. It assumes that you understand about first ionisation energy.
Important! If you have come straight to this page via a search engine, you should read the page on first ionisation energy before you go any further.

Defining second ionisation energy Second ionisation energy is defined by the equation: It is the energy needed to remove a second electron from each ion in 1 mole of gaseous 1+ ions to give gaseous 2+ ions. More ionisation energies You can then have as many successive ionisation energies as there are electrons in the original atom. The first four ionisation energies of aluminium, for example, are given by 1st I.E. = 577 kJ mol -1 2nd I.E. = 1820 kJ mol -1 3rd I.E. = 2740 kJ mol -1 4th I.E. = 11600 kJ mol1

In order to form an Al 3+(g) ion from Al(g) you would have to supply: 577 + 1820 + 2740 = 5137 kJ mol -1 That's a lot of energy. Why, then, does aluminium form Al3+ ions? It can only form them if it can get that energy back from somewhere, and whether that's feasible depends on what it is reacting with. For example, if aluminium reacts with fluorine or oxygen, it can recover that energy in various changes involving the fluorine or oxygen - and so aluminium fluoride or aluminium oxide contain Al3+ ions. If it reacts with chlorine, it can't recover sufficient energy, and so

solid anhydrous aluminium chloride isn't actually ionic - instead, it forms covalent bonds. Why doesn't aluminium form an Al 4+ ion? The fourth ionisation energy is huge compared with the first three, and there is nothing that aluminium can react with which would enable it to recover that amount of extra energy. Why do successive ionisation energies get larger? Once you have removed the first electron you are left with a positive ion. Trying to remove a negative electron from a positive ion is going to be more difficult than removing it from an atom. Removing an electron from a 2+ or 3+ (etc) ion is going to be progressively more difficult. Why is the fourth ionisation energy of aluminium so large? The electronic structure of aluminium is 1s22s22p63s23px1. The first three electrons to be removed are the three electrons in the 3p and 3s orbitals. Once they've gone, the fourth electron is removed from the 2p level - much closer to the nucleus, and only screened by the 1s2 (and to some extent the 2s2) electrons. Using ionisation energies to work out which group an element is in This big jump between two successive ionisation energies is typical of suddenly breaking in to an inner level. You can use this to work out which group of the Periodic Table an element is in from its successive ionisation energies. Magnesium (1s22s22p63s2) is in group 2 of the Periodic Table and has successive ionisation energies:

Here the big jump occurs after the second ionisation energy. It means that there are 2 electrons which are relatively easy to remove (the 3s2 electrons), while the third one is much more difficult (because it comes from an inner level - closer to the nucleus and with less screening). Silicon (1s22s22p63s23px13py1) is in group 4 of the Periodic Table and has successive ionisation energies:

Here the big jump comes after the fourth electron has been removed. The first 4 electrons are coming from the 3-level orbitals; the fifth from the 2-level. The lesson from all this: Count the easy electrons - those up to (but not including) the big jump. That is the same as the group number. Another example: Decide which group an atom is in if it has successive ionisation energies:

The ionisation energies are going up one or two thousand at a time for the first five. Then there is a huge jump of about 15000. There are 5 relatively easy electrons - so the element is in group 5. Exploring the patterns in more detail If you plot graphs of successive ionisation energies for a particular element, you can see the fluctuations in it caused by the different electrons being removed. Not only can you see the big jumps in ionisation energy when an electron comes from an inner level, but you can also see the minor fluctuations within a level depending on whether the electron is coming from an s or a p orbital, and even whether it is paired or unpaired in that orbital. Chlorine has the electronic structure 1s22s22p63s23px23py23pz1. This graph plots the first eight ionisation energies of chlorine. The green labels show which electron is being removed for each of the ionisation energies.

If you put a ruler on the first and second points to establish the trend, you'll find that the third, fourth and fifth points lie above the value you would expect. That is because the first two electrons are coming from pairs in the 3p levels and are therefore rather easier to remove than if they were unpaired. Again, if you put a ruler on the 3rd, 4th and 5th points to establish their trend, you'll find that the 6th and 7th points lie well above the values you would expect from a continuation of the trend. That is because the 6th and 7th electrons are coming from the 3s level - slightly closer to the nucleus and slightly less well screened. The massive jump as you break into the inner level at the 8th electron is fairly obvious!
Warning! People sometimes get confused with these graphs because they forget that they are removing electrons from the atom. For example, the first point refers to the first electron being lost - from a 3p orbital. Basically, you start from the outside of the atom and work in towards the middle. If you start from the 1s orbital and work outwards, you are doomed to failure!

To plot any more ionisation energies for chlorine needs a change of vertical scale. The seventeenth ionisation energy of chlorine is nearly 400,000 kJ mol -1, and the vertical scale has to be squashed to accommodate this.

This is now a "log graph" - plotted by finding the logarithm of each ionisation energy (press the "log" button on your calculator). This doesn't simply squash the vertical scale. It distorts it as well, to such an extent that the only useful thing the graph now shows is the major jumps where the next electron to be removed comes from an inner level. The distortion is so great in the first 8 ionisation energies, for example, that the patterns shown by the previous graph are completely (and misleadingly) destroyed.

Chemical Bonding
Chemguide: Support for CIE A level Chemistry Learning outcome 3(a)
This statement deals with ionic (electrovalent) bonding. Before you go on, you should find and read the statement in your copy of the syllabus. You should read the page which looks at ionic (electrovalent) bonding. This page actually goes beyond what this statement asks for, but you should read the whole of it. That includes the bit below the red warning notice, although you don't need to try to remember this last bit. What is important is that you break away from any view of bonding that you may have built up at an earlier stage which over-rates the importance of noble gas structures. As the page points out, there are far more ions which don't have noble gas structures than there are which do have noble gas structures. And I think you should at least read the last bit of that page so that you can see that the real reasons for the formation of a particular ion lie in the energetics of the process. You will notice that the syllabus statement talks about "dot and cross

diagrams", whereas there isn't a single dot and cross diagram on my Chemguide page. That's because they are a complete waste of time to draw for ionic compounds! It is much easier and quicker to write, say, 2,8,6, than to draw three circles and put the correct numbers of dots or crosses on each one.
In an exam, of course, if you are asked to show ionic bonding using dot and cross diagrams, then that is what you must do. Don't expect to get any credit for writing, for example, 2,8,6, if the question specifically asks for a dot and cross diagram.

IONIC (ELECTROVALENT) BONDING


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

A simple view of ionic bonding


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

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

The sodium has lost an electron, so it no longer has equal numbers of electrons and protons. Because it has one more proton than electron, it has a charge of 1+. If electrons are lost from an atom, positive ions are formed. Positive ions are sometimes called cations. The chlorine has gained an electron, so it now has one more electron than proton. It therefore has a charge of 1-. If electrons are gained by an atom, negative ions are formed. A negative ion is sometimes called an anion. The nature of the bond The sodium ions and chloride ions are held together by the strong electrostatic attractions between the positive and negative charges. The formula of sodium chloride You need one sodium atom to provide the extra electron for one chlorine atom, so they combine together 1:1. The formula is therefore NaCl. Some other examples of ionic bonding magnesium oxide

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

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

Again, noble gas structures are formed. It takes two potassiums to supply the electrons the oxygen needs. The formula of potassium oxide is K2O.

THE A'LEVEL VIEW OF IONIC BONDING


Electrons are transferred from one atom to another resulting in the formation of positive and negative ions. The electrostatic attractions between the positive and negative ions hold the compound together. So what's new? At heart - nothing. What needs modifying is the view that there is something magic about noble gas structures. There are far more ions which don't have noble gas structures than there are which do. Some common ions which don't have noble gas structures You may have come across some of the following ions in a basic course like GCSE. They are all perfectly stable , but not one of them has a noble gas structure. Fe3+ [Ar]3d5 Cu2+ [Ar]3d9 Zn2+ [Ar]3d10 Ag+ [Kr]4d10 Pb2+ [Xe]4f145d106s2 Noble gases (apart from helium) have an outer electronic structure ns2np6.
Note: If you aren't happy about writing electronic structuresusing of s, p and d notation, follow this link before you go on. Return to this page via the menus or by using the BACK button on your browser.

Apart from some elements at the beginning of a transition series (scandium forming Sc3+ with an argon structure, for example), all transition elements and any metals following a transition series (like tin and lead in Group 4, for example) will have structures like those above.

That means that the only elements to form positive ions with noble gas structures (apart from odd ones like scandium) are those in groups 1 and 2 of the Periodic Table and aluminium in group 3 (boron in group 3 doesn't form ions). Negative ions are tidier! Those elements in Groups 5, 6 and 7 which form simple negative ions all have noble gas structures. If elements aren't aiming for noble gas structures when they form ions, what decides how many electrons are transferred? The answer lies in the energetics of the process by which the compound is made.
Warning! From here to the bottom of this page goes beyond anything you are likely to need for A'level purposes. It is included for interest only.

What determines what the charge is on an ion? Elements combine to make the compound which is as stable as possible - the one in which the greatest amount of energy is evolved in its making. The more charges a positive ion has, the greater the attraction towards its accompanying negative ion. The greater the attraction, the more energy is released when the ions come together. That means that elements forming positive ions will tend to give away as many electrons as possible. But there's a down-side to this. Energy is needed to remove electrons from atoms. This is calledionisation energy. The more electrons you remove, the greater the total ionisation energy becomes. Eventually the total ionisation energy needed becomes so great that the energy released when the attractions are set up between positive and negative ions isn't large enough to cover it. The element forms the ion which makes the compound most stable - the one in which most energy is released over-all. For example, why is calcium chloride CaCl 2 rather than CaCl or CaCl3? If one mole of CaCl (containing Ca+ ions) is made from its elements, it is possible to estimate that about 171 kJ of heat is evolved. However, making CaCl 2 (containing Ca2+ ions) releases more heat. You get 795 kJ. That extra amount of heat evolved makes the compound more stable, which is why you get CaCl 2 rather than CaCl.

What about CaCl3 (containing Ca3+ ions)? To make one mole of this, you can estimate that you would have to put in 1341 kJ. This makes this compound completely non-viable. Why is so much heat needed to make CaCl 3? It is because the third ionisation energy (the energy needed to remove the third electron) is extremely high (4940 kJ mol -1) because the electron is being removed from the 3-level rather than the 4-level. Because it is much closer to the nucleus than the first two electrons removed, it is going to be held much more strongly.
Note: It would pay you to read about ionisation energies if you really want to understand this. You could also go to a standard text book and investigate Born-Haber Cycles.

A similar sort of argument applies to the negative ion. For example, oxygen forms an O2- ion rather than an O- ion or an O3ion, because compounds containing the O2- ion turn out to be the most energetically stable.

Learning outcome 3(b) This statement is in two parts and deals with simple covalent bonding, and co-ordinate (dative covalent) bonding. Before you go on, you should find and read the statement in your copy of the syllabus. Learning outcome 3(b)(i) This covers simple covalent bonding in terms of dot and cross diagrams. You should read a part of the page which looks at covalent bonding single bonds. Stop when you get to the heading which says "A more sophisticated view of covalent bonding". We will come back to that later on in this section. Up to that point, you will have come across hydrogen, chlorine, hydrogen chloride and methane from the syllabus statement. Then follow the link at the bottom of the page to find out about double bonding in oxygen, carbon dioxide and ethene. All you need to read for now on that page is the first section. Ignore everything when it starts to get more sophisticated.

Learning outcome 3(b)(ii) This covers co-ordinate covalent bonding, also called dative covalent bonding. It doesn't matter which of these terms you have been taught - you must be familiar with both of them, and realise that they mean exactly the same thing. You should read a part of the page which looks at co-ordinate (dative covalent) bonding. The syllabus only mentions the ammonium ion, and the bonding in the aluminium chloride dimer. That means that you must be familiar with both of those. However, you might be asked about other things as well if they involve the same principle of one atom donating both electrons to the bonding pair. So look at the other examples, and make sure they make sense to you. Make sure that you understand everything down as far as (but not including) "The bonding in hydrated metal ions". All that you will eventually have to know from the rest of the page is the bonding in carbon monoxide. Don't worry about it for now.

COVALENT BONDING - SINGLE BONDS


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

A simple view of covalent bonding


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

diagram.

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

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

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

Covalent bonding at A'level


Cases where there isn't any difference from the simple view If you stick closely to modern A'level syllabuses, there is little need to move far from the simple (GCSE) view. The only thing which must be changed is the over-reliance on the concept of noble gas structures. Most of the simple molecules you draw do in fact have all their atoms with noble gas structures. For example:

Even with a more complicated molecule like PCl 3, there's no problem. In this case, only the outer electrons are shown for simplicity. Each atom in this structure has inner layers of electrons of 2,8. Again, everything present has a noble gas structure.

Cases where the simple view throws up problems Boron trifluoride, BF3

A boron atom only has 3 electrons in its outer level, and there is no possibility of it reaching a noble gas structure by simple sharing of electrons. Is this a problem? No. The boron has formed the maximum number of bonds that it can in the circumstances, and this is a perfectly valid structure. Energy is released whenever a covalent bond is formed. Because energy is being lost from the system, it becomes more stable after every covalent bond is made. It follows, therefore, that an atom will tend to make as many covalent bonds as possible. In the case of boron in BF3, three bonds is the maximum possible because boron only has 3 electrons to share.
Note: You might perhaps wonder why boron doesn't form ionic bonds with fluorine instead. Boron doesn't form ions because the total energy needed to remove three

electrons to form a B3+ ion is simply too great to be recoverable when attractions are set up between the boron and fluoride ions.

Phosphorus(V) chloride, PCl5 In the case of phosphorus 5 covalent bonds are possible - as in PCl5. Phosphorus forms two chlorides - PCl3 and PCl5. When phosphorus burns in chlorine both are formed - the majority product depending on how much chlorine is available. We've already looked at the structure of PCl3. The diagram of PCl5 (like the previous diagram of PCl 3) shows only the outer electrons.

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

A more sophisticated view of covalent bonding


The bonding in methane, CH4
Warning! If you aren't happy with describing electron arrangements in s and p notation, and with the shapes of s and p orbitals, you need to read about orbitals before you go on. Use the BACK button on your

browser to return quickly to this point.

What is wrong with the dots-and-crosses picture of bonding in methane? We are starting with methane because it is the simplest case which illustrates the sort of processes involved. You will remember that the dots-and-crossed picture of methane looks like this.

There is a serious mis-match between this structure and the modern electronic structure of carbon, 1s22s22px12py1. The modern structure shows that there are only 2 unpaired electrons for hydrogens to share with, instead of the 4 which the simple view requires. You can see this more readily using the electrons-in-boxes notation. Only the 2-level electrons are shown. The 1s2electrons are too deep inside the atom to be involved in bonding. The only electrons directly available for sharing are the 2p electrons. Why then isn't methane CH2? Promotion of an electron When bonds are formed, energy is released and the system becomes more stable. If carbon forms 4 bonds rather than 2, twice as much energy is released and so the resulting molecule becomes even more stable. There is only a small energy gap between the 2s and 2p orbitals, and so it pays the carbon to provide a small amount of energy to promote an electron from the 2s to the empty 2p to give 4 unpaired electrons. The extra energy released when the bonds form more than compensates for the initial input.

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

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

Now that we've got 4 unpaired electrons ready for bonding, another problem arises. In methane all the carbon-hydrogen bonds are identical, but our electrons are in two different kinds of orbitals. You aren't going to get four identical bonds unless you start from four identical orbitals. Hybridisation The electrons rearrange themselves again in a process called hybridisation. This reorganises the electrons into four identical hybrid orbitals called sp3hybrids (because they are made from one s orbital and three p orbitals). You should read "sp3" as "s p three" - not as "s p cubed". sp3 hybrid orbitals look a bit like half a p orbital, and they arrange themselves in space so that they are as far apart as possible. You can picture the nucleus as being at the centre of a tetrahedron (a triangularly based pyramid) with the orbitals pointing to the corners. For clarity, the nucleus is drawn far larger than it really is.

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

Four molecular orbitals are formed, looking rather like the original sp3 hybrids, but with a hydrogen nucleus embedded in each lobe. Each orbital holds the 2 electrons that we've previously drawn as a dot and a cross. The principles involved - promotion of electrons if necessary, then hybridisation, followed by the formation of molecular orbitals - can be applied to any covalently-bound molecule.
Note: You will find this bit on methane repeated in the organic section of this site. That article on methane goes on to look at the formation of carbon-carbon single bonds in ethane.

The bonding in the phosphorus chlorides, PCl3 and PCl5 What's wrong with the simple view of PCl3? This diagram only shows the outer (bonding) electrons.

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

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

Each of the 3 chlorines then forms a covalent bond by merging the atomic orbital containing its unpaired electron with one of the phosphorus unpaired electrons to make 3 molecular orbitals. You might wonder whether all this is worth the bother! Probably not! It is worth it with PCl5, though. What's wrong with the simple view of PCl5? You will remember that the dots-and-crosses picture of PCl 5 looks awkward because the phosphorus doesn't end up with a noble gas structure. This diagram also shows only the outer electrons.

In this case, a more modern view makes things look better by abandoning any pretence of worrying about noble gas structures. If the phosphorus is going to form PCl 5 it has first to generate 5 unpaired electrons. It does this by promoting one of the electrons in the 3s orbital to the next available higher energy orbital. Which higher energy orbital? It uses one of the 3d orbitals. You might have expected it to use the 4s orbital because this is the orbital that fills before the 3d when atoms are being built from scratch. Not so! Apart from when you are building the atoms in the first place, the 3d always counts as the lower energy orbital.

This leaves the phosphorus with this arrangement of its electrons:

The 3-level electrons now rearrange (hybridise) themselves to give 5 hybrid orbitals, all of equal energy. They would be called sp 3d hybrids because that's what they are made from.

The electrons in each of these orbitals would then share space with electrons from five chlorines to make five new molecular orbitals and hence five covalent bonds. Why does phosphorus form these extra two bonds? It puts in an amount of energy to promote an electron, which is more than paid back when the new bonds form. Put simply, it is energetically profitable for the phosphorus to form the extra bonds. The advantage of thinking of it in this way is that it completely ignores the question of whether you've got a noble gas structure, and so you don't worry about it. A non-existent compound - NCl5 Nitrogen is in the same Group of the Periodic Table as phosphorus, and you might expect it to form a similar range of compounds. In fact, it doesn't. For example, the compound NCl 3exists, but there is no such thing as NCl 5. Nitrogen is 1s22s22px12py12pz1. The reason that NCl 5 doesn't exist is that in order to form five bonds, the nitrogen would have to promote one of its 2s electrons. The problem is that there aren't any 2d orbitals to promote an electron into - and the energy gap to the next level (the 3s) is far too great. In this case, then, the energy released when the extra bonds are made isn't enough to compensate for the energy needed to promote an electron - and so that promotion doesn't happen. Atoms will form as many bonds as possible provided it is

energetically profitable.

COVALENT BONDING - DOUBLE BONDS


This page explains how double covalent bonds arise. It starts with a simple picture of double covalent bonding, and then takes a more sophisticated view of the bonding in ethene.
Warning! This page assumes that you have already read the page on single covalent bonds. If you have come straight to this page via a search engine follow this link before you go on.

A simple view of double covalent bonds


A double covalent bond is where two pairs of electrons are shared between the atoms rather than just one pair. Some simple molecules containing double bonds Oxygen, O2 Two oxygen atoms can both achieve stable structures by sharing two pairs of electrons as in the diagram.

The double bond is shown conventionally by two lines joining the atoms. Each line represents one pair of shared electrons. Carbon dioxide, CO2

Ethene, C2H4 Ethene has a double bond between the two carbon atoms.

A more sophisticated view of the bonding in ethene


It is important to explore the bonding in ethene in more detail because it has a direct impact on its chemistry. Unless you have some understanding of the true nature of the double bond, you can't really understand the way that ethene behaves. An orbital view of the bonding in ethene Ethene is built from hydrogen atoms (1s1) and carbon atoms (1s22s22px12py1). Promotion of an electron The carbon atom doesn't have enough unpaired electrons to form the required number of bonds, so it needs to promote one of the 2s2 pair into the empty 2pz orbital. This is exactly the same as happens whenever carbon forms bonds - whatever else it ends up joined to. The carbon atom is now in an excited state.

Hybridisation In the case of ethene, there is a difference from methane because each carbon is only joining to three other atoms rather than four. When the carbon atoms hybridise their outer orbitals before forming bonds, this time they only hybridise three of the orbitals rather than all four. They use the 2s electron and two of the 2p electrons, but leave the other 2p electron unchanged.

Note: You might wonder why it chooses to hybridise these three orbitals rather than just use

the three p orbitals which already have the same energy. It's because it uses the orbitals with the lowest energy first.

The new orbitals formed are calledsp2 hybrids, because they are made by an s orbital and two p orbitals reorganising themselves. sp2 orbitals look rather like sp3 orbitals that we discussed in the bonding in methane in the page on single bonds, except that they are shorter and fatter. The three sp2 hybrid orbitals arrange themselves as far apart as possible which is at 120 to each other in a plane. The remaining p orbital is at right angles to them. The two carbon atoms and four hydrogen atoms would look like this before they joined together:

The various atomic orbitals which are pointing towards each other now merge to give molecular orbitals, each containing a bonding pair of electrons. Molecular orbitals made by end-to-end overlap of atomic orbitals are called sigma bonds.

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

For clarity, the sigma bonds are shown using lines - each line representing one pair of shared electrons. The various sorts of line show the directions the bonds point in. An ordinary line represents a bond in the plane of the screen (or the paper if you've printed it), a broken line is a bond going back away from you, and a wedge shows a bond coming out towards you.
Note: The really interesting bond in ethene is the pi bond. In almost all cases where you will draw the structure of ethene, the sigma bonds will be shown as lines.

Be clear about what a pi bond is. It is a region of space in which you can find the two electrons which make up the bond. Those two electrons can live anywhere within that space. It would be quite misleading to think of one living in the top and the other in the bottom.
Taking chemistry further: This is a good example of the curious behaviour of electrons. How do the electrons get from one half of the pi bond to the other if they are never found in between? It's an unanswerable question if you think of electrons as particles. If you want to follow this up, you will have to read some fairly highpowered stuff on the wave nature of electrons.

Even if your syllabus doesn't expect you to know how a pi bond is formed, it will expect you to know that it exists. The pi bond dominates the chemistry of ethene. It is very vulnerable to attack - a very negative region of space above and below the plane of the molecule. It is also somewhat distant from the control of the nuclei and so is a weaker bond than the sigma bond joining the two carbons.
Important! Check your syllabus! Find out whether you actually need to know how a pi bond is formed. Don't forget to look under ethene as well as in the bonding section of your syllabus. If you don't need to know it, there's no point in learning it! You will, however, need to know that a pi bond exists - that the two bonds between the carbon atoms in ethene aren't both the same. If you are working to a UK-based syllabus, but haven't got a copy of it, find out how to download one.

All double bonds (whatever atoms they might be joining) will consist of a sigma bond and a pi bond. This orbital view of the double bond is only really important at this level with regard to organic compounds. If you want to read more about this, follow the first link below which leads you to the menu for a section specifically on organic bonding. You will find the description of ethene repeated, but will also find

information about the bonding in benzene and in the carbon-oxygen double bond.

CO-ORDINATE (DATIVE COVALENT) BONDING


This page explains what co-ordinate (also called dative covalent) bonding is. You need to have a reasonable understanding of simple covalent bonding before you start.
Important! If you are uncertain about covalent bonding follow this link before you go on with this page.

Co-ordinate (dative covalent) bonding


A covalent bond is formed by two atoms sharing a pair of electrons. The atoms are held together because the electron pair is attracted by both of the nuclei. In the formation of a simple covalent bond, each atom supplies one electron to the bond - but that doesn't have to be the case. A coordinate bond (also called a dative covalent bond) is a covalent bond (a shared pair of electrons) in which both electrons come from the same atom. For the rest of this page, we shall use the term co-ordinate bond but if you prefer to call it a dative covalent bond, that's not a problem! The reaction between ammonia and hydrogen chloride If these colourless gases are allowed to mix, a thick white smoke of solid ammonium chloride is formed. Ammonium ions, NH4+, are formed by the transfer of a hydrogen ion from the hydrogen chloride to the lone pair of electrons on the ammonia molecule.

When the ammonium ion, NH4+, is formed, the fourth hydrogen is attached by a dative covalent bond, because only the hydrogen's nucleus is transferred from the chlorine to the nitrogen. The hydrogen's electron is left behind on the chlorine to form a negative chloride ion. Once the ammonium ion has been formed it is impossible to tell any difference between the dative covalent and the ordinary covalent bonds. Although the electrons are shown differently in the diagram, there is no difference between them in reality. Representing co-ordinate bonds In simple diagrams, a co-ordinate bond is shown by an arrow. The arrow points from the atom donating the lone pair to the atom accepting it.

Dissolving hydrogen chloride in water to make hydrochloric acid Something similar happens. A hydrogen ion (H+) is transferred from the chlorine to one of the lone pairs on the oxygen atom.

The H3O+ ion is variously called the hydroxonium ion, the hydronium ion or the oxonium ion. In an introductory chemistry course (such as GCSE), whenever you

have talked about hydrogen ions (for example in acids), you have actually been talking about the hydroxonium ion. A raw hydrogen ion is simply a proton, and is far too reactive to exist on its own in a test tube. If you write the hydrogen ion as H+(aq), the "(aq)" represents the water molecule that the hydrogen ion is attached to. When it reacts with something (an alkali, for example), the hydrogen ion simply becomes detached from the water molecule again. Note that once the co-ordinate bond has been set up, all the hydrogens attached to the oxygen are exactly equivalent. When a hydrogen ion breaks away again, it could be any of the three. The reaction between ammonia and boron trifluoride, BF 3 If you have recently read the page on covalent bonding, you may remember boron trifluoride as a compound which doesn't have a noble gas structure around the boron atom. The boron only has 3 pairs of electrons in its bonding level, whereas there would be room for 4 pairs. BF3 is described as being electron deficient. The lone pair on the nitrogen of an ammonia molecule can be used to overcome that deficiency, and a compound is formed involving a co-ordinate bond.

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

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

The structure of aluminium chloride Aluminium chloride sublimes (turns straight from a solid to a gas) at about 180C. If it simply contained ions it would have a very high melting and boiling point because of the strong attractions between the positive and negative ions. The implication is that it when it sublimes at this relatively low temperature, it must be covalent. The dots-and-crosses diagram shows only the outer electrons. AlCl3, like BF3, is electron deficient. There is likely to be a similarity, because aluminium and boron are in the same group of the Periodic Table, as are fluorine and chlorine.

Measurements of the relative formula mass of aluminium chloride show that its formula in the vapour at the sublimation temperature is not AlCl3, but Al2Cl6. It exists as a dimer (two molecules joined together). The bonding between the two molecules is co-ordinate, using lone pairs on the chlorine atoms. Each chlorine atom has 3 lone pairs, but only the two important ones are shown in the line diagram.

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

Energy is released when the two co-ordinate bonds are formed, and so the dimer is more stable than two separate AlCl 3molecules.
Note: Aluminium chloride is complicated because of the way it

keeps changing its bonding as the temperature increases. If you are interested in exploring this in more detail, you could have a look at the page about the Period 3 chlorides. It isn't particularly relevant to the present page, though. If you choose to follow this link, use the BACK button on your browser to return quickly to this page later.

The bonding in hydrated metal ions Water molecules are strongly attracted to ions in solution - the water molecules clustering around the positive or negative ions. In many cases, the attractions are so great that formal bonds are made, and this is true of almost all positive metal ions. Ions with water molecules attached are described as hydrated ions. Although aluminium chloride is covalent, when it dissolves in water, ions are produced. Six water molecules bond to the aluminium to give an ion with the formula Al(H2O)63+. It's called the hexaaquaaluminium ion - which translates as six ("hexa") water molecules ("aqua") wrapped around an aluminium ion. The bonding in this (and the similar ions formed by the great majority of other metals) is co-ordinate (dative covalent) using lone pairs on the water molecules. Aluminium is 1s22s22p63s23px1. When it forms an Al3+ ion it loses the 3-level electrons to leave 1s22s22p6. That means that all the 3-level orbitals are now empty. The aluminium re-organises (hybridises) six of these (the 3s, three 3p, and two 3d) to produce six new orbitals all with the same energy. These six hybrid orbitals accept lone pairs from six water molecules. You might wonder why it chooses to use six orbitals rather than four or eight or whatever. Six is the maximum number of water molecules it is possible to fit around an aluminium ion (and most other metal ions). By making the maximum number of bonds, it releases most energy and so becomes most energetically stable.

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

Because of the movement of electrons towards the centre of the ion, the 3+ charge is no longer located entirely on the aluminium, but is now spread over the whole of the ion.
Note: Dotted arrows represent lone pairs coming from water molecules behind the plane of the screen or paper. Wedge shaped arrows represent bonds from water molecules in front of the plane of the screen or paper.

Two more molecules


Note: It looks as if only one current UK syllabus wants these two. Check yours! If you haven't got a copy of your syllabus, follow this link to find out how to get one.

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

Nitric acid, HNO3 In this case, one of the oxygen atoms can be thought of as attaching to the nitrogen via a co-ordinate bond using the lone pair on the nitrogen atom.

In fact this structure is misleading because it suggests that the two oxygen atoms on the right-hand side of the diagram are joined to the nitrogen in different ways. Both bonds are actually identical in length and strength, and so the arrangement of the electrons must be identical. There is no way of showing this using a dots-andcrosses picture. The bonding involves delocalisation.
If you are interested: The bonding is rather similar to the bonding in the ethanoate ion (although without the negative charge). You will find thisdescribed on a page about theacidity of organic acids. Important: Don't read the next bit until you have read and understood the pages I have pointed you to. The rest of this page explains a tricky example which cropped up in a CIE exam question. You won't understand it until you are confident about the more straightforward examples.

At the time of writing (June 2010), they had already asked quite an awkward question, which you almost certainly wouldn't have come across as a part of your course. You were asked for the shape of the chlorine dioxide molecule, ClO 2, and told that the chlorine-oxygen bonds were both double.

The temptation is to think of this as being like carbon dioxide - linear. In fact, if you jump to that conclusion, you would be totally wrong. Incidentally, that is why it is important to have read about the sulfur dioxide case, even though it isn't specifically mentioned by the syllabus. Let's work through it for chlorine dioxide . . . Chlorine is in Group 7, and so has 7 electrons in its outer level. It is forming a total of 4 bonds to oxygens (2 double bonds), and so that adds another 4 electrons to the outer level - making a total of 11. It is a neutral molecule, not an ion, and so that's the final total of electrons. 11 electrons will be in five-and-a-half pairs. What that means, of course, is that there are 5 pairs, and an odd single electron. If you haven't come across molecules with unpaired electrons before, that might worry you but such things are perfectly possible. 4 of the 5 pairs will be used to form the 2 double bonds. So, there will be 2 double bond units, a lone pair and a single electron. These will arrange themselves in an approximately tetrahedral fashion to minimise the repulsions. When you think about the shape that the atoms take up, the molecule will be bent (or V-shaped or, as CIE tend to call it, "non-linear"). This question was only worth 2 marks - 1 for the shape, and 1 for the explanation. On occasions, CIE certainly make you work for your marks!
Warning: This method won't work without some modification for many ions containing metals, and no simple method gives reliable results where the central atom is a transition metal. The method will, however, cope with all the substances that you are likely to meet in this section of the syllabus. When you deal with transition metal chemistry, you will be expected to know the shapes of some ions formed by transition metals, but not to work them out. At that point, learn the ones your syllabus wants you to know. It is important to know exactly which molecules and ions your syllabus expects you to be able to work out the shapes for in this part of the syllabus. You should also check past exam papers. If you are working to a UK-based syllabus for 16 - 18 year olds, and haven't got copies of your syllabus and past papers follow this link to find out how to get them.

First you need to work out how many electrons there are around the central atom: Write down the number of electrons in the outer level of the central atom. That will be the same as the Periodic Table group number, except in the case of the noble gases which form compounds, when it will be 8. Add one electron for each bond being formed. (This allows for the electrons coming from the other atoms.) Allow for any ion charge. For example, if the ion has a 1-

charge, add one more electron. For a 1+ charge, deduct an electron. Now work out how many bonding pairs and lone pairs of electrons there are: Divide by 2 to find the total number of electron pairs around the central atom. Work out how many of these are bonding pairs, and how many are lone pairs. You know how many bonding pairs there are because you know how many other atoms are joined to the central atom (assuming that only single bonds are formed). For example, if you have 4 pairs of electrons but only 3 bonds, there must be 1 lone pair as well as the 3 bonding pairs. Finally, you have to use this information to work out the shape: Arrange these electron pairs in space to minimise repulsions. How this is done will become clear in the examples which follow.
Don't panic! This is all much easier to do in practice than it is to describe in a long list like this one!

Two electron pairs around the central atom The only simple case of this is beryllium chloride, BeCl2. The electronegativity difference between beryllium and chlorine isn't enough to allow the formation of ions. Beryllium has 2 outer electrons because it is in group 2. It forms bonds to two chlorines, each of which adds another electron to the outer level of the beryllium. There is no ionic charge to worry about, so there are 4 electrons altogether - 2 pairs. It is forming 2 bonds so there are no lone pairs. The two bonding pairs arrange themselves at 180 to each other, because that's as far apart as they can get. The molecule is described as beinglinear.

Three electron pairs around the central atom The simple cases of this would be BF 3 or BCl3. Boron is in group 3, so starts off with 3 electrons. It is forming 3 bonds, adding another 3 electrons. There is no charge, so the total is 6 electrons - in 3 pairs. Because it is forming 3 bonds there can be no lone pairs. The 3 pairs

arrange themselves as far apart as possible. They all lie in one plane at 120 to each other. The arrangement is calledtrigonal planar.

In the diagram, the other electrons on the fluorines have been left out because they are irrelevant. Four electron pairs around the central atom There are lots of examples of this. The simplest is methane, CH4.
Note: Elsewhere on the site, you will find the shape of methane worked out in detail using modern bonding theory. Here we are doing it the quick and easy way! If you are interested in the bonding in methane you can find it in the organic section by following this link, or in a page oncovalent bonding by following this one.

Carbon is in group 4, and so has 4 outer electrons. It is forming 4 bonds to hydrogens, adding another 4 electrons - 8 altogether, in 4 pairs. Because it is forming 4 bonds, these must all be bonding pairs. Four electron pairs arrange themselves in space in what is called a tetrahedral arrangement. A tetrahedron is a regular triangularlybased pyramid. The carbon atom would be at the centre and the hydrogens at the four corners. All the bond angles are 109.5.

Note: It is important that you understand the use of various sorts of line to show the 3-dimensional arrangement of the bonds. In diagrams of this sort, an ordinary line represents a bond in the plane of the screen or paper. A dotted line shows a bond

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

Other examples with four electron pairs around the central atom Ammonia, NH3 Nitrogen is in group 5 and so has 5 outer electrons. Each of the 3 hydrogens is adding another electron to the nitrogen's outer level, making a total of 8 electrons in 4 pairs. Because the nitrogen is only forming 3 bonds, one of the pairs must be a lone pair. The electron pairs arrange themselves in a tetrahedral fashion as in methane.

In this case, an additional factor comes into play. Lone pairs are in orbitals that are shorter and rounder than the orbitals that the bonding pairs occupy. Because of this, there is more repulsion between a lone pair and a bonding pair than there is between two bonding pairs. That forces the bonding pairs together slightly - reducing the bond angle from 109.5 to 107. It's not much, but the examiners will expect you to know it. Remember this: Greatest repulsion lone pair - lone pair lone pair - bond pair Least repulsion bond pair - bond pair

Be very careful when you describe the shape of ammonia. Although the electron pair arrangement is tetrahedral, when you describe the

shape, you only take notice of the atoms. Ammonia ispyramidal like a pyramid with the three hydrogens at the base and the nitrogen at the top. Water, H2O

Following the same logic as before, you will find that the oxygen has four pairs of electrons, two of which are lone pairs. These will again take up a tetrahedral arrangement. This time the bond angle closes slightly more to 104, because of the repulsion of the two lone pairs. The shape isn't described as tetrahedral, because we only "see" the oxygen and the hydrogens - not the lone pairs. Water is described as bent or V-shaped. The ammonium ion, NH4+ The nitrogen has 5 outer electrons, plus another 4 from the four hydrogens - making a total of 9. But take care! This is a positive ion. It has a 1+ charge because it has lost 1 electron. That leaves a total of 8 electrons in the outer level of the nitrogen. There are therefore 4 pairs, all of which are bonding because of the four hydrogens. The ammonium ion has exactly the same shape as methane, because it has exactly the same electronic arrangement. NH 4+ is tetrahedral.
Note: To simplify diagrams, bonding electrons won't be shown from now on. Each line, of course, represents a bonding pair. It is essential, however, to draw lone pairs.

Methane and the ammonium ion are said to be isoelectronic. Two species (atoms, molecules or ions) are isoelectronic if they have exactly the same number and arrangement of electrons (including

the distinction between bonding pairs and lone pairs). The hydroxonium ion, H3O+ Oxygen is in group 6 - so has 6 outer electrons. Add 1 for each hydrogen, giving 9. Take one off for the +1 ion, leaving 8. This gives 4 pairs, 3 of which are bond pairs. The hydroxonium ion is isoelectronic with ammonia, and has an identical shape - pyramidal.

Five electron pairs around the central atom A simple example: phosphorus(V) fluoride, PF5 (The argument for phosphorus(V) chloride, PCl 5, would be identical.) Phosphorus (in group 5) contributes 5 electrons, and the five fluorines 5 more, giving 10 electrons in 5 pairs around the central atom. Since the phosphorus is forming five bonds, there can't be any lone pairs. The 5 electron pairs take up a shape described as a trigonal bipyramid - three of the fluorines are in a plane at 120 to each other; the other two are at right angles to this plane. The trigonal bipyramid therefore has two different bond angles - 120 and 90.

A tricky example, ClF3 Chlorine is in group 7 and so has 7 outer electrons. The three fluorines contribute one electron each, making a total of 10 - in 5 pairs. The chlorine is forming three bonds - leaving you with 3 bonding pairs and 2 lone pairs, which will arrange themselves into a trigonal bipyramid. But don't jump to conclusions. There are actually three different ways in which you could arrange 3 bonding pairs and 2 lone pairs into a trigonal bipyramid. The right arrangement will be the one with the minimum amount of repulsion - and you can't decide that without first drawing all the possibilities.

These are the only possible arrangements. Anything else you might think of is simply one of these rotated in space. We need to work out which of these arrangements has the minimum amount of repulsion between the various electron pairs. A new rule applies in cases like this: If you have more than four electron pairs arranged around the central atom, you can ignore repulsions at angles of greater than 90. One of these structures has a fairly obvious large amount of repulsion.

In this diagram, two lone pairs are at 90 to each other, whereas in the other two cases they are at more than 90, and so their repulsions can be ignored. ClF3 certainly won't take up this shape because of the strong lone pair-lone pair repulsion. To choose between the other two, you need to count up each sort of repulsion. In the next structure, each lone pair is at 90 to 3 bond pairs, and so each lone pair is responsible for 3 lone pair-bond pair repulsions.

Because of the two lone pairs there are therefore 6 lone pair-bond pair repulsions. And that's all. The bond pairs are at an angle of 120 to each other, and their repulsions can be ignored. Now consider the final structure.

Each lone pair is at 90 to 2 bond pairs - the ones above and below the plane. That makes a total of 4 lone pair-bond pair repulsions compared with 6 of these relatively strong repulsions in the last structure. The other fluorine (the one in the plane) is 120 away, and feels negligible repulsion from the lone pairs. The bond to the fluorine in the plane is at 90 to the bonds above and below the plane, so there are a total of 2 bond pair-bond pair repulsions. The structure with the minimum amount of repulsion is therefore this last one, because bond pair-bond pair repulsion is less than lone pair-bond pair repulsion. ClF3 is described as T-shaped.
Warning! If your syllabus expects you to discuss examples with more than 4 pairs of electrons around the central atom, check past exam papers to see if nasty questions like this one involving ClF3 ever come up. If so, don't leave this example until you are sure that you understand it. It is by far the most complicated one on this page.

Six electron pairs around the central atom A simple example: SF6 6 electrons in the outer level of the sulphur, plus 1 each from the six fluorines, makes a total of 12 - in 6 pairs. Because the sulphur is forming 6 bonds, these are all bond pairs. They arrange themselves entirely at 90, in a shape described as octahedral.

Two slightly more difficult examples XeF4 Xenon forms a range of compounds, mainly with fluorine or oxygen, and this is a typical one. Xenon has 8 outer electrons, plus 1 from each fluorine - making 12 altogether, in 6 pairs. There will be 4 bonding pairs (because of the four fluorines) and 2 lone pairs.

There are two possible structures, but in one of them the lone pairs would be at 90. Instead, they go opposite each other. XeF4 is described as square planar. ClF4Chlorine is in group 7 and so has 7 outer electrons. Plus the 4 from the four fluorines. Plus one because it has a 1- charge. That gives a total of 12 electrons in 6 pairs - 4 bond pairs and 2 lone pairs. The shape will be identical with that of XeF 4.

Learning outcomes 3(d) and 3(e) These statements cover the formation of covalent bonds using some simple orbital theory, including the the cases of ethane, ethene and benzene. Before you go on, you should find and read the statements in your copy of the syllabus. You should notice that benzene is mentioned in bold type. That means that it will only be asked about in the second year of a two year course - not at AS. First, you should read the page which looks at bonding in methane and ethane. This page (and the next two that you will find below) come from the organic chemistry part of Chemguide. If you have already read the whole of the covalent bonding page referred to in learning outcome 3(b), you will recognise the bit about methane. Methane isn't specifically mentioned by the syllabus, but you must be sure that you understand that before you go on to the ethane at the bottom of the page. The concept of hybridisation also isn't specifically mentioned in the syllabus. In fact, most A level syllabuses don't expect you to know about it. However, it isn't really very difficult, and if you take a small amount of time in understanding it, it makes the whole covalent bonding topic much more satisfying - certainly as far as organic chemistry is concerned. Now read the page which about bonding in ethene. By the time you have finished this page, it is really important that you understand the difference between a sigma bond (formed by end-to-end overlap between atomic orbitals) and a pi bond (formed by sideways overlap).

For benzene, you have two options. If I was teaching it, I would want to include the benzene case to follow logically on from ethene, even though benzene won't be examined at AS. Alternatively, you could leave this for now, and come back to it again later when it is more directly relevant to the chemistry of benzene. The danger with that is that you could end up missing it out. This statement is repeated at the beginning of the organic section of the syllabus, but not specifically when the chemistry of benzene is introduced. I suggest, to be on the safe side, that you read this now and re-read it later on. The page you want at this stage is bonding in benzene. Towards the bottom of that page, you will find a link to another page about the Kekul structure for benzene. If you are coming to the benzene structure early on in the course, it would pay you to have a brief look at the first few paragraphs on that page, just looking at what the Kekul structure is, and the paragraph about the shape.

BONDING IN METHANE AND ETHANE


Warning! If you aren't happy with describing electron arrangements in s and p notation, and with the shapes of s and p orbitals, you really should read about orbitals. Use the BACK button on your browser to return quickly to this point.

Methane, CH4
The simple view of the bonding in methane You will be familiar with drawing methane using dots and crosses diagrams, but it is worth looking at its structure a bit more closely. There is a serious mis-match between this structure and the modern electronic structure of carbon, 1s22s22px12py1. The modern structure shows that there are only 2 unpaired electrons for hydrogens to share with, instead of the 4 which the simple view requires. You can see this more readily using the electrons-in-boxes notation. Only the 2level electrons are shown. The 1s2electrons are too deep inside the atom to be involved in bonding. The only

electrons directly available for sharing are the 2p electrons. Why then isn't methane CH2? Promotion of an electron When bonds are formed, energy is released and the system becomes more stable. If carbon forms 4 bonds rather than 2, twice as much energy is released and so the resulting molecule becomes even more stable. There is only a small energy gap between the 2s and 2p orbitals, and so it pays the carbon to provide a small amount of energy to promote an electron from the 2s to the empty 2p to give 4 unpaired electrons. The extra energy released when the bonds form more than compensates for the initial input.

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


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

Now that we've got 4 unpaired electrons ready for bonding, another problem arises. In methane all the carbon-hydrogen bonds are identical, but our electrons are in two different kinds of orbitals. You aren't going to get four identical bonds unless you start from four identical orbitals. Hybridisation The electrons rearrange themselves again in a process called hybridisation. This reorganises the electrons into four identical hybrid orbitals called sp3hybrids (because they are made from one s orbital and three p orbitals). You should read "sp3" as "s p three" - not as "s p cubed". sp3 hybrid orbitals look a bit like half a p orbital, and they arrange themselves in space so that they are as far apart as possible. You can picture the nucleus as being at the centre of a tetrahedron (a triangularly based pyramid) with the orbitals pointing to the corners. For clarity, the nucleus is drawn far larger than it really is.

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

Four molecular orbitals are formed, looking rather like the

original sp3 hybrids, but with a hydrogen nucleus embedded in each lobe. Each orbital holds the 2 electrons that we've previously drawn as a dot and a cross. The principles involved - promotion of electrons if necessary, then hybridisation, followed by the formation of molecular orbitals - can be applied to any covalently-bound molecule. The shape of methane When sp3 orbitals are formed, they arrange themselves so that they are as far apart as possible. That is a tetrahedral arrangement, with an angle of 109.5. Nothing changes in terms of the shape when the hydrogen atoms combine with the carbon, and so the methane molecule is also tetrahedral with 109.5 bond angles.

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

The hydrogens bond with the two carbons to produce molecular orbitals just as they did with methane. The two carbon atoms bond by merging their remaining sp 3 hybrid orbitals end-to-end to make a new molecular orbital. The bond formed by this endto-end overlap is called a sigma bond. The bonds between the carbons and hydrogens are also sigma bonds.

In any sigma bond, the most likely place to find the pair of electrons is on a line between the two nuclei. The shape of ethane around each carbon atom The shape is again determined by the way the sp 3 orbitals are arranged around each carbon atom. That is a tetrahedral arrangement, with an angle of 109.5. When the ethane molecule is put together, the arrangement around each carbon atom is again tetrahedral with approximately 109.5 bond angles. Why only "approximately"? This time, each carbon atoms doesn't have four identical things attached. There will be a small amount of distortion because of the attachment of 3 hydrogens and 1 carbon, rather than 4 hydrogens. Free rotation about the carbon-carbon single bond The two ends of this molecule can spin quite freely about the sigma bond so that there are, in a sense, an infinite number of possibilities for the shape of an ethane molecule. Some possible shapes are:

In each case, the left hand CH3 group has been kept in a constant position so that you can see the effect of spinning the right hand one. Other alkanes All other alkanes will be bonded in the same way: The carbon atoms will each promote an electron and then hybridise to give sp3 hybrid orbitals. The carbon atoms will join to each other by forming sigma bonds by the end-to-end overlap of their sp3 hybrid orbitals. Hydrogen atoms will join on wherever they are needed by overlapping their 1s1 orbitals with sp3 hybrid orbitals on the carbon atoms. BONDING IN ETHENE

Important! You will find this much easier to understand if you first read the article about the bonding in methane. You may also find it useful to read the article on orbitals if you aren't sure about simple orbital theory.

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

There is only a small energy gap between the 2s and 2p orbitals, and an electron is promoted from the 2s to the empty 2p to give 4 unpaired electrons. The extra energy released when these electrons are used for bonding more than compensates for the initial input. The carbon atom is now said to be in an excited state.
Note: If you haven't read about bonding in methane, follow this link before you go any further. Use the BACK button on your browser to come back here when you have finished. It is important that you have first met the idea of hybridisation in the more simple methane case.

Hybridisation In the case of ethene, there is a difference from, say, methane or ethane, because each carbon is only joining to three other atoms rather than four. When the carbon atoms hybridise their outer orbitals before forming bonds, this time they only hybridise three of the orbitals rather than all four. They use the 2s electron and two of the 2p electrons, but leave the other 2p electron unchanged.

The new orbitals formed are calledsp2 hybrids, because they are made by an s orbital and two p orbitals reorganising themselves. sp2 orbitals look rather like sp3 orbitals that you have already come across in the bonding in methane, except that they are shorter and fatter. The three sp2 hybrid orbitals arrange themselves as far apart as possible which is at 120 to each other in a plane. The remaining p orbital is at right angles to them. The two carbon atoms and four hydrogen atoms would look like this before they joined together:

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

The p orbitals on each carbon aren't pointing towards each other, and so we'll leave those for a moment. In the diagram, the black dots represent the nuclei of the atoms. Notice that the p orbitals are so close that they are overlapping sideways. This sideways overlap also creates a

molecular orbital, but of a different kind. In this one the electrons aren't held on the line between the two nuclei, but above and below the plane of the molecule. A bond formed in this way is called a pi bond.

For clarity, the sigma bonds are shown using lines - each line representing one pair of shared electrons. The various sorts of line show the directions the bonds point in. An ordinary line represents a bond in the plane of the screen (or the paper if you've printed it), a broken line is a bond going back away from you, and a wedge shows a bond coming out towards you.
Note: The really interesting bond in ethene is the pi bond. In almost all cases where you will draw the structure of ethene, the sigma bonds will be shown as lines.

Be clear about what a pi bond is. It is a region of space in which you can find the two electrons which make up the bond. Those two electrons can live anywhere within that space. It would be quite misleading to think of one living in the top and the other in the bottom.
Taking chemistry further: This is another example of the curious behaviour of electrons. How do the electrons get from one half of the pi bond to the other if they are never found in between? It's an unanswerable question if you think of electrons as particles.

Even if your syllabus doesn't expect you to know how a pi bond is formed, it will expect you to know that it exists. The pi bond dominates the chemistry of ethene. It is very vulnerable to attack - a very negative region of space above and below the plane of the molecule. It is also somewhat distant from the control of the nuclei and so is a weaker bond than the sigma bond joining the two carbons.
Important! Check your syllabus! Find out whether you actually need to know how a pi bond is formed. Don't forget to look in the bonding section of your syllabus as well as under ethene. If you don't need to know it, there's no point in learning it! You will, however, need to know that a pi bond exists - that the two bonds between

the carbon atoms in ethene aren't both the same. If you are working to a UK-based syllabus for 16 - 18 year olds, and haven't got a copy of your syllabus, find out how to download one

All double bonds (whatever atoms they might be joining) will consist of a sigma bond and a pi bond. The shape of ethene The shape of ethene is controlled by the arrangement of the sp2orbitals. Notice two things about them: They all lie in the same plane, with the other p orbital at right angles to it. When the bonds are made, all of the sigma bonds in the molecule must also lie in the same plane. Any twist in the molecule would mean that the p orbitals wouldn't be parallel and touching any more, and you would be breaking the pi bond. There is no free rotation about a carbon-carbon double bond. Ethene is a planar molecule. 2 The sp orbitals are at 120 to each other. When the molecule is constructed, the bond angles will also be 120. (That's approximate! There will be a slight distortion because you are joining 2 hydrogens and a carbon atom to each carbon, rather than 3 identical groups.)

Learning outcome 3(f) This statement allows the examiners to ask you about the shapes and bond angles of other molecules similar to the ones already mentioned previously in this section. Before you go on, you should find and read the statement in your copy of the syllabus. There is no short-cut to this! You need to understand how to work out shapes of molecules and their bond angles, using both inorganic and organic examples, as in learning outcomes 3(c) and 3(e). There is no substitute for understanding this. If you try to learn examples parrot-fashion, you won't be able to cope with unfamiliar cases. And CIE like to use unfamiliar examples, because it is a good test of candidates' ability. The best way of testing yourself is to look for examples from past question papers, and make sure that you can do them by looking at the mark schemes and examiner's reports. However, don't then waste time trying to learn these extra examples because, next time, the CIE examiners will most probably have come up with something

entirely different! As I said just now, there is no real alternative to understanding this. Learning outcome 3(g)
This statement deals with hydrogen bonding. Before you go on, you should find and read the statement in your copy of the syllabus. Electronegativity You can't understand hydrogen bonding, and some of the later statements in this section, unless you first understand the concept of electronegativity. This is very important in chemistry, and to my mind should have its own separate statement at this point in the syllabus. So before you make a start on this learning outcome, you should first spend some time making sure that you fully understand electronegativity - what it means, what causes electronegativity differences, and what effects it has. To do that, first read the page about electronegativity in the bonding section of Chemguide. Ignore the bit below the red warning note. When you are sure that you understand that page, follow the link at the bottom to look at it all again in an organic chemistry context. Read that second page down to, and including, "Bond polarity and inductive effects". Read the rest of the page for interest if you want to, but you will meet this later on when it will have more meaning to you. It is really important that you get this sorted out properly before you go on. You mustn't think "This isn't a syllabus statement, and so I don't need to bother with it." It should be a syllabus statement, and you most definitely do need to bother with it! Hydrogen bonding Read this page on hydrogen bonding. Near the top of the page, you will find a mention of van der Waals forces, and a link to find out more about them. van der Waals forces are coming up soon in learning outcome 3(i), but you won't understand the hydrogen bonding page without having a reasonable idea what van der Waals forces are. There are two ways you could do this: Follow that link, and read the van der Waals page quickly, so that you have some idea what they are about. Then finish off looking at hydrogen bonding, and come back to take a proper look at van der Waals forces later. Alternatively, stop at this point, and do learning outcome 3(i) before you do this one on hydrogen bonding. To my mind, it is actually more logical to look at van der Waals forces first, before you look at hydrogen bonding. Whichever order you do it in, though, you must sort out the electronegativity work first.

ELECTRONEGATIVITY
This page explains what electronegativity is, and how and why it varies around the Periodic Table. It looks at the way that electronegativity differences affect bond type and explains what is meant by polar bonds and polar molecules.

If you are interested in electronegativity in an organic chemistry context, you will find a link at the bottom of this page.

What is electronegativity
Definition Electronegativity is a measure of the tendency of an atom to attract a bonding pair of electrons. The Pauling scale is the most commonly used. Fluorine (the most electronegative element) is assigned a value of 4.0, and values range down to caesium and francium which are the least electronegative at 0.7. What happens if two atoms of equal electronegativity bond together? Consider a bond between two atoms, A and B. Each atom may be forming other bonds as well as the one shown - but these are irrelevant to the argument. If the atoms are equally electronegative, both have the same tendency to attract the bonding pair of electrons, and so it will be found on average half way between the two atoms. To get a bond like this, A and B would usually have to be the same atom. You will find this sort of bond in, for example, H2 or Cl2 molecules.
Note: It's important to realise that this is an average picture. The electrons are actually in a molecular orbital, and are moving around all the time within that orbital.

This sort of bond could be thought of as being a "pure" covalent bond - where the electrons are shared evenly between the two atoms. What happens if B is slightly more electronegative than A? B will attract the electron pair rather more than A does.

That means that the B end of the bond has more than its fair share of electron density and so becomes slightly negative. At the same time, the A end (rather short of

electrons) becomes slightly positive. In the diagram, " " (read as "delta") means "slightly" - so + means "slightly positive". Defining polar bonds This is described as a polar bond. A polar bond is a covalent bond in which there is a separation of charge between one end and the other - in other words in which one end is slightly positive and the other slightly negative. Examples include most covalent bonds. The hydrogenchlorine bond in HCl or the hydrogen-oxygen bonds in water are typical. What happens if B is a lot more electronegative than A? In this case, the electron pair is dragged right over to B's end of the bond. To all intents and purposes, A has lost control of its electron, and B has complete control over both electrons. Ions have been formed.

A "spectrum" of bonds The implication of all this is that there is no clear-cut division between covalent and ionic bonds. In a pure covalent bond, the electrons are held on average exactly half way between the atoms. In a polar bond, the electrons have been dragged slightly towards one end. How far does this dragging have to go before the bond counts as ionic? There is no real answer to that. You normally think of sodium chloride as being a typically ionic solid, but even here the sodium hasn't completely lost control of its electron. Because of the properties of sodium chloride, however, we tend to count it as if it were purely ionic.
Note: Don't worry too much about the exact cut-off point between polar covalent bonds and ionic bonds. At A'level, examples will tend to avoid the grey areas - they will be obviously covalent or obviously ionic. You will, however, be expected to realise that those grey areas exist.

Lithium iodide, on the other hand, would be described as being "ionic with some covalent character". In this case, the pair of electrons hasn't moved entirely over to the iodine end of the bond. Lithium iodide, for example, dissolves in organic solvents like ethanol - not something which ionic substances normally do. Summary No electronegativity difference between two atoms leads to a pure non-polar covalent bond. A small electronegativity difference leads to a polar covalent bond. A large electronegativity difference leads to an ionic bond. Polar bonds and polar molecules In a simple molecule like HCl, if the bond is polar, so also is the whole molecule. What about more complicated molecules? In CCl4, each bond is polar.

Note: Ordinary lines represent bonds in the plane of the screen or paper. Dotted lines represent bonds going away from you into the screen or paper. Wedged lines represent bonds coming out of the screen or paper towards you.

The molecule as a whole, however, isn't polar - in the sense that it doesn't have an end (or a side) which is slightly negative and one which is slightly positive. The whole of the outside of the molecule is somewhat negative, but there is no overall separation of charge from top to bottom, or from left to right. By contrast, CHCl3 is polar.

The hydrogen at the top of the molecule is less electronegative than carbon and so is slightly positive. This means that the molecule now has a slightly positive "top" and a slightly negative "bottom", and so is overall a polar molecule. A polar molecule will need to be "lop-sided" in some way.

Patterns of electronegativity in the Periodic Table


The most electronegative element is fluorine. If you remember that fact, everything becomes easy, because electronegativity must always increase towards fluorine in the Periodic Table.
Note: This simplification ignores the noble gases. Historically this is because they were believed not to form bonds - and if they don't form bonds, they can't have an electronegativity value. Even now that we know that some of them do form bonds, data sources still don't quote electronegativity values for them.

Trends in electronegativity across a period As you go across a period the electronegativity increases. The chart shows electronegativities from sodium to chlorine - you have to ignore argon. It doesn't

have an electronegativity, because it doesn't form bonds.

Trends in electronegativity down a group As you go down a group, electronegativity decreases. (If it increases up to fluorine, it must decrease as you go down.) The chart shows the patterns of electronegativity in Groups 1 and 7.

Explaining the patterns in electronegativity The attraction that a bonding pair of electrons feels for a particular nucleus depends on: the number of protons in the nucleus; the distance from the nucleus; the amount of screening by inner electrons.
Note: If you aren't happy about the concept of screening orshielding, it would pay you to read the page on ionisation energies before you go on. The factors influencing ionisation energies are just the same as those influencing electronegativities. Use the BACK button on your browser to return to this page.

Why does electronegativity increase across a period? Consider sodium at the beginning of period 3 and chlorine at the end (ignoring the noble gas, argon). Think of sodium chloride as if it were covalently bonded.

Both sodium and chlorine have their bonding electrons in the 3-level. The electron pair is screened from both nuclei by the 1s, 2s and 2p electrons, but the chlorine nucleus has 6 more protons in it. It is no wonder the electron pair gets dragged so far towards the chlorine that ions are formed. Electronegativity increases across a period because the number of charges on the nucleus increases. That attracts the bonding pair of electrons more strongly. Why does electronegativity fall as you go down a group? Think of hydrogen fluoride and hydrogen chloride.

The bonding pair is shielded from the fluorine's nucleus only by the 1s2 electrons. In the chlorine case it is shielded by all the 1s22s22p6 electrons. In each case there is a net pull from the centre of the fluorine or chlorine of +7. But fluorine has the bonding pair in the 2-level rather than the 3-level as it is in chlorine. If it is closer to the nucleus, the attraction is greater. As you go down a group, electronegativity decreases because the bonding pair of electrons is increasingly distant from the attraction of the nucleus.

Warning! As far as I am aware, none of the UK-based A level (or equivalent) syllabuses any longer want the next bit. It used to be on the AQA syllabus, but has been removed from their new syllabus. At the time of writing, it does, however, still appear on at least one overseas A level syllabus (Malta, but there may be others that I'm not aware of). If in doubt, check your syllabus. Otherwise, ignore the rest of this

page. It is an alternative (and, to my mind, more awkward) way of looking at the formation of a polar bond. Reading it unnecessarily just risks confusing you.

The polarising ability of positive ions


What do we mean by "polarising ability"? In the discussion so far, we've looked at the formation of polar bonds from the point of view of the distortions which occur in a covalent bond if one atom is more electronegative than the other. But you can also look at the formation of polar covalent bonds by imagining that you start from ions. Solid aluminium chloride is covalent. Imagine instead that it was ionic. It would contain Al 3+ and Cl- ions. The aluminium ion is very small and is packed with three positive charges - the "charge density" is therefore very high. That will have a considerable effect on any nearby electrons.

We say that the aluminium ions polarise the chloride ions. In the case of aluminium chloride, the electron pairs are dragged back towards the aluminium to such an extent that the bonds become covalent. But because the chlorine is more electronegative than aluminium, the electron pairs won't be pulled half way between the two atoms, and so the bond formed will be polar. Factors affecting polarising ability Positive ions can have the effect of polarising (electrically distorting) nearby negative ions. The polarising ability depends on the charge density in the positive ion. Polarising ability increases as the positive ion gets smaller and the number of charges gets larger. As a negative ion gets bigger, it becomes easier to polarise. For example, in an iodide ion, I -, the outer electrons are in the 5-level - relatively distant from the

nucleus. A positive ion would be more effective in attracting a pair of electrons from an iodide ion than the corresponding electrons in, say, a fluoride ion where they are much closer to the nucleus. Aluminium iodide is covalent because the electron pair is easily dragged away from the iodide ion. On the other hand, aluminium fluoride is ionic because the aluminium ion can't polarise the small fluoride ion sufficiently to form a covalent bond.

INTERMOLECULAR BONDING - HYDROGEN BONDS This page explains the origin of hydrogen bonding - a relatively strong form of intermolecular attraction. If you are also interested in the weaker intermolecular forces (van der Waals dispersion forces and dipoledipole interactions), there is a link at the bottom of the page. The evidence for hydrogen bonding Many elements form compounds with hydrogen. If you plot the boiling points of the compounds of the Group 4 elements with hydrogen, you find that the boiling points increase as you go down the group.

The increase in boiling point happens because the molecules are getting larger with more electrons, and so van der Waals dispersion forces become greater.
Note: If you aren't sure about van der Waals dispersion forces, it would pay you to follow this link

before you go on.

If you repeat this exercise with the compounds of the elements in Groups 5, 6 and 7 with hydrogen, something odd happens.

Although for the most part the trend is exactly the same as in group 4 (for exactly the same reasons), the boiling point of the compound of hydrogen with the first element in each group is abnormally high. In the cases of NH3, H2O and HF there must be some additional intermolecular forces of attraction, requiring significantly more heat energy to break. These relatively powerful intermolecular forces are described as hydrogen bonds.

The origin of hydrogen bonding


The molecules which have this extra bonding are:

Note: The solid line represents a bond in the plane of the screen or paper. Dotted bonds are going back into the screen or paper away from you, and wedge-shaped ones are coming out towards you.

Notice that in each of these molecules: The hydrogen is attached directly to one of the most electronegative elements, causing the hydrogen to acquire a significant amount of positive charge. Each of the elements to which the hydrogen is attached is not only significantly negative, but also has at least one "active" lone pair. Lone pairs at the 2-level have the electrons contained in a relatively small volume of space which therefore has a high density of negative charge. Lone pairs at higher levels are more diffuse and not so attractive to positive things.
Note: If you aren't happy about electronegativity, you should follow this link before you go on.

Consider two water molecules coming close together.

The + hydrogen is so strongly attracted to the lone pair that it is almost as if you were beginning to form a coordinate (dative covalent) bond. It doesn't go that far, but the attraction is significantly stronger than an ordinary dipole-dipole interaction. Hydrogen bonds have about a tenth of the strength of an average covalent bond, and are being constantly broken and reformed in liquid water. If you liken the covalent bond between the oxygen and hydrogen to a stable marriage, the hydrogen bond has "just good friends" status. On the same scale, van der Waals attractions represent mere passing acquaintances! Water as a "perfect" example of hydrogen bonding Notice that each water molecule can potentially form four

hydrogen bonds with surrounding water molecules. There are exactly the right numbers of + hydrogens and lone pairs so that every one of them can be involved in hydrogen bonding. This is why the boiling point of water is higher than that of ammonia or hydrogen fluoride. In the case of ammonia, the amount of hydrogen bonding is limited by the fact that each nitrogen only has one lone pair. In a group of ammonia molecules, there aren't enough lone pairs to go around to satisfy all the hydrogens. In hydrogen fluoride, the problem is a shortage of hydrogens. In water, there are exactly the right number of each. Water could be considered as the "perfect" hydrogen bonded system.
Note: You will find more discussion on the effect of hydrogen bonding on the properties of water in the page on molecular structures.

More complex examples of hydrogen bonding


The hydration of negative ions When an ionic substance dissolves in water, water molecules cluster around the separated ions. This process is called hydration. Water frequently attaches to positive ions by co-ordinate (dative covalent) bonds. It bonds to negative ions using hydrogen bonds.
Note: If you are interested in the bonding in hydrated positive ions, you could follow this link to coordinate (dative covalent) bonding.

The diagram shows the potential hydrogen bonds formed to a chloride ion, Cl-. Although the lone pairs in the chloride ion are at the 3-level and wouldn't normally be active enough to form hydrogen bonds, in this case they are made more attractive by the full negative charge on the chlorine.

However complicated the negative ion, there will always be lone pairs that the hydrogen atoms from the water molecules can hydrogen bond to. Hydrogen bonding in alcohols An alcohol is an organic molecule containing an -O-H group. Any molecule which has a hydrogen atom attached directly to an oxygen or a nitrogen is capable of hydrogen bonding. Such molecules will always have higher boiling points than similarly sized molecules which don't have an -O-H or an -N-H group. The hydrogen bonding makes the molecules "stickier", and more heat is necessary to separate them. Ethanol, CH3CH2-O-H, and methoxymethane, CH3-OCH3, both have the same molecular formula, C2H6O.

Note: If you haven't done any organic chemistry yet, don't worry about the names.

They have the same number of electrons, and a similar length to the molecule. The van der Waals attractions (both dispersion forces and dipole-dipole attractions) in

each will be much the same. However, ethanol has a hydrogen atom attached directly to an oxygen - and that oxygen still has exactly the same two lone pairs as in a water molecule. Hydrogen bonding can occur between ethanol molecules, although not as effectively as in water. The hydrogen bonding is limited by the fact that there is only one hydrogen in each ethanol molecule with sufficient + charge. In methoxymethane, the lone pairs on the oxygen are still there, but the hydrogens aren't sufficiently + for hydrogen bonds to form. Except in some rather unusual cases, the hydrogen atom has to be attached directly to the very electronegative element for hydrogen bonding to occur. The boiling points of ethanol and methoxymethane show the dramatic effect that the hydrogen bonding has on the stickiness of the ethanol molecules: ethanol (with hydrogen bonding) 78.5C methoxymethane (without hydrogen bonding) 24.8C The hydrogen bonding in the ethanol has lifted its boiling point about 100C. It is important to realise that hydrogen bonding exists in addition to van der Waals attractions. For example, all the following molecules contain the same number of electrons, and the first two are much the same length. The higher boiling point of the butan-1-ol is due to the additional hydrogen bonding.

Comparing the two alcohols (containing -OH groups), both boiling points are high because of the additional hydrogen bonding due to the hydrogen attached directly to the oxygen - but they aren't the same. The boiling point of the 2-methylpropan-1-ol isn't as high as the butan-1-ol because the branching in the molecule makes the van der Waals attractions less effective than in the longer butan-1-ol. Hydrogen bonding in organic molecules containing

nitrogen Hydrogen bonding also occurs in organic molecules containing N-H groups - in the same sort of way that it occurs in ammonia. Examples range from simple molecules like CH3NH2(methylamine) to large molecules like proteins and DNA. The two strands of the famous double helix in DNA are held together by hydrogen bonds between hydrogen atoms attached to nitrogen on one strand, and lone pairs on another nitrogen or an oxygen on the other one. Learning outcome 3(h) This statement introduces bond energies, bond lengths and bond polarities, and looks at the relationship between them and the reactivity of a particular bond. Before you go on, you should find and read the statement in your copy of the syllabus. I see this as only an introductory topic at this point. You will meet examples relating to this throughout the course, in both inorganic and organic chemistry. The examples I have discussed lower down this page aren't meant to be learnt at this stage. Bond energy Bond energy is a measure of the strength of a particular covalent bond. In fact, it is sometimes referred to as "bond strength". You may find bond energies also referred to as bond enthalpies. There is a subtle difference between the two terms, but they are often used as if they mean the same thing. Bond energies in simple diatomic molecules Diatomic molecules are ones which contain two atoms. For a simple covalent molecule, X-Y, bond energy is the amount of energy needed to break one mole of the covalent bond to produce individual atoms, starting from the original substance in the gas state, and ending with gaseous atoms. X and Y can be the same or different. So, if they are the same, for example in chlorine gas, it is the energy needed to carry out this change per mole of bonds: The bond energy of the Cl-Cl bond is +244 kJ mol-1. That means that it takes 244 kJ to break 1 mole of Cl-Cl bonds.

Note: Don't worry if the units aren't familiar for now. You can just take the size of the number as a way of comparing bond strengths. As far as the values are concerned, I am using those from the CIE Data Booklet as far as possible, because that's what you will get in an exam. You will find a copy of this towards the end of the syllabus. The values I give for bond energies in this CIE part of Chemguide might be 1 or 2 kJ different from values I have used elsewhere on the site, because those came from a different source.

And for bromine, the reaction is still from gaseous bromine molecules to separate gaseous atoms.

In this case, the bond energy is +193 kJ mol -1. That means that the Br-Br bond is weaker than the Cl-Cl bond. If the atoms are different at each end of the bond, it works in exactly the same way. For example, the bond energy of the H-Cl bond is +431 kJ mol -1, and is given by the equation:

You have to be very careful about this. Bond energies only work if you are going from the substance in the gas state to gaseous atoms. So what if you have liquid bromine, for example? There will be extra energy needed to first convert the liquid into the gas. Similarly, in a real reaction, you aren't going to end up with gaseous atoms - they will react with other things, and there will be other energy changes there as well. In chemistry, it is often helpful to break complicated processes down into small steps. You will come across this a lot in the future. Bond energies in more complicated molecules The bond energy of the C-H bond, for example in methane, CH4, is quoted by CIE as 410 kJ mol -1. However, if you took methane to pieces one hydrogen at a time, it needs a different amount of energy to break each of the four C-H bonds. Every time you break a hydrogen off the carbon, the environment of those left behind changes. And the strength of a bond is affected by what else is around it. In cases like this, the bond energy quoted is an average value. In the methane case, you can work out how much energy is needed to break a mole of methane gas into gaseous carbon and hydrogen atoms. That comes to +1640 kJ and involves breaking 4 moles of C-H bonds. The average bond energy is therefore +1640/4 kJ, which is +410 kJ per mole of bonds. That means that many bond energies are actually quoted as averagebond energies, although it might not actually say so. In fact, tables of bond energies give average values in another sense as well. The bond energy of, say, the C-H bond varies depending on what is around it in the molecule. That means that if you use the C-H value in some calculation, you can't be sure that it exactly fits the molecule you are working with. When you come to doing simple calculations with bond energies later on in the course, you have to remember that the answers you get are only approximations to what happens in real reactions. Bond length Bond length measures the distance between the two nuclei in a covalent bond. The important thing about bond length is its relationship with bond energy. In a covalent bond, the two atoms are held together because both nuclei are attracted to the same pair of electrons. In a longer

bond, the shared electron pair is further from at least one of the two nuclei, and so the attractions are weaker. For example, in the hydrogen halides, as the bond gets longer, the distance of the electron pair from the nucleus of the halogen atom is going to get much greater, and that weakens the bond. bond energy (kJ mol-1) H-F H-Cl H-Br H-I +562 +431 +366 +299

The short H-F bond is very strong, and that has a major effect on the chemistry of fluorine. The bonds then get weaker as they get longer as you go down the Group. You will find examples of these effects in the chemistry of these elements when you do some Group 7 chemistry later. Bond polarity If you are working through these Section 3 syllabus statements in order, you should already have read the page about electronegativityin the bonding section of Chemguide. You need to be sure that you understand what electronegativity means and what causes electronegativity differences. And of course, you must know what is meant by a polar bond. A polar bond is one in which two atoms of unequal electronegativity share a pair of electrons. The electrons are pulled towards the atom with the greater electronegativity. That leaves the more electronegative atom slightly negatively charged and the less electronegative one slightly positively charged. What does all this mean for bond reactivity? When a reaction happens, some bonds are broken and new ones are made. Obviously bond energies will play a major part in this. But in many cases, the reaction is triggered by some sort of attraction between two molecules. This is where the polarity of the bonds plays a part. To illustrate this, I would like you to look at a page in the organic chemistry section of Chemguide about nucleophilic substitution reactions. Just read down as far as the red warning notice. Don't worry if this is all unfamiliar. All I am trying to illustrate is that

when you look at a reaction in this sort of detail, you have to consider both bond polarities, and bond energies. In this particular case, the success of the reaction is determined by the strength of the carbon-halogen bond, and not the polarity of the bond. And here is another example, also from organic chemistry. Have a look at a page about electrophilic addition reactions. Remember that you aren't trying to learn this at the moment. I just want to give you an impression of the way in which we can explain how reactions happen, and the role of bond polarities in this. Now look at a page about the reactions of ethene with hydrogen halides.. Ignore any mention of cyclohexene, and concentrate on the section titled "Electrophilic addition reactions involving the other hydrogen halides". Notice that, again, what determines how fast the reactions happen is the strength of the hydrogen-halogen bonds, not their polarity. Hydrogen fluoride is the most polar molecule, but the reaction is the slowest because of the difficulty of breaking the hydrogen-fluorine bond. And one more example: Hydrogen and chlorine react to make hydrogen chloride. Hydrogen and iodine react to make hydrogen iodide. In each case, you have to break the H-H bonds in the hydrogen gas, and the halogen-halogen bonds in the chlorine or iodine. The bond energies for the two halogens are: bond energy (kJ mol-1) Cl-Cl I-I +244 +151

So which is the more reactive halogen, chlorine or iodine? Well, if you look at the bond energies, it is much easier to break iodine-iodine bonds than chlorine-chlorine bonds. So you might expect the iodine to react faster. Not so! In the presence of a flame or bright sunlight, chlorine and hydrogen react together explosively. On the other hand, you have to heat hydrogen and iodine together continuously, and even then only a proportion of them actually react. By contrast with the chlorine case, the hydrogen and iodine reaction is really feeble. The problem is that the bond energies of the halogens are just one factor in the overall process, and you need to look at the strengths of the bonds being formed as well. This is a problem for later on in the course. I just wanted to point out at this stage that

you can't safely draw any conclusions from what happens in just one step of a complicated series of reactions. Summary Bond length plays a part in the reactivity of a covalent bond because it affects bond strength. Bond polarity affects reactions because it can help in getting the molecules attracted and lined up in the right way. In reactions involving covalent molecules, bond energies will play a part in determining reactivity, but you can't make assumptions about this just by looking at one stage of the overall process.
Note: I worry about this learning outcome being included at this early stage of the syllabus. There is a real danger of making oversimplifications in order not to scare students too much. Explaining reactivity really has to be done on a case-bycase basis. But there is hope! In the six sets of exam papers, plus the specimen paper, that I had available at the time of writing (June 2010), I couldn't find a single question which related directly to this statement.

NUCLEOPHILIC SUBSTITUTION
Background
Bonding in the halogenoalkanes Halogenoalkanes (also known as haloalkanes or alkyl

halides) are compounds containing a halogen atom (fluorine, chlorine, bromine or iodine) joined to one or more carbon atoms in a chain. The interesting thing about these compounds is the carbonhalogen bond, and all the nucleophilic substitution reactions of the halogenoalkanes involve breaking that bond. The polarity of the carbon-halogen bonds With the exception of iodine, all of the halogens are more electronegative than carbon. Electronegativity values (Pauling scale) C 2.5 F 4.0 Cl 3.0 Br 2.8 I 2.5
Note: If you aren't sure about electronegativity and bond polarity follow this link before you read on. Use the BACK button on your browser to return to this page.

That means that the electron pair in the carbon-halogen bond will be dragged towards the halogen end, leaving the halogen slightly negative ( -) and the carbon slightly positive ( +) - except in the carbon-iodine case. Although the carbon-iodine bond doesn't have a permanent dipole, the bond is very easily polarised by anything approaching it. Imagine a negative ion approaching the bond from the far side of the carbon atom:

The fairly small polarity of the carbon-bromine bond will be increased by the same effect. The strengths of the carbon-halogen bonds
Note: If you haven't done any work on bond strengths, or are a bit rusty, it doesn't matter. Just realise that the bigger the number, the stronger the

bond. And don't worry if you have found slightly different numbers in a different data source - there is a lot of variability in the quoted values, but the overall pattern is still the same.

Look at the strengths of various bonds (all values in kJ mol 1 ). C-H 413 C-F 467 C-Cl 346 C-Br 290 C-I 228 In all of these nucleophilic substitution reactions, the carbon-halogen bond has to be broken at some point during the reaction. The harder it is to break, the slower the reaction will be. The carbon-fluorine bond is very strong (stronger than C-H) and isn't easily broken. It doesn't matter that the carbonfluorine bond has the greatest polarity - the strength of the bond is much more important in determining its reactivity. You might therefore expect fluoroalkanes to be very unreactive - and they are! We shall simply ignore them from now on. In the other halogenoalkanes, the bonds get weaker as you go from chlorine to bromine to iodine. That means that chloroalkanes react most slowly, bromoalkanes react faster, and iodoalkanes react faster still. Rates of reaction: RCl < RBr < RI Where "<" is read as "is less than" - or, in this instance, "is slower than", and R represents any alkyl group.

Warning! Before you read on it is essential that you know exactly what your syllabus says about these reactions so that you can extract the right amount of detail from what follows. The problem is that there are two different mechanisms depending on the type of halogenoalkane you are using (whether primary, secondary or tertiary). Some syllabuses try to make things simpler by restricting you to just one of the two mechanisms. Where the syllabus is vague, look at recent exam papers and mark schemes, or any support material published by your examiners. Haven't got a syllabus or recent exam papers? If you are working to one of the UK-based syllabuses

for 16 - 18 year olds, follow this link to find out how to get them. You mustknow what your examiners expect in this topic.

Be sure that you understand the relationship between these simplified diagrams and the full structures. The mechanisms The reactions are examples of electrophilic addition. With ethene and HBr:

and with cyclohexene:

Electrophilic addition reactions involving the other hydrogen halides


The facts Hydrogen chloride and the other hydrogen halides add on in exactly the same way. For example, hydrogen chloride adds to ethene to make chloroethane: The only difference is in how fast the reactions happen with the different hydrogen halides. The rate of reaction increases as you go from HF to HCl to HBr to HI. HF HCl HBr HI fastest reaction slowest reaction

The reason for this is that as the halogen atoms get bigger, the strength of the hydrogen-halogen bond falls. Bond strengths (measured in kilojoules per mole) are: H-F H-Cl H-Br H-I 568 432 366 298

Note: You may find slightly different values depending on which data source you use. It doesn't matter - the differences are minor and the pattern is always the same.

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

one half to the other.


Note: This diagram shows a side view of an ethene molecule. The dotted lines to two of the hydrogens show bonds going back into the screen or paper away from you. The wedge shapes show bonds coming out towards you.

The pi electrons are not as fully under the control of the carbon nuclei as the electrons in the sigma bond and, because they lie exposed above and below the rest of the molecule, they are relatively open to attack by other things.
Note: Check your syllabus to see if you need to know how a pi bond is formed. Haven't got a syllabus? If you are working towards a UK-based exam, find out how to get one by following this link. If you do need to know about the bonding in ethene in detail, follow this link as well.

Electrophiles An electrophile is something which is attracted to electron-rich regions in other molecules or ions. Because it is attracted to a negative region, an electrophile must be something which carries either a full positive charge, or has a slight positive charge on it somewhere.
Note: The ending ". . phile" means a liking for. For example, a francophile is someone who likes the French; an anglophile is someone who likes the English.

Ethene and the other alkenes are attacked by electrophiles. The electrophile is normally the slightly positive ( +) end of a molecule like hydrogen bromide, HBr.
Note: If you aren't sure about why some bonds are polar, read the page on electronegativity. Use the BACK button on your browser to return to this page.

Electrophiles are strongly attracted to the exposed electrons in

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

Summary: electrophilic addition reactions An addition reaction is a reaction in which two molecules join together to make a bigger one. Nothing is lost in the process. All the atoms in the original molecules are found in the bigger one. An electrophilic addition reaction is an addition reaction which happens because what we think of as the "important" molecule is attacked by an electrophile. The "important" molecule has a region of high electron density which is attacked by something carrying some degree of positive charge.
Note: When we talk about reactions of alkenes like ethene, we think of the ethene as being attacked by other molecules such as hydrogen bromide. Because ethene is the molecule we are focusing on, we quite arbitrarily think of it as the central molecule and hydrogen bromide as its attacker. There's no real justification for this, of course, apart from the fact that it helps to put things in some sort of logical pattern. In reality, the molecules just collide and may react if they have enough energy and if they are lined up correctly.

Understanding the electrophilic addition mechanism


The mechanism for the reaction between ethene and a molecule X-Y

It is very unlikely that any two different atoms joined together will have the same electronegativity. We are going to assume that Y is more electronegative than X, so that the pair of electrons is pulled slightly towards the Y end of the bond. That means that the X atom carries a slight positive charge.
Note: Once again, if you aren't sure about electronegativity and bond polarity follow this link before you read on. Use the BACK button on your browser to return to this page.

The slightly positive X atom is an electrophile and is attracted to the exposed pi bond in the ethene. Now imagine what happens as they approach each other.

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

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

Important term An ion in which the positive charge is carried on a carbon atom is called a carbocation or acarbonium ion (an older term). In the final stage of the reaction the electrons in the lone pair on the Y- ion are strongly attracted towards the positive carbon atom. They move towards it and form a co-ordinate (dative covalent) bond between the Y and the carbon.
Help! A co-ordinate (dative covalent) bond is simply a covalent bond in which both shared electrons originate from the same atom. The bond formed between the X and the other carbon atom was also a co-ordinate bond. Once a coordinate bond has been formed there is no difference whatsoever between it and any other covalent bond.

How to write this mechanism in an exam The movements of the various electron pairs are shown using curly arrows.
Help! If you aren't sure about the use of curly arrows in mechanisms, you must follow this link before you go on. Use the BACK button on your browser to return to this page.

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

Learning outcome 3(i) This statement deals with van der Waals forces. Before you go on, you should find and read the statement in your copy of the syllabus. Electronegativity Important: If you haven't read my comment about electronegativity on the page about learning outcome 3(g), go and do that before you do anything else. van der Waals forces First, read the page about van der Waals forces. That will give you all that you need (plus a bit more) to satisfy this syllabus statement, apart from the bit about liquid bromine. You will find an explanation of how van der Waals forces affect the halogens (including bromine) on a page about the physical properties of Group 7. You just need the section titled "Trends in Melting Point and Boiling Point" just over half-way down that page. Hydrogen bonding If you delayed reading about hydrogen bonding in section 3(g), go back and do it now before you forget!

INTERMOLECULAR BONDING - VAN DER WAALS FORCES


This page explains the origin of the two weaker forms of intermolecular attractions - van der Waals dispersion forces and dipole-dipole attractions. If you are also interested in hydrogen bonding there is a link at the bottom of the page.

What are intermolecular attractions?


Intermolecular versus intramolecular bonds Intermolecular attractions are attractions between one molecule and a neighbouring molecule. The forces of attraction

which hold an individual molecule together (for example, the covalent bonds) are known as intramolecular attractions. These two words are so confusingly similar that it is safer to abandon one of them and never use it. The term "intramolecular" won't be used again on this site. All molecules experience intermolecular attractions, although in some cases those attractions are very weak. Even in a gas like hydrogen, H2, if you slow the molecules down by cooling the gas, the attractions are large enough for the molecules to stick together eventually to form a liquid and then a solid. In hydrogen's case the attractions are so weak that the molecules have to be cooled to 21 K (-252C) before the attractions are enough to condense the hydrogen as a liquid. Helium's intermolecular attractions are even weaker - the molecules won't stick together to form a liquid until the temperature drops to 4 K (-269C).

van der Waals forces: dispersion forces


Dispersion forces (one of the two types of van der Waals force we are dealing with on this page) are also known as "London forces" (named after Fritz London who first suggested how they might arise). The origin of van der Waals dispersion forces Temporary fluctuating dipoles Attractions are electrical in nature. In a symmetrical molecule like hydrogen, however, there doesn't seem to be any electrical distortion to produce positive or negative parts. But that's only true on average. The lozenge-shaped diagram represents a small symmetrical molecule - H2, perhaps, or Br2. The even shading shows that on average there is no electrical distortion. But the electrons are mobile, and at any one instant they might find themselves towards one end of the molecule, making that end -. The other end will be temporarily short of electrons and so becomes +.
Note: (read as "delta") means "slightly" - so + means "slightly positive".

An instant later the electrons may well have moved up to the

other end, reversing the polarity of the molecule. This constant "sloshing around" of the electrons in the molecule causes rapidly fluctuating dipoles even in the most symmetrical molecule. It even happens in monatomic molecules - molecules of noble gases, like helium, which consist of a single atom. If both the helium electrons happen to be on one side of the atom at the same time, the nucleus is no longer properly covered by electrons for that instant.

How temporary dipoles give rise to intermolecular attractions I'm going to use the same lozenge-shaped diagram now to represent any molecule which could, in fact, be a much more complicated shape. Shape does matter (see below), but keeping the shape simple makes it a lot easier to both draw the diagrams and understand what is going on. Imagine a molecule which has a temporary polarity being approached by one which happens to be entirely non-polar just at that moment. (A pretty unlikely event, but it makes the diagrams much easier to draw! In reality, one of the molecules is likely to have a greater polarity than the other at that time - and so will be the dominant one.) As the right hand molecule approaches, its electrons will tend to be attracted by the slightly positive end of the left hand one. This sets up an induced dipole in the approaching molecule, which is orientated in such a way that the + end of one is attracted to the - end of the other.

An instant later the electrons in the left hand molecule may well have moved up the other end. In doing so, they will repel the electrons in the right hand one. The polarity of both molecules reverses, but you still have +

attracting -. As long as the molecules stay close to each other the polarities will continue to fluctuate in synchronisation so that the attraction is always maintained. There is no reason why this has to be restricted to two molecules. As long as the molecules are close together this synchronised movement of the electrons can occur over huge numbers of molecules.

This diagram shows how a whole lattice of molecules could be held together in a solid using van der Waals dispersion forces. An instant later, of course, you would have to draw a quite different arrangement of the distribution of the electrons as they shifted around - but always in synchronisation. The strength of dispersion forces Dispersion forces between molecules are much weaker than the covalent bonds within molecules. It isn't possible to give any exact value, because the size of the attraction varies considerably with the size of the molecule and its shape. How molecular size affects the strength of the dispersion forces The boiling points of the noble gases are helium -269C neon -246C argon -186C krypton -152C xenon -108C radon -62C All of these elements have monatomic molecules. The reason that the boiling points increase as you go down the group is that the number of electrons increases, and so also does the radius of the atom. The more electrons you have, and the more distance over which they can move, the bigger the possible temporary dipoles and therefore the bigger the dispersion forces.

Because of the greater temporary dipoles, xenon molecules are "stickier" than neon molecules. Neon molecules will break away from each other at much lower temperatures than xenon molecules - hence neon has the lower boiling point. This is the reason that (all other things being equal) bigger molecules have higher boiling points than small ones. Bigger molecules have more electrons and more distance over which temporary dipoles can develop - and so the bigger molecules are "stickier". How molecular shape affects the strength of the dispersion forces The shapes of the molecules also matter. Long thin molecules can develop bigger temporary dipoles due to electron movement than short fat ones containing the same numbers of electrons. Long thin molecules can also lie closer together - these attractions are at their most effective if the molecules are really close. For example, the hydrocarbon molecules butane and 2methylpropane both have a molecular formula C 4H10, but the atoms are arranged differently. In butane the carbon atoms are arranged in a single chain, but 2-methylpropane is a shorter chain with a branch.

Butane has a higher boiling point because the dispersion forces are greater. The molecules are longer (and so set up bigger temporary dipoles) and can lie closer together than the shorter, fatter 2-methylpropane molecules.

van der Waals forces: dipole-dipole interactions


Warning! There's a bit of a problem here with modern syllabuses. The majority of the syllabuses talk as if dipole-dipole interactions were quite distinct from van der Waals forces. Such a syllabus will talk about van der Waals forces (meaning dispersion forces) and, separately, dipole-dipole interactions. All intermolecular attractions are known collectively as van der Waals forces. The various different types were first explained by different people at different times. Dispersion forces, for example, were described by London in 1930; dipole-dipole interactions by Keesom in 1912. This oddity in the syllabuses doesn't matter in the least as far as understanding is concerned - but you obviously must know what your particular examiners mean by the terms they use in the

questions. Check your syllabus. If you are working to a UK-based syllabus for 16 - 18 year olds, but don't have a copy of it, follow this link to find out how to get one.

A molecule like HCl has a permanent dipole because chlorine is more electronegative than hydrogen. These permanent, in-built dipoles will cause the molecules to attract each other rather more than they otherwise would if they had to rely only on dispersion forces.
Note: If you aren't happy about electronegativity and polar molecules, follow this link before you go on.

It's important to realise that all molecules experience dispersion forces. Dipole-dipole interactions are not an alternative to dispersion forces - they occur in addition to them. Molecules which have permanent dipoles will therefore have boiling points rather higher than molecules which only have temporary fluctuating dipoles. Surprisingly dipole-dipole attractions are fairly minor compared with dispersion forces, and their effect can only really be seen if you compare two molecules with the same number of electrons and the same size. For example, the boiling points of ethane, CH3CH3, and fluoromethane, CH3F, are

Why choose these two molecules to compare? Both have identical numbers of electrons, and if you made models you would find that the sizes were similar - as you can see in the diagrams. That means that the dispersion forces in both molecules should be much the same. The higher boiling point of fluoromethane is due to the large permanent dipole on the molecule because of the high electronegativity of fluorine. However, even given the large permanent polarity of the molecule, the boiling point has only

been increased by some 10. Here is another example showing the dominance of the dispersion forces. Trichloromethane, CHCl 3, is a highly polar molecule because of the electronegativity of the three chlorines. There will be quite strong dipole-dipole attractions between one molecule and its neighbours.

On the other hand, tetrachloromethane, CCl 4, is non-polar. The outside of the molecule is uniformly - in all directions. CCl 4 has to rely only on dispersion forces.

So which has the highest boiling point? CCl 4 does, because it is a bigger molecule with more electrons. The increase in the dispersion forces more than compensates for the loss of dipoledipole interactions. The boiling points are: CHCl3 61.2C CCl4 76.8C
Note: A student pointed out to me that many web and book sources and teachers describe dispersion forces as being the weakest of the intermolecular forces, quoting values of, perhaps, up to 4 kJ/mol. That conflicts with what I have said above that "dipole-dipole attractions are fairly minor compared with dispersion forces". I have discussed this question of the strength of dispersion forces on a separate page, where I have tried to show that those web and book sources and teachers are wrong!

Trends in Melting Point and Boiling Point

You will see that both melting points and boiling points rise as you go down the Group. If you explore the graphs, you will find that fluorine and chlorine are gases at room temperature, bromine is a liquid and iodine a solid. Nothing very surprising there! Explaining the trends in melting point and boiling point All of the halogens exist as diatomic molecules - F2, Cl2, and so on. The intermolecular attractions between one molecule and its neighbours are van der Waals dispersion forces.
Note: If you aren't sure about van der Waals dispersion forces, you will find them covered in detail in another part of this site. You won't understand the next bit unless you are happy about dispersion forces and how they vary with the size of the molecule. Use the BACK button on your browser to return quickly to this page.

As the molecules get bigger there are obviously more electrons which can move around and set up the temporary dipoles which create these attractions. The stronger intermolecular attractions as the molecules get bigger means that you have to supply more heat energy to turn them into either a liquid or a gas - and so their melting and boiling points rise.

Learning outcome 3(j) This statement introduces metallic bonding. Before you go on, you should find and read the statement in your copy of the syllabus.

You need to read the page about metallic bonding in the bonding section of Chemguide. You could follow the invitation to "To explore the structure of metals" at the bottom of the page if you want to. This is part of the next learning outcome in Section 3.

METALLIC BONDING
This page introduces the bonding in metals. It explains how the metallic bond arises and why its strength varies from metal to metal.

What is a metallic bond?


Metallic bonding in sodium Metals tend to have high melting points and boiling points suggesting strong bonds between the atoms. Even a metal like sodium (melting point 97.8C) melts at a considerably higher temperature than the element (neon) which precedes it in the Periodic Table. Sodium has the electronic structure 1s22s22p63s1. When sodium atoms come together, the electron in the 3s atomic orbital of one sodium atom shares space with the corresponding electron on a neighbouring atom to form a molecular orbital - in much the same sort of way that a covalent bond is formed. The difference, however, is that each sodium atom is being touched by eight other sodium atoms - and the sharing occurs between the central atom and the 3s orbitals on all of the eight other atoms. And each of these eight is in turn being touched by eight sodium atoms, which in turn are touched by eight atoms and so on and so on, until you have taken in all the atoms in that lump of sodium. All of the 3s orbitals on all of the atoms overlap to give a vast number of molecular orbitals which extend over the whole piece of metal. There have to be huge numbers of molecular orbitals, of course, because any orbital can only hold two electrons. The electrons can move freely within these molecular orbitals, and so each electron becomes detached from its parent atom. The electrons are said to be delocalised. The metal is held together by the strong forces of attraction between the positive nuclei and the delocalised electrons.

This is sometimes described as "an array of positive ions in a sea of electrons". If you are going to use this view, beware! Is a metal made up of atoms or ions? It is made of atoms. Each positive centre in the diagram represents all the rest of the atom apart from the outer electron, but that electron hasn't been lost - it may no longer have an attachment to a particular atom, but it's still there in the structure. Sodium metal is therefore written as Na - not Na+. Metallic bonding in magnesium If you work through the same argument with magnesium, you end up with stronger bonds and so a higher melting point. Magnesium has the outer electronic structure 3s 2. Both of these electrons become delocalised, so the "sea" has twice the electron density as it does in sodium. The remaining "ions" also have twice the charge (if you are going to use this particular view of the metal bond) and so there will be more attraction between "ions" and "sea". More realistically, each magnesium atom has one more proton in the nucleus than a sodium atom has, and so not only will there be a greater number of delocalised electrons, but there will also be a greater attraction for them. Magnesium atoms have a slightly smaller radius than sodium atoms, and so the delocalised electrons are closer to the nuclei. Each magnesium atom also has twelve near neighbours rather than sodium's eight. Both of these factors increase the strength of the bond still further. Metallic bonding in transition elements Transition metals tend to have particularly high melting points and boiling points. The reason is that they can involve the 3d electrons in the delocalisation as well as the 4s. The more electrons you can involve, the stronger the attractions tend to be.
Note: If you aren't happy about the electronic structure of transition metals, then you might like to follow this link to revise

it.

The metallic bond in molten metals In a molten metal, the metallic bond is still present, although the ordered structure has been broken down. The metallic bond isn't fully broken until the metal boils. That means that boiling point is actually a better guide to the strength of the metallic bond than melting point is. On melting, the bond is loosened, not broken.

Learning outcomes 3(k) and 3(l) These two statements deal with the effect of the type of bonding on the physical properties of a substance. Before you go on, you should find and read the statements in your copy of the syllabus. You should work through all the topics on the structures menu of Chemguide. You will come back to much of this again in Sections 4 and 9, so even if it takes a bit of time and effort, it will be worth it. In the ionic structures page, you can miss out the bit about caesium chloride. Learning outcome 3(m)
This statement is an introduction to the idea that energy changes during a reaction are caused by bond breaking and making. Before you go on, you should find and read the statement in your copy of the syllabus. This is dealt with numerically in Section 5, and this current page is only intended to give a brief introduction. Chemical reactions involve breaking the bonds in one set of substances, and reorganising them to make new bonds in new substances. Breaking bonds needs energy That's fairly obvious really. Whatever sort of bond you are talking about, it involves one thing being attracted to something else. To separate them, you are going to have to put in some energy. Making bonds releases energy That's not so obvious. Whenever you set up an attraction of any sort, energy is released. Think of it like this. Suppose you have a covalent bond between two atoms X and Y. To split them into the two individual atoms takes, say, 350 kJ for a mole of bond. You can show this happening on an energy diagram.

Nows suppose you remade the bond from the same atoms. Obviously, you would be exactly reversing what you had just done in breaking the bond, and will end up exactly where you started. That means that you will get out exactly the same amount of energy as you put in to break the bond.

Exothermic and endothermic changes In any reaction, you may well be breaking several bonds, and so you will need to put in enough energy to do this. Then you will make a number of bonds and, at that stage, you will get energy released again. Why are some reactions exothermic? Suppose you need to put in 1000 kJ to break all the necessary bonds in your starting substances. Suppose also that the new bonds formed are stronger than the original ones, so that you get 1200 kJ of energy released as the new bonds are made. You get out 200 kJ of energy more than you put in. That's an exothermic reaction - one which releases the extra energy as heat. Why are some reactions endothermic? Suppose you need to put in 1000 kJ to break all the necessary bonds in your starting substances. Suppose this time that the new bonds formed are weaker than the original ones, so that you only get 900 kJ of energy released as the new bonds are made.

You put in 100 kJ of energy more than you get out. That's an endothermic reaction - one which absorbs heat energy.

Das könnte Ihnen auch gefallen