Você está na página 1de 85

ATOMIC STRUCTURE AND BONDING MENU

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

ATOMIC PROPERTIES MENU


Simple background . . . Revises the simple knowledge you should already have about the structure of atoms from introductory courses (e.g. GCSE). Atomic orbitals . . . Explains what atomic orbitals are and discusses their shapes and relative energies. This is essential pre -reading before you go on to any of the remaining topics in this section. Electronic structures . . . How to work out and write the electronic structures for atoms and simple monatomic ions (containing only one atom - e.g. Cl- or Mg2+) using s, p, d notation. Ionisation energies . . . Explains what ionisation energies are and how and why they vary around the Periodic Table. Hydrogen's atomic emission spectrum . . . An introduction to the atomic hydrogen emission spectrum, and how it can be used to find the ionisation energy of hydrogen. Electron affinities . . . Explains what electron affinities are and how and why they vary around the Periodic Table. Atomic and ionic radii . . . Looks at the various measures of atomic radius, and explains how and why atomic radii vary around the Periodic Table. Also co nsiders how the radii of positive and negative ions differ from the atoms they come fr om.

A SIMPLE VIEW OF ATOMIC STRUCTURE

The sub-atomic particles


Protons, neutrons and electrons.

proton neutron electron

relative mass 1 1 1/1836

relative charge +1 0 -1

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

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 particl es, all with the same energy, are passed between two electrically charged plates:

y y

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

Neutrons continue in a straight line.

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

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

y y

Protons are deflected on a curved path towards the negative plate. Electrons are deflected on a curved path towards the positive plate. If the electrons and protons are travelling with the same speed, then the lighter electrons are deflected far more strongly t han 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 dia grams 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 theyprobably 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 a s nucleons. Virtually all the mass of the atom is concentrated in the nucleus, because the electrons weigh so little.

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

This information can be given simply in the form:

How many protons and neutrons has this atom got? The atomic number counts the number of protons (9); the mass number counts protons + neutrons (19). If there are 9 protons, t here 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 wh at sort of element you are talking about. So if an atom has 8 protons (atomic number = 8), it m ust 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 same number of protons, but the number of neutrons varies. protons 6 6 6 neutrons 6 7 8
12C, 13C

and 14C. They all have the

carbon-12 carbon-13 carbon-14

mass number 12 13 14

These different atoms of carbon are called isotopes. The fact that they have varying numbers of neutrons makes no difference whatsoever to the chemical reactions of the carbon. Isotopes are atoms which have the same atomic number but different mass number s. 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 = 1 7) 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 le vel 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 n ucleus. Electrons will always go into the lowest possible energy level (nearest the nucleus) - provided there is space. To work out the electronic arrangement of an atom

y y y

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

e.g. to find the electronic arrangement in chlorine

y y y

The Periodic Table gives you the atomic number of 17. Therefore there are 17 protons and 17 electrons. The arrangement of the electrons will be 2, 8, 7 (i.e. 2 in the first level, 8 in the second, and 7 in the third).

The electronic arrangements of the first 20 elements

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

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 oute r 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 a d can be misleading. It gives the impression that the n electrons are circling the nucleus in orbits like planets around the sun. As you will find when you look at the A'level view of the atom, it is impossible to know exactly how they are actually moving.

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

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

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

ATOMIC ORBITALS

This page explains what atomic orbitals are in a way that makes them understandable for introductory courses such as UK A lev el 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 l ooks 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 wher e 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 momentu m of an electron.) That makes it impossible to plot an orbit for an electron a round 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 electronsas 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 justfor 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. T he diagram shows a cross-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 spa ce 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 abou t 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 nucl eus. 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 th e 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 thenucleus? 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 ea ch 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 2p x, 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 (3p x, 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 a s 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 p robably won't be expected to draw them or name them. Apart from a passing reference, you won't come across f orbitals at al l.

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 inclu ding 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 o them. f

Fitting electrons into orbitals


You can think of an atom as a very bizarre house (like an inverted pyramid!) - with the nucleus living on the ground floor, and then various rooms (orbitals) on the higher floors occupied by the electrons. On the first floor there is only 1 room (the 1s orbital); on the second floor there are 4 rooms (the 2s, 2px, 2py and 2pz orbitals); on the third floor there are 9 rooms (one 3s orbital, three 3p orbitals and five 3d orbitals ); and so on. But the rooms aren't very big . . . Each orbital can only hold 2 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 th e 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 liv in different orbitals, that's fine - but if they are both in the same orbital e 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 1s 2. 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 betwe en 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 ene rgies 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 a lways 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 betw een the energy levels that the 4f orbitals don't fill until after the 6s, for example.

ELECTRONIC STRUCTURES

This page explores how you write electronic structures for atoms using s, p, and d notation. It assumes that you know about s imple 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 s imple 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 ask ed 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 1s 22s1. Beryllium adds a second electron to this same level - 1s22s2. Now the 2p levels start to fill. These levels all have the same energy, and so the electrons go in singly at first. B C N 1s22s22px1 1s22s22px12py1 1s22s22px12py12pz1

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

The next electrons to go in will have to pair up with those already there. O F Ne 1s22s22px22py12pz1 1s22s22px22py22pz1 1s22s22px22py22pz2

You can see that it is going to get progressively tedious to write the full electronic structures of atoms as the number of e lectrons 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 1s 22s22p6. This is what is normally done if the electrons are in an inner layer. If the el ectrons 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 1s 22s22p63s23px23py23pz1. 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 o f 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: 1s 22s22px22py22pz2 You wouldn't do this with helium because it takes longer to write [He] than it does 1s 2. On this basis the structure of chlorine would be written [Ne]3s 23px23py23pz1. 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 Mg S Ar 1s22s22p6 3s2 1s22s22p6 3s23px23py13pz1 1s22s22p6 3s23px23py23pz2 [Ne]3s2 [Ne]3s23px23py13pz1 [Ne]3s23px23py23pz2

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

The beginning of the fourth period At this point the 3-level orbitals aren't all full - the 3d levels haven't been used yet. But if you refer back to the energies of the orbitals, you will see that the next lowest energy orbital is the 4s - so that fills next. K Ca 1s22s22p63s23p64s1 1s22s22p63s23p64s2

There is strong evidence for this in the similarities in the chemistry of elements like sodium (1s 22s22p63s1) and potassium (1s 22s22p63s23p64s1) 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 ns 1 (where n is a number between 2 and 7). All group 2 elements have an outer electronic structure of ns 2. 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 add ed to the atom is in a d orbital. The first series of these contains the elements from scandium to zinc, which at GCSE you probably called transition elements or transition metals. The terms "transition element" and "d-block element" don't quite have the same meaning, but it doesn't matter in the present context.

If you are interested: A transition element is defined as one which has partially filled d orbitals either in the element or any of its compounds. Zinc (at the righthand end of the d-block) always has a completely full 3d level (3d10 ) and so doesn't count as a transition element.

d electrons are almost always described as, for example, 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 th at they will have to pair up.

d5 means

d8 means Notice in what follows that all the 3-level orbitals are written together, even though the 3d electrons are added to the atom after the 4s. Sc Ti V Cr 1s22s22p63s23p63d14s2 1s22s22p63s23p63d24s2 1s22s22p63s23p63d34s2 1s22s22p63s23p63d54s1

Whoops! Chromium breaks the sequence. In chromium, the electrons in the 3d and 4s orbitals rearrange so that there is one ele ctron in each orbital. It would be convenient if the sequence was tidy - but it's not! Mn Fe Co Ni Cu Zn 1s22s22p6 3s23p63d54s2 1s22s22p6 3s23p63d64s2 1s22s22p6 3s23p63d74s2 1s22s22p6 3s23p63d84s2 1s22s22p6 3s23p63d104s1 1s22s22p6 3s23p63d104s2 (another awkward one!) (back to being tidy again)

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 1s 22s22p63s23p63d104s24px24py24pz1.

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

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

y y y

Use the Periodic Table to find the atomic number, and hence number of electrons. Fill up orbitals in the order 1s, 2s, 2p, 3s, 3p, 4s, 3d, 4p - until you run out of electrons. The 3d is the awkward one - remember that specially. Fill p and d orbitals singly as far as possible before pairing electr ons 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 t o 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 fo rgetting 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 5s 25px25py25pz1. 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 m ake 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 6s 2. Including all the inner levels: 1s 22s22p63s23p63d104s24p64d105s25p66s2. It would be easy to include 5d10 as well by mistake, but the d level always fills after the next s level - so 5d fills after 6s just as 3d fills after 4s. As long as you counted the number of electrons you could easily spot this mistake because you would have 10 too many.

Note: Don't worry too much about these complicated structures. You need to know how to work them out in principle, but your examine are much more likely to ask you for something rs simple like sulphur or iron.

IONISATION ENERGY

This page explains what first ioni sation 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 lin k 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 mome t. n

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 v ery 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 a mount 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 expl ained 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 ele ctrons 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 o ne 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 usefu l!) If the outer electron looks in towards the nucleus, it doesn't see the nucleus sharply. Betwee n 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+ f rom 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 1s 1. 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 1s 2. 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 of fset 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 1s 2 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 1s 2 electrons. The major difference is the increasing number of protons in the nucleus as you go from lithium to neon. That causes greater a ttraction between the nucleus and the electrons and so increases the ionisation energies. In fact the increasing nuclear charge also drags the oute r 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 t hird level and are screened by the 1s 22s22p6 electrons. They all have the same sort of environment, but there is an increasing nuclear charge. Why the drop between groups 2 and 3 (Be-B and Mg-Al)?

The explanation lies with the structures of boron and aluminium. The outer electron is removed more easily from these atoms than the general trend in their period would suggest. Be B 1s22s2 1s22s22px1 1st I.E. = 900 kJ mol -1 1st I.E. = 799 kJ mol -1

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

y y

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

The explanation for the drop between magnesium and aluminium is the same, except that everything is happening at the 3-level rather than the 2level. Mg Al 1s22s22p63s2 1s22s22p63s23px 1 1st I.E. = 736 kJ mol -1 1st I.E. = 577 kJ mol -1

The 3p electron in aluminium is slightly more distant from the nucleus than the 3s, and partially screened by the 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 energ is abnormally high. This is quite simply wrong! The reason for the y fluctuation is because the group 3 value is lower than you might expect for the reasons we've looked at.

Why the drop between groups 5 and 6 (N-O and P-S)? Once again, you might expect the ionisation energy of the group 6 element to be higher than that of group 5 because of the extra proton. What is offsetting it this time? N O 1s22s22px12py 12pz1 1s22s22px22py 12pz1 1st I.E. = 1400 kJ mol -1 1st I.E. = 1310 kJ mol -1

The screening is identical (from the 1s 2 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 2p x2 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 th e 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 exp ected a much larger ionisation energy in sodium, but offsetting the nuclear charge is a greater distance from the nucleus and more screening. Li Na 1s22s1 1s22s22p63s1 1st I.E. = 519 kJ mol -1 1st I.E. = 494 kJ mol -1

Lithium's outer electron is in the second level, and only has the 1s 2 electrons to screen it. The 2s 1 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]3d n4s2 (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 orbita are slightly closer to the nucleus than the 4s - and so offer some ls screening. You will find this commented on in the page aboutelectronic 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 el ectrons. The 3d electrons have some screening effect, and the extra proton and the extra 3d electron more or less cancel each othe r out as far as attraction from the centre of the atom is concerned. The rise at zinc is easy to explain. Cu Zn [Ar]3d104s1 [Ar]3d104s2 1st I.E. = 745 kJ mol -1 1st I.E. = 908 kJ mol -1

In each case, the electron is coming from the same orbital, with identical screening, but the zinc has one extra proton in the nucleus and so the attraction is greater. There will be a degree of repulsion between the paired up electrons in the 4s orbital, but in this cas e 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 ele ctronic 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 b able to find any reference to this anywhere. In fact, I haven't been een 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 who le 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 ener gy, 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 thereactions of Group 2 metals with water.

THE ATOMIC HYDROGEN EMISSION SPECTRUM

This page introduces the atomic hydrogen emission spectrum, showing how it arises from electron movements between energy leve ls within the atom. It also looks at how the spectrum can be used to find the ionisation energy of hydrogen.

What is an emission spectrum?


Observing hydrogen's emission spectrum A hydrogen discharge tube is a slim tube containing hydrogen gas at low pressure with an electrode at each end. If you put a high voltage across this (say, 5000 volts), the tube lights up with a bright pink glow. If the light is passed through a prism or diffraction grating, it is split into its various colours. What you would see is a small part of the hydrogen emission spectrum. Most of the spectrum is invisible to the eye because it is either in the infra -red or the ultra-violet. The photograph shows part of a hydrogen discharge tube on the left, and the three most easily seen lines in the visible part of the spectrum on the right. (Ignore the "smearing" - particularly to the left of the red line. This is c aused by flaws in the way the photograph was taken. See note below.)

Note: This photograph is by courtesy of Dr Rod Nave of the Department of Physics and Astronomy at Georgia State University, Atlanta The photograph comes from notes about the hydrogen . spectrum in his HyperPhysics pages on the University site. If you are interested in more than an introductory look at the subject, that is a good place togo.

Ideally the photo would show three clean spectral lines - dark blue, cyan and red. The red smearing which appears to the left of the red line, and other similar smearing (much more difficult to see) to the left of the other two lines probably comes, according to Dr Nave, from stray reflections in the setup, or possibly from flaws in the diffraction grating. Ihave chosen to use this photograph anyway because a) I think it is a stunning image, and b) it is the only one I have ever come across which includesa hydrogen discharge tube and its spectrum in the same image.

Extending hydrogen's emission spectrum into the UV and IR There is a lot more to the hydrogen spectrum than the three lines you can see with the naked eye. It is possible to detect pa tterns of lines in both the ultra-violet and infra-red regions of the spectrum as well. These fall into a number of "series" of lines named after the person who discovered them. The diagram below shows three of these series, but there are others in the infra-red to the left of the Paschen series shown in the diagram. The diagram is quite complicated, so we will look at it a bit at a time. Look first at the Lyman series on the right of the diagram - this is the most spread out one and easiest to see what is happening.

Note: The frequency scale is marked in PHz - that's petaHertz. You are familiar with prefixes like kilo (meaning a thousand or 103 times), and mega (meaning a million or 106 times). Peta means 1015 times. So a value like 3 PHz means 3 x 1015 Hz. If you are worried about "Hertz", it just means "cycles per second".

The Lyman series is a series of lines in the ultra-violet. Notice that the lines get closer and closer together as the frequency increases. Eventually, they get so close together that it becomes impossible to see them as anything other than a continuous spectrum. That's what t he shaded bit on the right-hand end of the series suggests. Then at one particular point, known as the series limit, the series stops. If you now look at the Balmer series or the Paschen series, you will see that the pattern is just the same, but the series ha ve become more compact. In the Balmer series, notice the position of the three visible lines from the photograph further up the page.

Complicating everything - frequency and wavelength You will often find the hydrogen spectrum drawn using wavelengths of light rather than frequencies. Unfortunately, because of the mathematical relationship between the frequency of light and its wavelength, you get two completely different views of the spectrum if you plot it against frequency or against wavelength. The relationship between frequency and wavelength The mathematical relationship is:

Rearranging this gives equations for either wavelength or frequency.

What this means is that there is an inverse relationship between the two - a high frequency means a low wavelength and vice versa.

Note: You will sometimes find frequency given the much more obvious symbol, f.

Drawing the hydrogen spectrum in terms of wavelength This is what the spectrum looks like if you plot it in terms of wavelength instead of frequency:

. . . and just to remind you what the spectrum in terms of frequency looks like:

Is this confusing? Well, I find it extremely confusing! So what do you do about it? For the rest of this page I shall only look at the spectrum plotted against frequency, because it is much easier to relate it to what is happening in the atom. Be aware that the spectrum looks different depending on how it is plotted, but, other than that, ignore the wavelength version unless it is obvious that your examiners want it. If you try to learn both versions, you are only going to get them muddled up!

Note: Syllabuses probably won't be very helpful about this. You need to look at past papers and mark schemes. If you are working towards a UK-based exam and don't have these things, you can find out how to get hold of them by going to thesyllabuses page.

Explaining hydrogen's emission spectrum


The Balmer and Rydberg Equations

By an amazing bit of mathematical insight, in 1885 Balmer came up with a simple formula for predicting the wavelength of any of the lines in what we now know as the Balmer series. Three years later, Rydberg generalised this so that it was possib le to work out the wavelengths of any of the lines in the hydrogen emission spectrum. What Rydberg came up with was:

RH is a constant known as the Rydberg constant. n1 and n2 are integers (whole numbers). n 2 has to be greater than n1. In other words, if n1 is, say, 2 then n2 can be any whole number between 3 and infinity. The various combinations of numbers that you can slot into this formula let you calculate the wavelength of any of the lines in the hydrogen emission spectrum - and there is close agreement between the wavelengths that you get using this formula and those found by analysing a real spectrum.

Note: If you come across a version of Balmer's original equation, it won't look like this. In Balmer's equation, n is always 2 - because that gives the wavelengths of the lines in the visible part 1 of the spectrum which is what he was interested in. His original equation was also organised differently. The modern versionshows more clearly what is going on.

You can also use a modified version of the Rydberg equation to calculate the frequency of each of the lines. You can work out this version from the previous equation and the formula relating wavelength and frequency further up the page.

Note: You may come across versions of the Rydberg equation where the n1 and n 2 are the other way around, or they may even be swapped for letters like m and n. Whichever version you use, the bigger number must always be the one at the bottom of the right-hand term - the one you take away. If you get them the wrong way around, it is immediately obvious if you start to do a calculation, because you will end up with a negative answer!

The origin of the hydrogen emission spectrum The lines in the hydrogen emission spectrum form regular patterns and can be represented by a (relatively) simple equation. Each line can be calculated from a combination of simple whole numbers. Why does hydrogen emit light when it is excited by being exposed to a high voltage and what is the significance of those whole numbers? When nothing is exciting it, hydrogen's electron is in the first energy level - the level closest to the nucleus. But if you supply energy to the atom, the electron gets excited into a higher energy level - or even removed from the atom altogether. The high voltage in a discharge tube provides that energy. Hydrogen molecules are first broken up into hydrogen atoms (hence the atomic hydrogen emission spectrum) and electrons are then promoted into higher energy levels. Suppose a particular electron was excited into the third energy level. This would tend to lose energy again by falling back d own to a lower level. It could do this in two different ways. It could fall all the way back down to the first level again, or it could fall back to the second level - and then, in a second jump, down to the first level.

Tying particular electron jumps to individual lines in the spectrum If an electron falls from the 3-level to the 2-level, it has to lose an amount of energy exactly the same as the energy gap between those two levels. That energy which the electron loses comes out as light (where "light" includes UV and IR as well as visible). Each frequency of light is associated with a particular energy by the equation:

The higher the frequency, the higher the energy of the light. If an electron falls from the 3-level to the 2-level, red light is seen. This is the origin of the red line in the hydrogen spectrum. By measuring the frequency of the red light, you can work out its energy. That energy must be exactly the same as the energy gap between the 3-level and the 2-level in the hydrogen atom. The last equation can therefore be re-written as a measure of the energy gap between two electron levels.

The greatest possible fall in energy will therefore produce the highest frequency line in the spectrum. The greatest fall will be from the infinity level to the 1-level. (The significance of the infinity level will be made clear later.)

The next few diagrams are in two parts - with the energy levels at the top and the spectrum at the bottom.

If an electron fell from the 6-level, the fall is a little bit less, and so the frequency will be a little bit lower. (Because of the scale of the diagram, it is impossible to draw in all the jumps involving all the levels between 7 and infinity!)

. . . and as you work your way through the other possible jumps to the 1 -level, you have accounted for the whole of the Lyman series. The spacings between the lines in the spectrum reflect th e way the spacings between the energy levels change.

If you do the same thing for jumps down to the 2-level, you end up with the lines in the Balmer series. These energy gaps are all much smaller than in the Lyman series, and so the frequencies produced are also much lower.

The Paschen series would be produced by jumps down to the 3-level, but the diagram is going to get very messy if I include those as well - not to mention all the other series with jumps down to the 4-level, the 5-level and so on.

The significance of the numbers in the Rydberg equation n1 and n2 in the Rydberg equation are simply the energy levels at either end of the jump producing a particular line in the spectrum.

For example, in the Lyman series, n 1 is always 1. Electrons are falling to the 1 -level to produce lines in the Lyman series. For the Balmer series, n 1 is always 2, because electrons are falling to the 2-level. n2 is the level being jumped from. We have already mentioned that the red line is produced by electrons falling from the 3-level to the 2-level. In this case, then, n2 is equal to 3.

The significance of the infinity level The infinity level represents the highest possible energy an electron can have as a part of a hydrogen atom. So what happens if the electron exceeds that energy by even the tiniest bit? The electron is no longer a part of the atom. The infinity level represents the point at which ionisation of the atom occurs to form a positively charged ion.

Using the spectrum to find hydrogen's ionisation energy


When there is no additional energy supplied to it, hydrogen's electron is found at the 1 -level. This is known as its ground state. If you supply enough energy to move the electron up to the infinity level, you have ionised the hydrogen. The ionisation energy per electron is therefore a measure of the distance between the 1-level and the infinity level. If you look back at the last few diagrams, you will find that that particular energy jump produces the series limit of the Lyman series.

Note: Up to now we have been talking about the energy released when an electron falls from a higher to a lower level. Obviously ifa certain amount of energy is released when an electron falls from the infinity level to the 1-level, that same amount will beneeded to push the electron from the 1-level up to the infinity level.

If you can determine the frequency of the Lyman series limit, you can use it to calculate the energy needed to move the elect ron in one atom from the 1-level to the point of ionisation. From that, you can calculate the ionisation energy per mole of atoms. The problem is that the frequency of a series limit is quite difficult to find accurately from a spectrum because the lines a re so close together in that region that the spectrum looks continuous.

Finding the frequency of the series limit graphically Here is a list of the frequencies of the seven most widely spaced lines in the Lyman series, together with the increase in fr equency as you go from one to the next.

As the lines get closer together, obviously the increase in frequency gets less. At the series limit, the gap between the lines would be literally zero.

That means that if you were to plot the increases in frequency against the actual frequency, you could extrapolate (contin ue) the curve to the point at which the increase becomes zero. That would be the frequency of the series limit. In fact you can actually plot two graphs from the data in the table above. The frequency difference is related to two frequencies. For example, the figure of 0.457 is found by taking 2.467 away from 2.924. So which of these two values should you plot the 0.457 against? It doesn't matter, as long as you are always consistent - in other words, as long as you always plot the difference against either the higher or the lower figure. At the point you are interested in (where the difference becomes zero), the two frequency numbers are the same. As you will see from the graph below, by plotting both of the possible curves on the same graph, it makes it ea sier to decide exactly how to extrapolate the curves. Because these are curves, they are much more difficult to extrapolate than if they were straight line s.

Both lines point to a series limit at about 3.28 x 10 15 Hz.

Note: Remember that 3.28 PHz is the same as 3.28 x 1015 Hz. You can use the Rydberg equation to calculate the series limit of the Lyman series as a check on this figure: n = 1 for the Lyman 1 2 series, and n 2 = infinity for the series limit. 1/(infinity) = zero. That gives a value for the frequency of 3.29 x 1015 Hz - in other words the two values agree to within 0.3%.

So . . . now we can calculate the energy needed to remove a single electron from a hydrogen atom. Remember the equation from higher up the page:

We can work out the energy gap between the ground state and the point at which the electron leaves the atom by substituting the value we've got f or frequency and looking up the value of Planck's constant from a data book.

That gives you the ionisation energy for a single a tom. To find the normally quoted ionisation energy, we need to multiply this by the number of atoms in a mole of hydrogen atoms (the Avogadro constant) and then divide by 1000 to convert it into kilojoules.

Note: It would be wrong to quote this to more than 3 significant figures. The value for the frequency obtained from the graph is only to that accuracy.

This compares well with the normally quoted value for hydrogen's ionisation energy of 1312 kJ mol -1.

ELECTRON AFFINITY

This page explains what electron affinity is, and then looks at the factors that affect its size. It assumes that you know ab out simple atomic orbitals, and can write electronic structures for simple atoms.

Important! If you aren't reasonable happy about atomic orbitals and electronic structures you should follow these links before you go any further.

First electron affinity


Ionisation energies are always concerned with the formation of positive ions. Electron affinities are the negative ion equiva lent, and their use is almost always confined to elements in groups 6 and 7 of the Periodic Table. Defining first electron affinity The first electron affinity is the energy released when 1 mole of gaseous atoms each acquire an electron to form 1 mole of ga seous 1- ions. This is more easily seen in symbol terms.

It is the energy released (per mole of X) when this change happens. First electron affinities have negative values. For example, the first electron affinity of chlorine is -349 kJ mol-1. By convention, the negative sign shows a release of energy.

The first electron affinities of the group 7 elements F Cl Br I -328 kJ mol-1 -349 kJ mol-1 -324 kJ mol-1 -295 kJ mol-1

Note: These values are based on the most recent research. If you are using a different data source, you may have slightly different numbers. That doesn't matter - the pattern will still be the same.

Is there a pattern? Yes - as you go down the group, first electron affinities become less (in the sense that less energy is evolved when the negative i ons are formed). Fluorine breaks that pattern, and will have to be accounted for separately. The electron affinity is a measure of the attraction between the incoming electron and the nucleus - the stronger the attraction, the more energy is released. The factors which affect this attraction are exactly the same as those relating to ionisation energies - nuclear charge, distance and screening.

Note: If you haven't read about ionisation energy recently, it might be a good idea to follow this link before you go on. These factors are discussed in more detail on that page than the are on y this one.

The increased nuclear charge as you go down the group is offset by extra screening electrons. Each outer electron in effect feels a pull of 7+ from the centre of the atom, irrespective of which element you are talking about. For example, a fluorine atom has an electronic structure of 1s 22s22px2 2py22pz1. It has 9 protons in the nucleus. The incoming electron enters the 2-level, and is screened from the nucleus by the two 1s 2 electrons. It therefore feels a net attraction from the nucleus of 7+ (9 protons less the 2 screening electrons). By contrast, chlorine has the electronic structure 1s 22s22p63s23px23py 23pz1. It has 17 protons in the nucleus. But again the incoming electron feels a net attraction from the nucleus of 7+ (17 protons less the 10 screening electrons in the first and second levels).

Note: If you want to be fussy, there is also a small amount ofscreening by the 2s electrons in fluorine and by the 3s electrons in chlorine. This will be approximately the same in both th ese cases and so doesn't affect the argument in any way (apart from complicating it!).

The over-riding factor is therefore the increased distance that the incoming electron finds itself from the nucleus as you go down the group. The greater the distance, the less the attraction and so the less energy is released as electron affinity.

Note: Comparing fluorine and chlorine isn't ideal, because fluorine breaks the trend in the group. However, comparing chlorine and bromine, say, makes things seem more difficult because of the more complicated electronic structures involved. What we have said so far is perfectly true and applies to the fluorine-chlorine case as much as to anything else in the group, but there's another factor which operates as well which we haven't considered yet - and that over-rides the effect of distance in the case of fluorine.

Why is fluorine out of line? The incoming electron is going to be closer to the nucleus in fluorine than in any other of these elements, so you would expect a high value of electron affinity. However, because fluorine is such a small atom, you are putting the new electron into a region o f space already crowded with electrons and there is a significant amount of repulsion. This repulsion lessens the attraction the incoming electron feels and so lessens the elect ron affinity. A similar reversal of the expected trend happens between oxygen and sulphur in Group 6. The first electron affinity of oxygen (-142 kJ mol-1) is smaller than that of sulphur (-200 kJ mol-1) for exactly the same reason that fluorine's is smaller than chlori ne's.

Comparing Group 6 and Group 7 values As you might have noticed, the first electron affinity of oxygen ( -142 kJ mol -1) is less than that of fluorine (-328 kJ mol -1). Similarly sulphur's (-200 kJ mol-1) is less than chlorine's (-349 kJ mol -1). Why?

It's simply that the Group 6 element has 1 less proton in the nucleus than its next door neighbour in Group 7. The amount of sc reening is the same in both. That means that the net pull from the nucleus is less in Group 6 than in Group 7, and so the electron affinities are less.

First electron affinity and reactivity The reactivity of the elements in group 7 falls as you go down the group - fluorine is the most reactive and iodine the least. Often in their reactions these elements form their negative ions. At GCSE the impression is sometimes given that the fall in reactivity is because the incoming electron is held less strongly as you go down the group and so the negative ion is less likely to form. That explana tion looks reasonable until you include fluorine! An overall reaction will be made up of lots of different steps all involving energy changes, and you cannot safely try to exp lain a trend in terms of just one of those steps. Fluorine is much more reactive than chlorine (despite the lower electron affini ty) because the energy released in other steps in its reactions more than makes up for the lower amount of energy released as electron affinity.

Second electron affinity


You are only ever likely to meet this with respect to the group 6 elements oxygen and sulphur which both form 2 - ions. Defining second electron affinity The second electron affinity is the energy required to add an electron to each ion in 1 mole of gaseous 1- ions to produce 1 mole of gaseous 2- ions. This is more easily seen in symbol terms.

It is the energy needed to carry out this change per mole of X -. Why is energy needed to do this? You are forcing an electron into an already negative ion. It's not going to go in willingly! 1st EA = -142 kJ mol-1 2nd EA = +844 kJ mol-1 The positive sign shows that you have to put in energy to perform this change. The second electron affinity of oxygen is part icularly high because the electron is being forced into a small, very electron-dense space.

ATOMIC AND IONIC RADIUS

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

Important! If you aren't reasonable happy about electronic structures you should follow this link before you go any further.

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

As you can see from the diagrams, the same atom could be found to have a different radius depending on what was around it. The left hand diagram shows bonded atoms. The atoms are pulled closely together and so the measured radius is less than if th ey are just touching. This is what you would get if you had metal atoms in a metallic structure, or atoms covalently bonded to each other. The type of atomic radius being measured here is called the metallic radius or the covalent radius depending on the bonding. The right hand diagram shows what happens if the atoms are jus t touching. The attractive forces are much less, and the atoms are essentially "unsquashed". This measure of atomic radius is called the van der Waals radius after the weak attractions present in this situation.
Note: If you want to explore these various types of bonding this link will take you to the bonding menu.

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

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

Leaving the noble gases out, atoms get smaller as you go across a period.

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

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

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

Trends in the transition elements

Although there is a slight contraction at the beginning of the series, the atoms are all much the same size. The size is determined by the 4s electrons. The pull of the increasing number of protons in the nucleus is more or less offset by the extra screening due to the increasing number of 3d electrons.

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

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

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

electronic structure of ion Li+ Na+ K+ Rb+ 2 2, 8 2, 8, 8 2, 8, 18, 8

ionic radius (nm)

0.076 0.102 0.138 0.152

Cs+

2, 8, 18, 18, 8

0.167

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

ionic radius (nm)

0.133 0.181 0.196 0.220

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

Trends in ionic radius across a period Let's look at the radii of the simple ions formed by elements as you go across Period 3 of the Periodic Table - the elements from Na to Cl.

Na+ no of protons electronic structure of ion ionic radius (nm) 11 2,8 0.102

Mg2+ 12 2,8 0.072

Al3+ 13 2,8 0.054

P315 2,8,8 (0.212)

S216 2,8,8 0.184

Cl17 2,8,8 0.181

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

First of all, notice the big jump in ionic radius as soon as you get into the negative ions. Is this surprising? Not at all - you have just added a whole extra layer of electrons. Notice that, within the series of positive ions, and the series of negative ions, that the ionic radii fall as you go across the per iod. We need to look at the positive and negative ions separately. The positive ions In each case, the ions have exactly the same electronic structure - they are said to be isoelectronic. However, the number of protons in the nucleus of the ions is increasing. That will tend to pull the electrons more and more towards the centre of the ion - causing the ionic radii to fall. That is pretty obvious! The negative ions Exactly the same thing is happening here, except that you have an extra layer of electrons. What needs commenting on, though is how similar in size the sulphide ion and the chloride ion are. The additional proton here is making hardly any difference.

The difference between the size of similar pairs of ions actually gets even smaller as you go down Groups 6 and 7. For exampl e, the Te2- ion is only 0.001 nm bigger than the I - ion. As far as I am aware there is no simple explanation for this - certainly not one which can be used at this level. This is a good illustration of what I said earlier - explaining things involving ionic radii in detail is sometimes very difficult.

Trends in ionic radius for some more isoelectroni c ions This is only really a variation on what we have just been talking about, but fits negative and positive isoelectronic ions in to the same series of results. Remember that isoelectronic ions all have exactly the same electron arrangement.

N3no of protons electronic structure of ion ionic radius (nm) 7 2, 8 (0.171)

O28 2, 8 0.140

F9 2, 8 0.133

Na+ 11 2, 8 0.102

Mg2+ 12 2, 8 0.072

Al3+ 13 2, 8 0.054

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

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

The relative sizes of ions and atoms You probably won't have noticed, but nowhere in what you have read so far has there been any need to talk about the relative sizes of the ions and the atoms they have come from. Neither (as far as I can tell from the syllabuses) do any of the current UK -based exams for 16 - 18 year olds ask for this specifically in their syllabuses. However, it is very common to find statements about the relative sizes of ions and atoms. I am fairly convinced that these st atements are faulty, and I would like to attack the problem head-on rather than just ignoring it. Important! For 10 years, until I rewrote this ionic radius section in August 2010, I included what is in the box below. You will find th is same information and explanation in all sorts of books and on any number of websites aimed at this level. At least one non-UK A level syllabus has a statement which specifically asks for this.

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

Positive ions Positive ions are smaller than the atoms they come from. Sodium is 2,8,1; Na + is 2,8. You've lost a whole layer of electrons, and the remaining 10 electrons are being pulled in by the full force of 11 protons.

Negative ions Negative ions are bigger than the atoms they come from. Chlorine is 2,8,7; Cl - is 2,8,8. Although the electrons are still all in the 3 -level, the extra repulsion produced by the incoming electron causes the atom to expand. There are still only 17 protons, but they are now havi ng to hold 18 electrons.

However, I was challenged by an experienced teacher about the negative ion explanation, and that forced me to think about it carefully for the first time. I am now convinced that the facts and the explanation relating to negative ions are simply illogical. As far as I can tell, no UK-based syllabus mentions the relative sizes of atoms and ions (as of August 2010), but you should check past papers and mark schemes to see whether questions have sneaked in. The rest of this page discusses the problems that I can see, and is really aimed at teachers and others, rather than at students. If you are a student, look carefully at your syllabus, and past exam questions and mark schemes, to find out whether you need to know about this. If you don't need to know about it, stop reading now (unless, of course, you are interested in a bit of controversy!). If you do need to know it, then you will have to learn what is in the box, even if, as I believe, it is wrong. If you like your chemistry to be simple, ignore the rest of the page, because you risk getting confused about what you need to know. If you have expert knowledge of this topic, and can find any flaws in what I am saying, then please contact me via the address on the about this site page.

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

Group 7

vdW radius (nm)

ionic radius of X - (nm)

F Cl Br I

0.147 0.175 0.185 0.198

0.133 0.181 0.196 0.220

Group 6

vdW radius (nm) O S Se Te 0.152 0.180 0.190 0.206

ionic radius of X 2- (nm) 0.140 0.184 0.198 0.221

Group 5

vdW radius (nm) N P 0.155 0.180

ionic radius of X3- (nm) 0.171 0.212

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

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

And what is it safe to say about the explanation? If there are any additional electron-electron repulsions on adding extra electrons, they must be fairly small. This is particularly shown if you consider some pairs of isoelectronic ions.

You would have thought that if repulsion was an important factor, then the radius of, say a sulphide ion, with two negative c harges would be significantly larger than a chloride ion with only one. The difference should actu ally be even more marked, because the sulphide electrons are being held by only 16 protons rather than the 17 in the chlorine case. On this repulsion theory, the sulphide ion shouldn't just be a little bit bigger than a chloride ion - it should be a lot bigger. The same effect is shown with selenide and bromide, and with telluride and iodide ions. In the last case, there is virtually no difference in the size s of the 2- and 1- ions. So if there is some repulsion playing a part in this, it certainly doesn't look as if it is playing a major part.

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

Conclusion It seems to me that, for negative ions, it is compl etely illogical to compare ionic radii with covalent radii if you want to use the electron repulsion explanation. If you compare the ionic radii of negative ions with the van der Waals radii of the atoms they come from, the uncertainties i n the data make it very difficult to make any reliable generalisations. The similarity in sizes of pairs of isoelectronic ions from Groups 6 and 7 calls into question how important repulsion is in any explanation. Having spent more than a week working on this, and discussi ng it with input from some very knowledgable people, I don't think there is any explanation which is simple enough to give to most students at this level. It would seem to me to be better that these ideas about relative sizes of atoms and ions are just dropped. At this level, you can describe and explain simple periodic trends in atomic radii in the way I did further up this page, wit hout even thinking about the relative sizes of the atoms and ions. Personally, I would be more than happy never to think abou t this again for the rest of my life!

BONDING MENU

Ionic bonding . . . Includes a simple view of ionic bonding and the way you need to modify this for A'level purposes. Covalent bonding . . . Includes a simple view of covalent bonding (single and double) and the modifications needed for A'level purposes. Co-ordinate (dative covalent) bonding . . . Explains what co-ordinate (dative covalent) bonding is, and looks at a wide range of examples. Electronegativity . . . Explains what electronegativity is and how it varies around the Periodic Table. Describes and explains how electronegativity differences determine the type of bond formed. Looks at polar bonds an d molecules. Shapes of simple molecules and ions . . . Explains how to work out the shapes of a wide range of simple molecules and ions.

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

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

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). Thes e 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 the ir outer levels are either completely full or completely empty.

Note: The central role given to noble gas structures is very much an over-simplification. We shall have to spend some time later on demolishing the concept!

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

The sodium has lost an electron, so it no longer has equal numbers of electrons and protons. Because it has one more proton t han 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 th erefore 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 therefo re 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 K 2O.

THE A'LEVEL VIEW OF IONIC BONDING y y


Electrons are transferred from one atom to another resulting in the formatio n 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 , b ut not one of them has a noble gas structure. Fe3+ Cu2+ Zn2+ Ag+ Pb2+ Noble gases (apart from helium) have an outer electronic structure ns 2np6. [Ar]3d5 [Ar]3d9 [Ar]3d10 [Kr]4d10 [Xe]4f 145d106s2

Note: If you aren't happy about writing electronic structures using of s, p and d notation, follow this link before you go on. Return to this page via the menus or by using the BACK button on your browser.

Apart from some elements at the beginning of a transition series (scandium forming Sc 3+ 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 l ike those above. That means that the only elements to form positive ions with noble gas structures (apart from odd ones like scandium) are tho se 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 int rest only. e

What determines what the charge is on an ion? Elements combine to make the compound which is as stable as possible - the one in which the greatest amount of energy is evolved in its making. The more charges a positive ion has, the greater the attraction towards its accompanying negative ion. The greater the attraction, the more energy is released when the ions come together. That means that elements forming positive ions will tend to give away as many electrons as possible. But there's a down -side to this. Energy is needed to remove electrons from atoms. This is called ionisation energy. The more electrons you remove, the greater the total ionisation energy becomes. Eventually the total ionisation energy needed becomes so great that the energy released when the attractions are set up between positive and negative ions isn't large enough to cover it. 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 mor e stable, which is why you get CaCl 2 rather than CaCl. What about CaCl 3 (containing Ca3+ ions)? To make one mole of this, you can estimate that you would have to put in 1341 kJ. This makes this compound completely non-viable. Why is so much heat needed to make CaCl3? It is because the third ionisation energy (the energy needed to remove the third electron) is extremely high (4940 kJ mol -1) because the electron is being removed from the 3-level rather than the 4-level. Because it is much closer to the nucleu s 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 O containing the O2- ion turn out to be the most energetically stable.

2-

ion rather than an O- ion or an O3- ion, because compounds

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 i t 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 d ouble 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 w hich 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 fr om 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 s tructures 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 tha t 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 thi ng 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 n oble 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 sim ple 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 a fter 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 BF 3, 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 PCl 3. The diagram of PCl 5 (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 o ff 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 n to read about orbitals before you go on. eed 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-andcrossed picture of methane looks like this.

There is a serious mis-match between this structure and the modern electronic structure of carbon, 1s 22s22px12py1. 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 1s2 electrons are too deep inside the atom to be involved in bonding. The only electrons directly available for sharing are the 2p electrons. Why then isn't methane CH2? Promotion of an electron 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 energ y 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 ide ntical orbitals. Hybridisation The electrons rearrange themselves again in a process called hybridisation. This reorganises the electrons into four identical hybrid orbitals called sp 3 hybrids (because they are made from one s orbital and three p orbitals). You should read "sp3" as "s p three" - not as "s p cubed".

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 possi ble. You can picture the nucleus as being at the centre of a tetr ahedron (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 sp 3 hybrids, but with a hydrogen nucleus embedded in each lobe. Each orbital holds the 2 electrons that we've previously drawn as a d ot 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 onmethane 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 PCl 3?

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 1s 22s22p63s23px13py13pz 1. 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 sp 3 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 n ot! It is worth it with PCl5, though. What's wrong with the simple view of PCl 5? 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 tha n 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 stru cture, 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 compoun ds. In fact, it doesn't. For example, the compound NCl3 exists, but there is no such thing as NCl 5. Nitrogen is 1s22s22px12py 12pz1. 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.

CO-ORDINATE (DATIVE COVALENT) BONDING

This page explains what co-ordinate (also called dative covalent) bonding is. You need to have a reasonable unde rstanding of simple covalent bonding before you start.

Important! If you are uncertain about covalent bonding follow this link before you go on with this page.

Co-ordinate (dative covalent) bonding

A covalent bond is formed by two atoms sharing a pair of electrons. The atoms are held together because the electron pair is attracted by both of the nuclei. In the formation of a simple covalent bond, each atom supplies one electron to the bond - but that doesn't have to be the case. A co-ordinate bond (also called a dative covalent bond) is a covalent bond (a shared pair of electrons) in which both electrons come from the sa me 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, NH 4+, is formed, the fourth hydrogen is attached by a dative covalent bond, bec ause 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 chlo ride 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 te st 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 hyd rogen 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 co-ordinate bonds drawn. The nitrogen end of the bond has become positive because the electron pair has moved away from the nitrogen towards the boron - which has therefore become negative. We shan't use this method again - it's more confusing than just using an arrow.

The structure of aluminium chlorid e 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 temper ature is not AlCl3, but Al2Cl6. It exists as a dimer (two mo lecules 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 3 molecules.

Note: Aluminium chloride is complicated because of the way it keeps changing its bonding as the temperature increases. If yo are interested in exploring this in more detail, you could have a u 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 mole cules 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 1s 22s22p63s23px1. When it forms an Al 3+ ion it loses the 3-level electrons to leave 1s 22s22p6. 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 acce pt 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 ma king 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 invo lved 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 alumin ium, 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, an d so the arrangement of the electrons must be identical. There is no way of showing this using a dots-and-crosses picture. The bonding involves delocalisation.

If you are interested: The bonding is rather similar to the bonding in the ethanoate ion (alt ough without the negative charge). You will find thisdescribed on a page about theacidity of h organic acids.

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 chemis try 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 valu es 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 o f bond in, for example, H2 or Cl2 molecules.

Note: It's important to realise that this is anaverage 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 hydrog en-chlorine bond in HCl or the hydrogen-oxygen bonds in water are typical.

What happens if B is a lot more electronegative than A? In this case, the electron pair is dragged right over to B's end of the bond. To all intents and purposes, A has lost control of its electron, and B has complete control over both electrons. Ions have been formed.

A "spectrum" of bonds The implication of all this is that there is no clear -cut division between covalent and ionic bonds. In a pure covalent bond, the electrons are held on average exactly half way between the atoms. In a polar bond, the electrons have been dragged slightly towards one end. How far does this dragging have to go before the bond counts as ionic? There is no real answer to that. You normally think of sodium chloride as being a typically ionic solid, but even here the sodium hasn't completely lost control of its electron. Because of the properties of sodium chloride, however, we tend to count it as if it were purely ionic.

Note: Don't worry too much about the exact cut-off point between polar covalent bonds and ionic bonds. At A'level, examples will tend to avoid the grey areas- they will be obviously covalent or obviously ionic. You will, however, be expected to realise that those grey areas exist.

Lithium iodide, on the other hand, would be described as being "ionic with some covalent character". In this case, the pair o f electrons hasn't moved entirely over to the iodine end of the bond. Lithium iodide, for example, dissolves in organic solvents like ethanol - not something which ionic substances normally do.

Summary

y y y

No electronegativity difference between two atoms leads to a pure non -polar covalent bond. A small electronegativity difference leads to a polar covalen t 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 to p to bottom, or from left to right. By contrast, CHCl3 is polar.

The hydrogen at the top of the molecule is less electronegative than carbon and so is slightly positive. This means that the molecule now has a slightly positive "top" and a slightly negative "bottom", and so is overall a polar molecule.

A polar molecule will need to be "lop-sided" in some way.

Patterns of electronegativity in the Periodic Table


The most electronegative element is fluorine. If you remember that fact, everything becomes easy, because electronegativity must always increase towards fluorine in the Periodic Table.

Note: This simplification ignores the noble gases. Historically this is because they were believed not to form bonds- and if they don't form bonds, they can't have an electronegativity value. Even now that we know that some of them do form bonds, data sources still don't quote electronegativity values for them.

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

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

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

y y y

the number of protons in the nucleus; the distance from the nucleus; the amount of screening by inner electrons.

Note: If you aren't happy about the concept of screening or shielding, it would pay you to read the page on ionisation energies before you go on. The factors influencing ionisation energies are just the same as those influencing electronegativities. Use the BACK button on your browser to return to this page.

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 cha rges 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 o nly by the 1s2 electrons. In the chlorine case it is shielded by all the 1s 22s22p6 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 at traction 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 remove from their new d syllabus. At the time of writing, it does, however, still appear on at least one overseas Alevel 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 a polar bond. Reading it unnecessarily just risks confusing you. of

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 conta in Al3+ and Cl- ions. The aluminium ion is very small and is packed with three positive charges - the "charge density" is therefore very high. That will have a considerable effect on any nearby electrons.

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

Factors affecting polarising ability Positive ions can have the effect of polarising (electrically distorting) nearby negative ions. The polarising ability depend s 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 fluorid e is ionic because the aluminium ion can't polarise the small fluoride ion sufficiently to form a covalent bond.

SHAPES OF MOLECU LES AND IONS

This page explains how to work out the shapes of molecules and ions containing only single bonds. If you are interested in th e shapes of molecules and ions containing double bonds, you will find a link at the bottom of the page. The examples on this page are all simple in the sense that they only contain two sorts of atoms joined by single bonds - for example, ammonia only contains a nitrogen atom joined to three hydrogen atoms by single bonds. If you are given a more complicated example, look carefully at the arrangement of the atoms before you start to make sure that there are only single bonds present. For example, if you had a molecule such as COCl 2, you would need to work out its structure, based on the fact that you know that carbon forms 4 covalent bonds, oxygen 2, and chlorine (normally) 1. If you did that, you would find that the carbon is joined to the oxygen by a double bond, and to the two chlorines by single bonds. That means that you couldn't use the techniques on this page, because this page only considers single bonds.

The electron pair repulsion theory


The shape of a molecule or ion is governed by the arrangement of the electron pairs around the central atom. All you need to do is to work out how many electron pairs there are at the bonding level, and then arrange them to produce the minimum amount of repulsion between them. You have to include both bonding pairs and lone pairs. How to work out the number of electron pairs You can do this by drawing dots-and-crosses pictures, or by working out the structures of the atoms using electrons -in-boxes and worrying about promotion, hybridisation and so on. But this is all very tedious! You can ge t exactly the same information in a much quicker and easier way for the examples you will meet if you are doing one of the UK-based exams for 16 - 18 year olds.

Warning: This method won't work without some modification for many ions containing metals, a no simple method gives reliable results where the central atom is a transition metal. The nd method will, however, cope with all the substances that you are likely to meet in this section of the syllabus. When you dealwith 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 wantsyou to know. It is important to know exactly which molecules and ions your syllabus expects you to be able to work out the shapes forin this part of the syllabus. You should also check past exam papers. If you are working to a UK-based syllabus for 16 - 18 year olds, and haven't got copies of your syllabus and past papers follow this link to find out how to get them.

First you need to work out how many electrons there are around the central atom:

y y y

Write down the number of electrons in the outer level of the central atom. That will be the same as the Periodic Table group numb er, except in the case of the noble gases which form compounds, when it will be 8. Add one electron for each bond being formed. (This allows for the electrons coming from the other atoms.) Allow for any ion charge. For example, if the ion has a 1 - charge, add one more electron. For a 1+ charge, deduct an electron.

Now work out how many bonding pairs and lone pairs of electrons there are:

y y

Divide by 2 to find the total number of electron pairs around the central atom. Work out how many of these are bonding pairs, and how many are lone pairs. You know how many bonding pairs there are because you know how many other atoms are joined to the central atom (assuming that only single bonds are formed). For example, if you have 4 pairs of electrons but only 3 bonds, there must be 1 lone pair as well as the 3 bonding pairs.

Finally, you have to use this information to work out the shape:

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

Three electron pairs around the central atom The simple cases of this would be 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 electro ns - in 3 pairs. Because it is forming 3 bonds there can be no lone pairs. The 3 pairs arrange themselves as far apart as possible. They all lie in one plane at 120 to each other. The arrangement is called trigonal planar.

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 doingit the quick and easy way! If you are interested in the bonding in methane you can find it in the organic section by following this link, or in a page oncovalent bonding by following this one.

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

Note: It is important that you understand the use of various sorts of line to show the 3-dimensional arrangement of the bonds. In diagrams of this sort, an ordinary line represents a bond in the plane of the screen or paper. A dotted line shows a bond going away from you into the screen or paper. A wedge shows a bond coming out towards you. It is my habit to draw diagrams like this with the bond at the top in the plane of the paper, the middle bond at the bottom oming out towards you, and the other two going back in. But that's all c 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 abou half-way down the page about drawing organic molecules. t 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 out er 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 p air. 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 descr ibe the shape, you only take notice of the atoms. Ammonia is pyramidal - like a pyramid with the three hydrogens at the base and the nitrogen at the top.

Water, H 2O

Following the same logic as before, you will find that the oxygen has four pairs of electrons, two of which are lone pairs. T hese 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 Vshaped.

The ammonium ion, NH 4+ 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 12 0 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, ClF 3 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 th an 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 l one 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 fe els negligible repulsion from the lone pairs. The bond to the fluorine in the plane i s 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 on on this page. e

Six electron pairs around the central atom A simple example: SF 6 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, plu s 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. X eF4 is described as square planar.

ClF4Chlorine is in group 7 and so has 7 outer electrons. Plus the 4 from the f our 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.

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 sodiu m (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 1s 22s22p63s1. When sodium atoms come together, the electron in the 3s atomic orbital of o ne 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 pi ece of metal. There have to be huge numbers of molecular orbitals, of cou rse, 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. Th e 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 me tal 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) a nd 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 greate r 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 increa se 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 3 d 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.

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 individu al molecule together (for example, the covalent bond s) 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, H 2, if you slow the molecules down by cooling the gas, the attractions are large enough for the molecules to stick together eventually t o 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 fo rces" (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 electric al 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 Br 2. 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 prop erly 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 p olarity 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 the other. - end of

An instant later the electrons in the left hand molecule may well have moved up the other end. In doing so, they will repel t he 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 e xact value, because the size of the attraction varies considerably with the size of the molecule and its shape. How molecular size affects the strength of the dispersion forces The boiling points of the noble gases are helium neon argon krypton xenon radon 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 d oes the radius of the atom. The more electrons you have, and the more distance over which they can move, the bigger the possible te mporary dipoles and therefore the bigger the dispersion forces. -269C -246C -186C -152C -108C -62C

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 mo lecules 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 th an short fat ones containing the same numbers of electrons. Long thin molecules can also lie closer together - these attractions are at their most effective if the molecules are really close. For example, the hydrocarbon molecules butane and 2-methylpropane both have a molecular formula 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 molec ules 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 e thane, 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 th at 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 electroneg ativity 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. CHCl3, is a highly polar molecule because of the electronegativity of the three chlorines. strong dipole-dipole attractions between one molecule and its neighbours.

Trichloromethane, There will be quite

On the other hand, tetrachloromethane, CCl 4, is non-polar. The outside of the molecule is uniformly dispersion forces.

- in all directions. CCl 4 has to rely only on

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 dipole-dipole interactions. The boiling points are: CHCl3 CCl4 61.2C 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 a bove 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!

INTERMOLECULAR BONDING - HYDROGEN BONDS

This page explains the origin of hydrogen bonding - a relatively strong form of intermolecular attraction. If you are also interested in the weaker intermolecular forces (van der Waals dispersion forces and dipole -dipole interactions), there is a link at the bottom of the page.

The evidence for hydrogen bonding


Many elements form compounds with hydrogen. If you plot the boiling points of the compounds of the Group 4 elements with hydrogen, you find that the boiling points increase as you go down the group.

The increase in boiling point happens because the molecules are getting larger with more electrons, and so van der Waals dispersion forces become greater.

Note: If you aren't sure about van der Waals dispersion forces, it would pay you to follow this link before you go on.

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

Although for the most part the trend is exactly the same as in group 4 (for exactly the same reasons), the boiling point of t he 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 intermolecula r 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:

y y

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 ha s a high density of negative charge. Lone pairs at higher levels are more diffuse and not so attractive to positive things.

Note: If you aren't happy about electronegativity, you should follow this link before you go on.

Consider two water molecules coming close together.

The + hydrogen is so strongly attracted to the lone pair that it is almost as if you were beginning to form a co -ordinate (dative covalent) bond. It doesn't go that far, but the attraction is significantly stronger than an ordinary dipole-dipole interaction. Hydrogen bonds have about a tenth of the strength of an average covalent bond, and are being constantly broken and reformed i n 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 bon ding 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 cou ld 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 onmolecular 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 hydrati on. 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 toco-ordinate (dative covalent) bonding.

The diagram shows the potential hydrogen bonds formed to a chloride ion, Cl -. Although the lone pairs in the chloride ion are at the 3-level and wouldn't normally be active enough to form hydrogen bonds, in this case they are made more attractive by the fu ll 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 hy drogen 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 mole cules will always have higher boiling points than similarly sized molecules which don't have an -O-H or an -N-H group. The hydrogen bonding makes the molecules "stickier", and more heat is necessary to separate them. Ethanol, CH3CH2-O-H, and methoxymethane, CH3-O-CH3, both have the same molecular formula, C 2H6O.

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 dipoledipole attractions) in each will be much the same. However, ethanol has a hydrogen atom attached directly to an oxygen - and that oxygen still has exactly the same two lone pairs as in a water molecule. Hydrogen bonding can occur between ethanol molecules, although not as effectively as in water. The hydrogen bonding is limited by the fact that there is only one hydrogen in each ethanol molecule with sufficient + charge. In methoxymethane, the lone pairs on the oxygen are still there, but the hydrogens aren't sufficiently + for hydrogen bonds to form. Except in some rather unusual cases, the hydrogen atom has to be attached directly to the very electronegative element for hydrogen bonding to occur. The boiling points of ethanol and methoxymethane show the dramatic effect that the hydrogen bonding has on the stickiness of the ethanol molecules: ethanol (with hydrogen bonding) methoxymethane (without hydrogen bonding) The hydrogen bonding in the ethanol has lifted its boiling point about 100C. 78.5C -24.8C

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 CH 3NH2 (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.

STRUCTURES MENU

Ionic structures . . . Takes NaCl as a typical example to show the relationship between the physical properties of an ionic compound and its structu re. Also compares the structures of NaCl and CsCl. Giant covalent structures . . . Looks at diamond, graphite and silicon dioxide and the way their physical properties are affected by their structures. Metallic structures . . . Looks at the relationship between the physical properties of metals and their structures. Molecular structures . . . Looks in detail at the structures of ice, iodine and poly(ethene) as well as dealing in general with the sort of physical pro perties you might expect of molecular substances. How to decide . . . Explains how you can decide which sort of structure a substance has by looking at its physical properties. Physical properties of Period 3 elements . . . Explains how the structures of the elements from Na to Ar in the Periodic Table affect their simple physical properties.

IONIC STRUCTURES

This page explains the relationship between the arrangement of the ions in a typical ionic solid like sodium chloride and its physical properties melting point, boiling point, brittleness, solubility and electrical behaviour. It also explains w hy caesium chloride has a different structure from sodium chloride even though sodium and caesium are both in Group 1 of the Periodic Table.

Note: If you need to revise how ionic bonding arises, then you might like to follow this link. It isn't important for understanding this page, however.

The structure of a typical ionic solid - sodium chloride


How the ions are arranged in sodium chloride

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

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

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

How to draw this structure Draw a perfect square:

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

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

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

Now all you have to do is put the ions in. Use different colours or different sizes for the two different ions, and don' t forget a key. It doesn't matter whether you end up with a sodium ion or a chloride ion in the centre of the cube - all that matters is that they alternate in all three dimensions.

You should be able to draw a perfectly adequate free-hand sketch of this in under two minutes - less than one minute if you're not too fussy!

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

The different structure of caesium chloride


We'll look first at the arrangement of the ions and then talk about why the structures of sodium chloride and caesium chlorid e are different afterwards.

Warning: Before you go on with this section, make sure that you actually need it for your syllabus. If you don't, jump down the page to "The physical properties of sodium chloride"

How the ions are arranged in caesium chloride Imagine a layer of chloride ions as shown below. The individual chloride ions aren't touching each other. That's really important - if they were touching, there would be repulsion.

Now let's place a similarly arranged layer of caesium ions on top of these.

Notice that the caesium ions aren't touching each other either, but that each caesium ion is resting on four chlori de ions from the layer below. Now let's put another layer of chloride ions on, exactly the same as the first layer. Again, the chloride ions in this layer are NOT touching those in the bottom layer - otherwise you are introducing repulsion. Since we are looking directly down on the structure, you can't see the bottom layer of chloride ions any more, of course.

If you now think about a caesium ion sandwiched between the two layers of chloride ions, it is touching four chloride ions in the bottom layer, and another four in the top one. Each caesium ion is touched by eight chloride ions. We say that it is 8 -co-ordinated. If we added another layer of caesium ions, you could similarly work out that each chloride ion was touching eight caesium ion s. The chloride ions are also 8-co-ordinated. Overall, then, caesium chloride is 8:8-co-ordinated.

The final diagram in this sequence takes a slightly tilted view of the structure so that you can see how the layers build up. These diagrams are quite difficult to draw without it looking as if ions of the same charge are touching each other. They aren't!

Diagrams of ionic crystals are usually simplified to show the most basic unit of the repeating pattern. For caesium chloride, you could, for example, draw a simple diagram showing the arrangement of the chloride ions around each caesium ion:

By reversing the colours (green chloride ion in the centre, and orange caesium ions surrounding it), you would have an exactl y equivalent diagram for the arrangement of caesium io ns around each chloride ion.

Note: These diagrams are difficult enough to draw convincingly on a computer. Trying to draw them freehand in an exam is seriously difficult. If you are doing a syllabus which wants you to know about the structure of caesium chloride, take a careful look at past exam papers and mark schemes to see exactly what sort of diagrams (if any) you need to use, and then practise them so that you can draw them quickly and well. If you haven't got any past papers and mark schemes, follow this link to the syllabuses page to find out how to get them if you are doing a UKbased exam.

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

Now imagine what would happen if you replaced the caesium ion with the smaller sodium ion. Sodium ions are, of course, smalle r than caesium ions because they have fewer layers of electrons around them. You still have to keep the chloride ions in contact with the sodium. The effect of this would be that the whole arrangement would shrink, bringing the chloride ions into contact with each other - and that introduces repulsion. Any gain in attractions because you have eight chlorides around the sodiu m rather than six is more than countered by the new repulsions between the chloride ions themselves. When sodium chloride is 6:6-co-ordinated, there are no such repulsions - and so that is the best way for it to organise itself. Which structure a simple 1:1 compound like NaCl or CsCl crystallises in depends on the radius ratio of the positive and the negative ions. If the radius of the positive ion is bigger than 73% of that of the negative ion, then 8:8 -co-ordination is possible. Less than that (down to 41%) then you get 6:6-co-ordination. In CsCl, the caesium ion is about 93% of the size of the chloride ion - so is easily within the range where 8:8-co-ordination is possible. But with NaCl, the sodium ion is only about 52% of the size of the chloride ion. That puts it in the range where you get 6:6 -co-ordination.

Note: What happens below 41%? At this point the negative ions will touch each other again even with 6:6 -co-ordination. A new arrangement (known as 4:4-co-ordination) then becomes necessary. This is beyond any syllabus that I am currently tracking.

The physical properties of sodium chloride


Sodium chloride is taken as typical of ionic compounds, and is chosen rather than, say, caesium chloride, because it is found on every syllabus at this level. Sodium chloride has a high melting and boiling point There are strong electrostatic attractions between the positive and negative ions, and it takes a lot of heat energy to overc ome them. Ionic substances all have high melting and boiling points. Differences between ionic substances will depend on things like:

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

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

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

Ions of the same charge are brought side-by-side and so the crystal repels itself to pieces!

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

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

Sodium chloride is insoluble in organic solvents This is also typical of ionic solids. The attractions between the solvent molecules and the ions aren't big enough to overcom e the attractions holding the crystal together.

The electrical behaviour of sodium chloride Solid sodium chloride doesn't conduct electricity, because there are no electrons which are free to move. When it melts, sodi um chloride undergoes electrolysis, which involves conduction of electricity because of the movement and discharge of the ions. In the process, sodium and chlori ne are produced. This is a chemical change rather than a physical process. The positive sodium ions move towards the negatively ch arged electrode (the cathode). When they get there, each sodium ion picks up an electron from the electrode to form a sodium atom.

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

The new electrons deposited on the anode are pumped off around the external circuit by the power source, eventually ending up on the cathode where they will be transferred to sodium ions. Molten sodium chloride conducts electricity because of the movement of the ions in the melt, and the discharge of the ions at the electrodes. Both of these have to happen if you are to get electrons flowing in the external circuit. In solid sodium chloride, of course, that ion movement can't happen and that stops any possibility of any curr ent flow in the circuit.

GIANT COVALENT STRUCTURES

This page decribes the structures of giant covalent substances like diamond, graphite and silicon dioxide (silicon(IV) oxide) , and relates those structures to the physical properties of the substances.

The structure of diamond


The giant covalent structure of diamond Carbon has an electronic arrangement of 2,4. In diamond, each carbon shares electrons with four other carbon atoms - forming four single bonds.

In the diagram some carbon atoms only seem to be forming two bonds (or even one bond), but that's not really the case. We are only showing a small bit of the whole structure. This is a giant covalent structure - it continues on and on in three dimensions. It is not a molecule, because the number of atoms joined up in a real diamond is completely variable - depending on the size of the crystal.

Note: We quoted the electronic structure of carbon as 2,4. That simple view is perfectly adequate to explain the bonding in diamond If you are interested in a more modern view, you could . read the page on bonding in methane and ethane in the organic section of this site. In the case of diamond, each carbon is bonded to 4 other carbons rather than hydrogens, but that makes no essential difference.

How to draw the structure of diamond Don't try to be too clever by trying to draw too much of the structure! Learn to draw the diagram given above. Do it in the following stages:

Practise until you can do a reasonable free-hand sketch in about 30 seconds.

The physical properties of diamond Diamond

y y y y

has a very high melting point (almost 4000C). Very strong carbon -carbon covalent bonds have to be broken throughout the structure before melting occurs. is very hard. This is again due to the need to break very strong covalent bonds operating in 3 -dimensions. doesn't conduct electricity. All the electrons are held tightly between the atoms, and aren't free to move. is insoluble in water and organic solvents. There are no possible attractions which could occur between solvent molecules and carbon atoms which could outweigh the attractions between the covalently bound carbon atoms.

The structure of graphite


The giant covalent structure of graphite Graphite has a layer structure which is quite difficult to draw convincingly in three dimensions. The diagram below shows the arrangement of the atoms in each layer, and the way the layers are spaced.

Notice that you can't really draw the side view of the layers to the same scale as the atoms in the layer without one or other part of the diagram being either very spread out or very squashed. In that case, it is important to give some idea of the distances involved. The distance between the layers is about 2.5 times the distance between the atoms within each layer. The layers, of course, extend over huge numbers of atoms - not just the few shown above. You might argue that carbon has to form 4 bonds because of its 4 unpaired electrons, whereas in this diagram it only seems to be forming 3 bonds to the neighbouring carbons. This diagram is something of a simplification, and shows the arrangement of atoms rather than th e bonding.

The bonding in graphite Each carbon atom uses three of its electrons to form simple bonds to its three close neighbours. That leaves a fourth electron in the bonding level. These "spare" electrons in each carbon atom become delocalised over the whole of the sheet of atoms in one layer. They are no longer associated directly with any particular atom or pair of ato ms, but are free to wander throughout the whole sheet.

If you are interested (beyond A'level): The bonding in graphite is like a vastly extended version of the bonding in benzene. Each carbon atom undergoes sp 2 hybridisation, and then the unhybridised p orbitals on each carbon atom overlap sideways to give a massive pi system above and below the plane of the sheet of atoms.

The important thing is that the delocalised ele ctrons are free to move anywhere within the sheet - each electron is no longer fixed to a particular carbon atom. There is, however, no direct contact between the delocalised electrons in one sheet and those in the neighbourin g sheets. The atoms within a sheet are held together by strong covalent bonds - stronger, in fact, than in diamond because of the additional bonding caused by the delocalised electrons. So what holds the sheets together? In graphite you have the ultimate example of van der Waals disper sion forces. As the delocalised electrons move around in the sheet, very large temporary dipoles can be set up which will induce opposite dipoles in the sheets above and below - and so on throughout the whole graphite crystal.

Note: If you aren't sure about van der Waals forces follow this link before you go on. Use the BACK button on your browser to return to this page.

The physical properties of graphite

Graphite

y y y y y

has a high melting point, similar to that of diamond. In order to melt graphite, it isn't enough to loosen one sheet from ano ther. You have to break the covalent bonding throughout the whole structure. has a soft, slippery feel, and is used in pencil s and as a dry lubricant for things like locks. You can think of graphite rather like a pack of cards - each card is strong, but the cards will slide over each other, or even fall off the pack altogether. When you use a pencil, s heets are rubbed off and stick to the paper. has a lower density than diamond. This is because of the relatively large amount of space that is "wasted" between the sheets . is insoluble in water and organic solvents - for the same reason that diamond is insoluble. Attractions between solvent molecules and carbon atoms will never be strong enough to overcome the strong covalent bonds in graphite. conducts electricity. The delocalised electrons are free to move throughout the sheets. If a piece of graphite is connected i nto a circuit, electrons can fall off one end of the sheet and be replaced with new ones at the other end.

Note: The logic of this is that a piece of graphite ought only to conduct electricity in 2 -dimensions because electrons can only move around in the sheets - and not from one sheet to its neighbours. In practice, a real piece of graphite isn't a perfect crystal, but a host of small crystals stuck together at all sorts of angles. Electrons will be able to find a route through the largepiece of graphite in all directions by moving from one small crystal to the next.

The structure of silicon dioxide, SiO 2


Silicon dioxide is also known as silicon(IV) oxide. The giant covalent structure of silicon dioxide There are three different crystal forms of silicon dioxide. The easiest one to remember and draw is based on the diamond stru cture. Crystalline silicon has the same structure as diamond. To turn it into silicon dioxide, all you need to do is to modify the silicon structure by incl uding some oxygen atoms.

Notice that each silicon atom is bridged to its neighbours by an oxygen atom. Don't forget that this is just a tiny part of a giant structure extending on all 3 dimensions.

Note: If you want to be fussy, the Si-O-Si bond angles are wrong in this diagram. In reality the "bridge" from one silicon atom to its neighbour isn't in a straightline, but via a "V" shape (similar to the shape around the oxygen atom in a water molecule). It's extremely difficult to draw that convincingly and tidily in adiagram involving this number of atoms. The simplification is perfectly acceptable.

The physical properties of silicon dioxide Silicon dioxide

y y y y

has a high melting point - varying depending on what the particular structure is (remember that the structure given is only one of three possible structures), but around 1700C. Very strong silicon -oxygen covalent bonds have to be broken throughout the structure before melting occurs. is hard. This is due to the need to break the very strong covalent bonds. doesn't conduct electricity. There aren't any delocalised electrons. All the electrons are held tightly between the atoms , and aren't free to move. is insoluble in water and organic solvents. There are no possible attractions which could occur between solvent molecules and the silicon or oxygen atoms which could overcome the covalent bonds in the giant structure.

METALLIC STRUCTURES

This page decribes the structure of metals, and relates that structure to the physical properties of the metal.

The structure of metals


The arrangement of the atoms Metals are giant structures of atoms held together by metallic bonds. "Giant" implies that large but variable numbers of atoms are involved depending on the size of the bit of metal.

Note: Before you go on, it might be a good idea to read the page on bonding in metals unless you are reasonably happy about the idea of the delocalised electrons ("sea of electrons") in metals.

12-co-ordination Most metals are close packed - that is, they fit as many atoms as possible into the available volume. Each ato m in the structure has 12 touching neighbours. Such a metal is described as 12-co-ordinated. Each atom has 6 other atoms touching it in each layer.

There are also 3 atoms touching any particular atom in the layer above and another 3 in the layer underneath.

This second diagram shows the layer immediately above the first layer. There will be a corresponding layer underneath. (There are actually two different ways of placing the third layer in a close packed structure, but that goes beyond the req uirements of current A'level syllabuses.) 8-co-ordination Some metals (notably those in Group 1 of the Periodic Table) are packed less efficiently, having only 8 touching neighbours. These are 8-coordinated.

The left hand diagram shows that no atoms are touching each other within a particular layer . They are only touched by the at oms in the layers above and below. The right hand diagram shows the 8 atoms (4 above and 4 below) touching the darker coloured one. Crystal grains It would be misleading to suppose that all the atoms in a piece of metal are arranged in a regular way. Any piece of metal is made up of a large number of "crystal grains", which are regions of regularity. At the grain boundaries atoms have become misaligned.

Note: Within a crystal grain you get rather subtle irregularities known as dislocations. It isn't important to know about these forUK A level Chemistry (or equivalent) purposes, although they turn out to be essential in discussing the workability of metals at a higher level. I haven't included a description of them here because it is quite difficult to iv sualise how they work, and I don't want to add unnecessary complications.

The physical properties of metals


Melting points and boiling points Metals tend to have high melting and boiling points because of the strength of the metallic bond. The strength of the bond va ries from metal to metal and depends on the number of electrons which each atom delocalises into the sea of electro ns, and on the packing. Group 1 metals like sodium and potassium have relatively low melting and boiling points mainly because each atom only has one electron to contribute to the bond - but there are other problems as well:

y y

Group 1 elements are also ineff iciently packed (8-co-ordinated), so that they aren't forming as many bonds as most metals. They have relatively large atoms (meaning that the nuclei are some distance from the delocalised electrons) which also weaken s the bond.

Electrical conductivity Metals conduct electricity. The delocalised electrons are free to move throughout the structure in 3 -dimensions. They can cross grain boundaries. Even though the pattern may be disrupted at the boundary, as long as atoms are touching each other, the metall ic bond is still present. Liquid metals also conduct electricity, showing that although the metal atoms may be free to move, the delocalisation remains in force until the metal boils.

Thermal conductivity Metals are good conductors of heat. Heat energy is picked up by the electrons as additional kinetic energy (it makes them move faster). The energy is transferred throughout the rest of the metal by the moving electrons.

Strength and workability Malleability and ductility Metals are described as malleable (can be beaten into sheets) and ductile (can be pulled out into wires). This is because of the ability of the atoms to roll over each other into new positions without breaking the metallic bond. If a small stress is put onto the metal, the layers of atoms will start to roll over each other. If the stress is released ag ain, they will fall back to their original positions. Under these circumstances, the metal is said to be elastic.

If a larger stress is put on, the atoms roll over each other into a new position, and the metal is permanently changed.

The hardness of metals This rolling of layers of atoms over each other is hindered by grain boundaries because the rows of atoms don't line up prope rly. It follows that the more grain boundaries there are (the smaller the individual crystal grains), the harder the metal becomes. Offsetting this, because the grain boundaries are areas where the atoms aren't in such good contact with each other, metals t end to fracture at grain boundaries. Increasing the number of grain boundaries not only makes the metal harder, but also makes it more brittle. Controlling the size of the crystal grains If you have a pure piece of metal, you can control the size of the gr ains by heat treatment or by working the metal. Heating a metal tends to shake the atoms into a more regular arrangement - decreasing the number of grain boundaries, and so making the metal softer. Banging the metal around when it is cold tends to produce lots of small grains. Cold working therefore makes a metal harder. To restore its workability, you would need to reheat it. You can also break up the regular arrangement of the atoms by inserting atoms of a slightly different size into the structure . Alloys such as brass (a mixture of copper and zinc) are harder than the original metals because the irregularity in the structure helps to stop rows of atoms from slipping over each other.

MOLECULAR STRUCTURES

This page describes how the physical properties of substances having molecular structures varies with the type of intermolecular attractions hydrogen bonding or van der Waals forces.
Important! There's not much point in reading this page unless you are reasonably happy about the origin of hydrogen bonding and van der Waals forces. Follow these links first if you aren't sure about these.

The physical properties of molecular substances


Molecules are made of fixed numbers of atoms joined together by covalent bonds, and can range from the very small (even down to single atoms, as in the noble gases) to the very large (as in polymers, proteins or even DNA). The covalent bonds holding the molecules together are very strong, but these are largely irrelevant to the physical propertie s of the substance. Physical properties are governed by the intermolecular forces - forces attracting one molecule to it s neighbours - van der Waals attractions or hydrogen bonds. Melting and boiling points Molecular substances tend to be gases, liquids or low melting point solids, because the intermolecular forces of attraction a re comparatively weak. You don't have to break any covalent bonds in order to melt or boil a molecular substance.

Note: This is really important! You can make yourself look extremely stupid if you imply in an exam that boiling water, for example splits it into hydrogen and oxygen by breaking covalent , bonds. Exactly the same water molecules are present in ice, water and steam.

The size of the melting or boiling point will depend on the strength of the intermolecular forces. The presence of hydrogen b onding will lift the melting and boiling points. The larger the molecule the more van der Waals attractions are possible - and those will also need more energy to break.

Solubility in water

Most molecular substances are insoluble (or only very sparingly soluble) in water. Those which do dissolve often react with the water, or else are capable of forming hydrogen bonds with the water. Why doesn't methane, CH4, dissolve in water? The methane itself isn't the problem. Methane is a gas, and so its molecules are already separate - the water doesn't need to pull them apart from one another. The problem is the hydrogen bonds between the water molecules. If methane were to dissolve, it would have to force its way be tween water molecules and so break hydrogen bonds. That costs a reasonable amount of ener gy. The only attractions possible between methane and water molecules are the much weaker van der Waals forces - and not much energy is released when these are set up. It simply isn't energetically profitable for the methane and water to mix. Why does ammonia, NH3, dissolve in water? Ammonia has the ability to form hydrogen bonds. When the hydrogen bonds between water molecules are broken, they can be repla ced by equivalent bonds between water and ammonia molecules. Some of the ammonia also reacts with the water to produce ammonium ions and hydroxide ions.

The reversible arrows show that the reaction doesn't go to completion. At any one time only about 1% of the ammonia has actua lly reacted to form ammonium ions. The solubility of ammonia is mainly due to the hydrogen bonding and not the reaction. Other common substances which are freely soluble in water because they can hydrogen bond with water molecules include ethanol (alcohol) and sucrose (sugar).

Solubility in organic solvents Molecular substances are often soluble in organic solvents - which are themselves molecular. Both the solute (the substance which is dissolving) and the solvent are likely to have molecules attracted to each other by van der Waals forces. Although these attractions will be disrupted when they mix, they are replaced by similar ones between the two different sorts of molecules.

Electrical conductivity Molecular substances won't conduct electricity. Even in cases where electrons may be delocalised within a particular molecule , there isn't sufficient contact between the molecules to allow the electrons to move through the whole solid or liquid.

Some individual examples


Iodine, I2 Iodine is a dark grey crystalline solid with a purple vapour. M.Pt: 114C. B.Pt: 184C. It is very, very slightly soluble in water, but dissolves freely in organic solvents. Iodine is therefore a low melting point solid. The crystallinity suggests a regular packing of the molecules.

The structure is described as face centred cubic - it is a cube of iodine molecules with another molecule at the centre of each face.

The orientation of the iodine molecules within this structure is quite difficult to draw (let alone remember!). If your sylla bus and past exam papers suggests that you need to remember it, look carefully at the next sequence of diagrams showing the layers.

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

Notice that as you look down on the cube, all the molecules on the left and right hand sides are aligned the same way. The on es in the middle are aligned in the opposite way. All these diagrams show an "exploded" view of the crystal. The iodine molecules are, of course, touching each other. Measurem ents of the distances between the centres of the atoms in the crystal show two different values:

The iodine atoms within each molecule are pulled closely together by the covalent bond. The van der Waals attraction between the molecules is much weaker, and you can think of the atoms in two separate molecules as just loosely touching each other.

Ice Ice is a good example of a hydrogen bonded solid. There are lots of different ways that the water molecules can be arranged in ice. This is one of them, but NOT the common one - I can't draw that in any way that makes sense! The one below is known as "cubic ice", or "ice Ic". It is based on the water molecules arranged in a diamond structure.

This is just a small part of a structure which extends over huge numbers of molecules in three dimensions. In the diagram, th e lines represent hydrogen bonds. The lone pairs that the hydrogen atoms are attracted to are left out for clarity. Cubic ice is only stable at temperatures below -80C. The ice you are familiar with has a different, hexagonal structure. It is called "ice Ih".

Note: Don't worry about this problem. If asked to draw ice in an exam at this level (16- 18 year olds), don't try to be too clever. It is probably best not to go beyond the top five molecules in the above diagram. This will show the essential features of the bonding in the structure witho getting bogged down in stuff which is far beyond this level. ut If you are interested in following this up, try a Google search (at the bottom of the Main Menu- link below) using the search term ice structure hexagonal cubic (or something similar). This will throw up lots of information together with an assortment of fairly dreadful diagrams which I for one don't have the visual im agination to unscramble!

The unusual density behaviour of water The hydrogen bonding forces a rather open structure on the ice - if you made a model of it, you would find a significant amount of wasted space. When ice melts, the structure breaks down and the molecules tend to fill up this wasted space.

This means that the water formed takes up less space than the original ice. Ice is a very unusual solid in this respect - most solids show an increase in volume on melting. When water freezes, the opposite happens - there is an expansion as the hydrogen bonded structure establishes. Most liquids contract on freezing. Remnants of the rigid hydrogen bonded structure are still present in very cold liquid water, and don't finally disappear unti l 4C. From 0C to 4C, the density of water increases as the molecules free themselves from the open structure a nd take up less space. After 4C, the thermal motion of the molecules causes them to move apart and the density falls. That's the normal behaviour with liquids on heating.

Note: You can find more about water (particularly its abnormally high boiling point) in the page on hydrogen bonding.

Polymers Bonding in polymers Polymers like poly(ethene) - commonly called polythene - consist of very long molecules. Poly(ethene) molecules are made by joining up lots of ethene molecules into chains of covalently bound carbon atoms with hydrogens attached. There may be short branches along the main chain, also consisting of carbon chains with attached hydrogens. The molecules are attracted to each other in the solid by van der Waals dispersion forces. By controlling the conditions under which ethene is polymerised, it is possible to control the amount of branching to give tw o distinct types of polythene. High density polythene High density polythene has virtually unbranched chains. The lack of branching allows molecules to lie close together in a regular way which is almost crystalline. Because the molecules lie close together, dispersion forces are more effective, and so the plasti c is relatively strong and has a somewhat higher melting point than low density polythene. High density polythene is used for containers for household chemicals like washing -up liquid, for example, or for bowls or buckets. Low density polythene Low density polythene has lots of short branches along the chain. These branches prevent the chains from lying close together in a tidy arrangement. As a result dispersion forces are less and the plastic is weaker and has a lower melting point. Its density is l ower, of course, because of the wasted space within the unevenly packed structure. Low density polythene is used for things like plastic bags.

DECIDING WHAT TYPE OF STRUCTURE A SUBSTANCE HAS

This page explains how you can decide what sort of structure a substance has by looking at its physical properties. The page originally had a brief kinetic theory description of solids, liquids and gases. That has now been transferred to a s eparate introduction to kinetic theory page in the physical chemistry section of Chemguide.

Deducing the type of bonding from physical properties


Melting and boiling points The best place to start is usually the physical state. Melting point isn't always a good guide to the size of the attractions between particles, because the attractive forces have only been loosened on melting - not broken entirely. Boiling point is a much better guide, because enough heat has now been supplied to break the attractive forces completely. The stronger the attractions, the higher the boiling point.

That being said, melting points are often used to judge the size of attractive forces between particles in solids, but you wi ll find the occasional oddity. Those oddities usually disappear if you consider boiling points instead.

As an example: You would expect stronger metallic bonding in aluminium than in magnesium, because aluminium has 3 electrons to delocaliseinto the "sea of electrons" rather than magnesium's 2. The boiling points reflect this: Al 2470C, Mg 1110C. However, aluminium's melting point is only 10C higherthan magnesium's: Al 660C, Mg 650C. (I've never found a good explanation for this!) If you need some more background on metallic bonding, you could follow this link.

So . . . If it is a gas, liquid or low melting point solid, it will consist of covalently bound molecules (except the noble gases which have molecules consisting of single atoms). The size of the melting point or boiling point gives a guide to the strength of the intermolecular forces. That is then the end of the problem. If it is a gas, liquid or low melting point solid then you are talking about a simple molecular substance. Full stop ! If it is a high melting point solid, it will be a giant structure - either ionic, metallic or giant covalent. You now have to sort out which of these it is.

Effect of water Solubility of a solid in water (without reaction) suggests it is ionic. There are exceptions to this, of course. Sugar (sucrose) is soluble in water despite being a covalent molecule. It is capable of extensive hydrogen bonding with water molecules. And there are a lot of ionic compounds which are insoluble in water, of course. Solubility of a low melting point solid or a liquid in water (without it reacting) suggests a small molecule capable of hydro gen bonding - or, at least, a small very polar molecule.

Conduction of electricity Conduction of electricity in the solid state suggests delocalised electrons, and therefore either a metal or graphite. The cl ue as to which you had would usually come from other data - appearance, malleability, etc (see below).

Note: Semi-conductors like silicon - a giant covalent structure with the same arrangement of atoms as diamond - also conduct electricity. The theory of semi-conductors is beyond A level (or equivalent) chemistry syllabuses.

If a substance doesn't conduct electricity as a solid, but undergoes electrolysis when it is molten, that would confirm that it was ionic.

Note: Electrolysis is the splitting up of a compound using electricity. For example, moltensodium chloride conducts electricity and is split into sodium and chlorine in the process.

Appearance etc Don't forget about obvious things like the shiny appearance of most metals, and their ease of working. Metals are malleable ( can be bent or beaten into shape easily) and ductile (can be pulled out into wires). By contrast, giant ionic or giant covalent structures tend to be brittle - shattering rather than bending.

THE STRUCTURES OF THE PERIOD 3 ELEMENTS

This page describes the structures of the Period 3 elements from sodium to argon, and shows how these structures can be used to explain the physical properties of the elements.

Variation in physical properties in period 3


Melting and boiling points

In a moment we shall explain all the ups and d owns in this graph.

Electrical conductivity Sodium, magnesium and aluminium are all good conductors of electricity. Silicon is a semiconductor. None of the rest conduct electricity.

Explaining the trends


Warning! To understand this section you must be familiar with metallic bonding and the structure of metals, giant covalent structures, simple molecular structures and van der Waals forces. If you are uncertain of any of this, now is the time to revise it.

Three metallic structures Sodium, magnesium and aluminium all have metallic structures, which accounts for their electrical conductivity and relatively high melting and boiling points. Melting and boiling points rise across the three metals because of the increasing number of electrons which each atom can con tribute to the delocalised "sea of electrons". The atoms also get smaller and have more protons as you go from sodium to magnesium to alumin ium. The attractions and therefore the melting and boiling points increase because:

y y y

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

Silicon - a giant covalent structure

Silicon is a non-metal, and has a giant covalent structure ex actly the same as carbon in diamond - hence the high melting point. You have to break strong covalent bonds in order to melt it. There are no obviously free electrons in the structure, and although it conducts electricity, it doesn't do so in the same wa y as metals. Silicon is a semiconductor.

Note: Explaining how semiconductors conduct electricity is beyond the scope of A'level chemistry syllabuses.

Four molecular elements Phosphorus, sulphur, chlorine and argon are simple molecular substances with o nly van der Waals attractions between the molecules. Their melting or boiling points will be lower than those of the first four members of the period which have giant structures. The presence of individual molecules prevents any possibility of electrons flowing, and so none of them conduct electricity. The sizes of the melting and boiling points are governed entirely by the sizes of the molecules:

Argon molecules consist of single argon atoms. Phosphorus There are several forms of phosphorus. The data in the graph at the top of the page applies to white phosphorus which contains P 4 molecules. To melt phosphorus you don't have to break any covalent bonds - just the much weaker van der Waals forces between the molecules. Sulphur Sulphur consists of S 8 rings of atoms. The molecules are bigger than phosphorus molecules, and so the van der Waals attractions will be stronger, leading to a higher melting and boiling point. Chlorine Chlorine, Cl 2, is a much smaller molecule with comparatively weak van der Waals att ractions, and so chlorine will have a lower melting and boiling point than sulphur or phosphorus. Argon Argon molecules are just single argon atoms, Ar. The scope for van der Waals attractions between these is very limited and so the melting and boiling points of argon are lower again.

Note: You might also be interested in the trends in ionisation energy, atomic radius and electronegativity in this period. You will find relevant descriptions and explanations if you follow these links - or they are available via the menus below. If you are exploring Period 3 in detail, this link will take you to a major section covering all the aspects of Period 3 chemistry needed for the UK A level syllab uses - including reactions of the elements, their oxides, chlorides and hydroxides.

Você também pode gostar