Anda di halaman 1dari 35

IONIC (ELECTROVALENT) BONDING

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

A simple view of ionic bonding

The importance of noble gas structures

At a simple level (like GCSE) a lot of importance is attached to the electronic


structures of noble gases like neon or argon which have eight electrons in their
outer energy levels (or two in the case of helium). These noble gas structures are
thought of as being in some way a "desirable" thing for an atom to have.

You may well have been left with the strong impression that when other atoms
react, they try to organise things such that their outer levels are either completely
full or completely empty.

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

Ionic bonding in sodium chloride

Sodium (2,8,1) has 1 electron more than a stable noble gas structure (2,8). If it
gave away that electron it would become more stable.

Chlorine (2,8,7) has 1 electron short of a stable noble gas structure (2,8,8). If it
could gain an electron from somewhere it too would become more stable.

The answer is obvious. If a sodium atom gives an electron to a chlorine atom, both
become more stable.
The sodium has lost an electron, so it no longer has equal numbers of electrons
and protons. Because it has one more proton than electron, it has a charge of 1+. If
electrons are lost from an atom, positive ions are formed.

Positive ions are sometimes called cations.

The chlorine has gained an electron, so it now has one more electron than proton.
It therefore has a charge of 1-. If electrons are gained by an atom, negative ions are
formed.

A negative ion is sometimes called an anion.

The nature of the bond

The sodium ions and chloride ions are held together by the strong electrostatic
attractions between the positive and negative charges.

The formula of sodium chloride

You need one sodium atom to provide the extra electron for one chlorine atom, so
they combine together 1:1. The formula is therefore NaCl.

Some other examples of ionic bonding

magnesium oxide

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

calcium chloride

This time you need two chlorines to use up the two outer electrons in the calcium.
The formula of calcium chloride is therefore CaCl 2.

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

Electrons are transferred from one atom to another resulting in the formation of
positive and negative ions.

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

So what's new? At heart - nothing. What needs modifying is the view that there is
something magic about noble gas structures. There are far more ions which don't
have noble gas structures than there are which do.

Some common ions which don't have noble gas structures

You may have come across some of the following ions in a basic course like GCSE.
They are all perfectly stable , but not one of them has a noble gas structure.

Fe3+ [Ar]3d5
Cu2+ [Ar]3d9
Zn2+ [Ar]3d10
Ag+ [Kr]4d10
Pb2+ [Xe]4f145d106s2

Noble gases (apart from helium) have an outer electronic structure ns 2np6.

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
Sc3+ with an argon structure, for example), all transition elements and any metals
following a transition series (like tin and lead in Group 4, for example) will have
structures like those above.

That means that the only elements to form positive ions with noble gas structures
(apart from odd ones like scandium) are those in groups 1 and 2 of the Periodic
Table and aluminium in group 3 (boron in group 3 doesn't form ions).

Negative ions are tidier! Those elements in Groups 5, 6 and 7 which form simple
negative ions all have noble gas structures.

If elements aren't aiming for noble gas structures when they form ions, what
decides how many electrons are transferred? The answer lies in the energetics of
the process by which the compound is made.

Warning! From here to the bottom of this page goes beyond anything you are
likely to need for A'level purposes. You should read it, but you almost certainly
won't be tested on it for the purposes of UK A level (or its equivalents). Check
your syllabus if you aren't sure.
Questions to test your understanding

If this is the first set of questions you have done, please read the introductory page before you
start. You will need to use the BACK BUTTON on your browser to come back here afterwards.

questions on ionic bonding

answers

There are no questions to test the rest of this page.

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 CaCl2 rather than CaCl or CaCl3?

If one mole of CaCl (containing Ca+ ions) is made from its elements, it is possible to
estimate that about 171 kJ of heat is evolved.

However, making CaCl2 (containing Ca2+ ions) releases more heat. You get 795 kJ.
That extra amount of heat evolved makes the compound more stable, which is why
you get CaCl2 rather than CaCl.

What about CaCl3 (containing Ca3+ ions)? To make one mole of this, you can
estimate that you would have to put in 1341 kJ. This makes this compound
completely non-viable. Why is so much heat needed to make CaCl 3? It is because
the third ionisation energy (the energy needed to remove the third electron) is
extremely high (4940 kJ mol-1) because the electron is being removed from the 3-
level rather than the 4-level. Because it is much closer to the nucleus than the first
two electrons removed, it is going to be held much more strongly.

Note: It would pay you to read about ionisation energies if you really want to
understand this.

You could also go to a standard text book and investigate Born-Haber Cycles.

A similar sort of argument applies to the negative ion. For example, oxygen forms
an O2- ion rather than an O- ion or an O3- ion, because compounds containing the
O2- ion turn out to be the most energetically stable.
COVALENT BONDING - SINGLE BONDS

This page explains what covalent bonding is. It starts with a simple picture of the
single covalent bond, and then modifies it slightly for A'level purposes.

It also goes on to a more sophisticated view involving hybridisation. This isn't


required by many UK-based syllabuses at this level. However, if you can follow it,
it will make the bonding in organic compounds easier to understand. I shall make
use of it throughout the rest of Chemguide.

You will find a link to a page on double covalent bonds at the bottom of the page.

A simple view of covalent bonding

The importance of noble gas structures

At a simple level (like GCSE) a lot of importance is attached to the electronic


structures of noble gases like neon or argon which have eight electrons in their
outer energy levels (or two in the case of helium). These noble gas structures are
thought of as being in some way a "desirable" thing for an atom to have.

You may well have been left with the strong impression that when other atoms
react, they try to achieve noble gas structures.

As well as achieving noble gas structures by transferring electrons from one atom
to another as in ionic bonding, it is also possible for atoms to reach these stable
structures by sharing electrons to give covalent bonds.

Some very simple covalent molecules

Chlorine

For example, two chlorine atoms could both achieve stable structures by sharing
their single unpaired electron as in the diagram.
The fact that one chlorine has been drawn with electrons marked as crosses and
the other as dots is simply to show where all the electrons come from. In reality
there is no difference between them.

The two chlorine atoms are said to be joined by a covalent bond. The reason that
the two chlorine atoms stick together is that the shared pair of electrons is
attracted to the nucleus of both chlorine atoms.

Hydrogen

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

Hydrogen chloride

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

Covalent bonding at A'level

Cases where there isn't any difference from the simple view

If you stick closely to modern A'level syllabuses, there is little need to move far
from the simple (GCSE) view. The only thing which must be changed is the over-
reliance on the concept of noble gas structures. Most of the simple molecules you
draw do in fact have all their atoms with noble gas structures.

For example:

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

Cases where the simple view throws up problems

Boron trifluoride, BF3

A boron atom only has 3 electrons in its outer level, and there is no possibility of it
reaching a noble gas structure by simple sharing of electrons. Is this a problem?
No. The boron has formed the maximum number of bonds that it can in the
circumstances, and this is a perfectly valid structure.

Energy is released whenever a covalent bond is formed. Because energy is being


lost from the system, it becomes more stable after every covalent bond is made. It
follows, therefore, that an atom will tend to make as many covalent bonds as
possible. In the case of boron in BF3, three bonds is the maximum possible
because boron only has 3 electrons to share.

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

Phosphorus(V) chloride, PCl5

In the case of phosphorus, 5 covalent bonds are possible - as in PCl 5.

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 PCl5 (like the previous diagram of PCl3) shows only the outer
electrons.

Notice that the phosphorus now has 5 pairs of electrons in the outer level -
certainly not a noble gas structure. You would have been content to draw PCl 3 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 most current A'level syllabuses. Don't be put off by this! It isn't
particularly difficult, and is extremely useful if you are going to understand the
bonding in some important organic compounds.
A more sophisticated view of covalent bonding

The bonding in methane, CH4

Warning! If you aren't happy with describing electron arrangements in s and p


notation, and with the shapes of s and p orbitals, you need to read about orbitals
before you go on.

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

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

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

There is a serious mis-match between this structure and the modern electronic
structure of carbon, 1s22s22px12py1. The modern structure shows that there are
only 2 unpaired electrons to share with hydrogens, 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 energy to promote an electron from the 2s to
the empty 2p to give 4 unpaired electrons. The extra
energy released when the bonds form more than
compensates for the initial input.

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

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

Now that we've got 4 unpaired electrons ready for bonding, another problem
arises. In methane all the carbon-hydrogen bonds are identical, but our electrons
are in two different kinds of orbitals. You aren't going to get four identical bonds
unless you start from four identical orbitals.

Hybridisation

The electrons rearrange themselves again in a process called hybridisation. This


reorganises the electrons into four identical hybrid orbitals
called sp3 hybrids (because they are made from one s
orbital and three p orbitals). You should read "sp 3" as "s p
three" - not as "s p cubed".

sp3 hybrid orbitals look a bit like half a p orbital, and they
arrange themselves in space so that they are as far apart as
possible. You can picture the nucleus as being at the centre of a tetrahedron (a
triangularly based pyramid) with the orbitals pointing to the corners. For clarity,
the nucleus is drawn far larger than it really is.

What happens when the bonds are formed?

Remember that hydrogen's electron is in a 1s orbital - a spherically symmetric


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

Four molecular orbitals are formed, looking rather like the original 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 dot and a cross.

The principles involved - promotion of electrons if necessary, then hybridisation,


followed by the formation of molecular orbitals - can be applied to any covalently-
bound molecule.

Note: You will find this bit on methane repeated in the organic section of this site.
That article on methane goes on to look at the formation of carbon-carbon single
bonds in ethane.
The bonding in the phosphorus chlorides, PCl3 and PCl5

What's wrong with the simple view of PCl3?

This diagram only shows the outer (bonding) electrons.

Nothing is wrong with this! (Although it doesn't account for the shape of the
molecule properly.) If you were going to take a more modern look at it, the
argument would go like this:

Phosphorus has the electronic structure 1s22s22p63s23px13py13pz1. If we look only


at the outer electrons as "electrons-in-boxes":

There are 3 unpaired electrons that can be used to form bonds with 3 chlorine
atoms. The four 3-level orbitals hybridise to produce 4 equivalent 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's unpaired electrons
to make 3 molecular orbitals.

You might wonder whether all this is worth the bother! Probably not! It is worth it
with PCl5, though.

What's wrong with the simple view of PCl5?

You will remember that the dots-and-crosses picture of PCl 5 looks awkward
because the phosphorus doesn't end up with a noble gas structure. This diagram
also shows only the outer electrons.

In this case, a more modern view makes things look better by abandoning any
pretence of worrying about noble gas structures.

If the phosphorus is going to form PCl5 it has first to generate 5 unpaired


electrons. It does this by promoting one of the electrons in the 3s orbital to the
next available higher energy orbital.

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

This leaves the phosphorus with this arrangement of its electrons:


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

The electrons in each of these orbitals would then share space with electrons
from five chlorines to make five new molecular orbitals - and hence five covalent
bonds.

Why does phosphorus form these extra two bonds? It puts in an amount of
energy to promote an electron, which is more than paid back when the new bonds
form. Put simply, it is energetically profitable for the phosphorus to form the extra
bonds.

The advantage of thinking of it in this way is that it completely ignores the


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

If you are a teacher or if you are likely to do chemistry at university: A paper


published in 2007 suggests that this explanation is seriously flawed. If you are
likely to do chemistry at university level, you will probably have to discard it later in
favour of a more accurate explanation. However, that explanation is way beyond
16 - 18 year old level.

If you get asked about this at the equivalent of UK A level, you will have to give the
explanation above - there is no alternative. Don't worry about it!

If you want a little bit more detail you will find it on this page.
A non-existent compound - NCl5

Nitrogen is in the same Group of the Periodic Table as phosphorus, and you
might expect it to form a similar range of compounds. In fact, it doesn't. For
example, the compound NCl3 exists, but there is no such thing as NCl5.

Nitrogen is 1s22s22px12py12pz1. The reason that NCl5 doesn't exist is that in order
to form five bonds, the nitrogen would have to promote one of its 2s electrons.
The problem is that there aren't any 2d orbitals to promote an electron into - and
the energy gap to the next level (the 3s) is far too great.

In this case, then, the energy released when the extra bonds are made isn't
enough to compensate for the energy needed to promote an electron - and so
that promotion doesn't happen.

Atoms will form as many bonds as possible provided it is energetically profitable.

CO-ORDINATE (DATIVE COVALENT) BONDING

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

Co-ordinate (dative covalent) bonding

A covalent bond is formed by two atoms sharing a pair of electrons. The atoms are
held together because the electron pair is attracted by both of the nuclei.

In the formation of a simple covalent bond, each atom supplies one electron to the
bond - but that doesn't have to be the case. A co-ordinate bond (also called a dative
covalent bond) is a covalent bond (a shared pair of electrons) in which both
electrons come from the same atom.

For the rest of this page, we shall use the term co-ordinate bond - but if you prefer to
call it a dative covalent bond, that's not a problem!

The reaction between ammonia and hydrogen chloride

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

Ammonium ions, NH4+, are formed by the transfer of a hydrogen ion from the
hydrogen chloride to the lone pair of electrons on the ammonia molecule.

When the ammonium ion, NH4+, is formed, the fourth hydrogen is attached by a
dative covalent bond, because only the hydrogen's nucleus is transferred from the
chlorine to the nitrogen. The hydrogen's electron is left behind on the chlorine to
form a negative chloride ion.

Once the ammonium ion has been formed it is impossible to tell any difference
between the dative covalent and the ordinary covalent bonds. Although the electrons
are shown differently in the diagram, there is no difference between them in reality.

Representing co-ordinate bonds

In simple diagrams, a co-ordinate bond is shown by an arrow. The arrow points from
the atom donating the lone pair to the atom accepting it.

Dissolving hydrogen chloride in water to make hydrochloric acid

Something similar happens. A hydrogen ion (H+) is transferred from the chlorine to
one of the lone pairs on the oxygen atom.

The H3O+ ion is variously called the hydroxonium ion, the hydronium ion or the
oxonium ion.

In an introductory chemistry course (such as GCSE), whenever you have talked


about hydrogen ions (for example in acids), you have actually been talking about the
hydroxonium ion. A raw hydrogen ion is simply a proton, and is far too reactive to
exist on its own in a test tube.
If you write the hydrogen ion as H+(aq), the "(aq)" represents the water molecule that
the hydrogen ion is attached to. When it reacts with something (an alkali, for
example), the hydrogen ion simply becomes detached from the water molecule
again.

Note that once the co-ordinate bond has been set up, all the hydrogens attached to
the oxygen are exactly equivalent. When a hydrogen ion breaks away again, it could
be any of the three.

The reaction between ammonia and boron trifluoride, BF3

If you have recently read the page on covalent bonding, you may remember boron
trifluoride as a compound which doesn't have a noble gas structure around the
boron atom. The boron only has 3 pairs of electrons in its bonding level, whereas
there would be room for 4 pairs. BF3 is described as being electron deficient.

The lone pair on the nitrogen of an ammonia molecule can be used to overcome
that deficiency, and a compound is formed involving a co-ordinate bond.

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

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

The structure of aluminium chloride

Aluminium chloride sublimes (turns straight from a solid to a gas) at about 180C. If
it simply contained ions it would have a very high melting and boiling point because
of the strong attractions between the positive and negative ions. The implication is
that it when it sublimes at this relatively low temperature, it must be covalent. The
dots-and-crosses diagram shows only the outer electrons.

AlCl3, like BF3, is electron deficient. There is likely to be a similarity,


because aluminium and boron are in the same group of the Periodic
Table, as are fluorine and chlorine.

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

Note: The uninteresting electrons on the chlorines have been


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

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

Note: Aluminium chloride is complicated because of the way it


keeps changing its bonding as the temperature increases. If you
are interested in exploring this in more detail, you could have a
look at the page about the Period 3 chlorides. It isn't particularly
relevant to the present page, though.

If you choose to follow this link, use the BACK button on your
browser to return quickly to this page later.

The bonding in hydrated metal ions

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

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


produced. Six water molecules bond to the aluminium to give an ion with the formula
Al(H2O)63+. It's called the hexaaquaaluminium ion - which translates as six ("hexa")
water molecules ("aqua") wrapped around an aluminium ion.

The bonding in this (and the similar ions formed by the great majority of other
metals) is co-ordinate (dative covalent) using lone pairs on the water
molecules.

Aluminium is 1s22s22p63s23px1. When it forms an Al3+ ion it loses the 3-level


electrons to leave 1s22s22p6.

That means that all the 3-level orbitals are now empty. The aluminium re-organises
(hybridises) six of these (the 3s, three 3p, and two 3d) to produce six new orbitals all
with the same energy. These six hybrid orbitals accept lone pairs from six water
molecules.

You might wonder why it chooses to use six orbitals rather than four or eight or
whatever. Six is the maximum number of water molecules it is possible to fit around
an aluminium ion (and most other metal ions). By making the maximum number of
bonds, it releases most energy and so becomes most energetically stable.

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

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

Note: Dotted arrows represent lone pairs coming from water


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

Two more molecules


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

Carbon monoxide, CO

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

Nitric acid, HNO3

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

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

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


in the ethanoate ion (although without the negative charge). You
will find thisdescribed on a page about the acidity of 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 chemistry context, you will find
a link at the bottom of this page.

What is electronegativity

Definition

Electronegativity is a measure of the tendency of an atom to attract a bonding pair


of electrons.

The Pauling scale is the most commonly used. Fluorine (the most electronegative
element) is assigned a value of 4.0, and values range down to caesium and
francium which are the least electronegative at 0.7.

What happens if two atoms of equal electronegativity bond together?

Consider a bond between two atoms, A and B. Each atom may be forming other
bonds as well as the one shown - but these are irrelevant to the argument.
If the atoms are equally electronegative, both have the same tendency to attract the
bonding pair of electrons, and so it will be found on average half way between the
two atoms. To get a bond like this, A and B would usually have to be the same atom.
You will find this sort of bond in, for example, H 2 or Cl2 molecules.

Note: It's important to realise that this is an average picture. The electrons are
actually in a molecular orbital, and are moving around all the time within that orbital.

This sort of bond could be thought of as being a "pure" covalent bond - where the
electrons are shared evenly between the two atoms.

What happens if B is slightly more electronegative than A?

B will attract the electron pair rather more than A does.

That means that the B end of the bond has more than its fair share of electron
density and so becomes slightly negative. At the same time, the A end (rather short
of electrons) becomes slightly positive. In the diagram, " " (read as "delta") means
"slightly" - so + means "slightly positive".

Defining polar bonds

This is described as a polar bond. A polar bond is a covalent bond in which there is
a separation of charge between one end and the other - in other words in which one
end is slightly positive and the other slightly negative. Examples include most
covalent bonds. The hydrogen-chlorine bond in HCl or the hydrogen-oxygen bonds
in water are typical.

What happens if B is a lot more electronegative than A?

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

A "spectrum" of bonds

The implication of all this is that there is no clear-cut division between covalent and
ionic bonds. In a pure covalent bond, the electrons are held on average exactly half
way between the atoms. In a polar bond, the electrons have been dragged slightly
towards one end.

How far does this dragging have to go before the bond counts as ionic? There is no
real answer to that. You normally think of sodium chloride as being a typically ionic
solid, but even here the sodium hasn't completely lost control of its electron.
Because of the properties of sodium chloride, however, we tend to count it as if it
were purely ionic.

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

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

Summary

No electronegativity difference between two atoms leads to a pure non-polar


covalent bond.

A small electronegativity difference leads to a polar covalent bond.


A large electronegativity difference leads to an ionic bond.

Polar bonds and polar molecules

In a simple molecule like HCl, if the bond is polar, so also is the whole molecule.
What about more complicated molecules?

In CCl4, each bond is polar.

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

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

By contrast, CHCl3 is polar.

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

Patterns of electronegativity in the Periodic Table

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

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

Trends in electronegativity across a period

As you go across a period the electronegativity increases. The chart shows


electronegativities from sodium to chlorine - you have to ignore argon. It doesn't
have an electronegativity, because it doesn't form bonds.
Trends in electronegativity down a group

As you go down a group, electronegativity decreases. (If it increases up to fluorine, it


must decrease as you go down.) The chart shows the patterns of electronegativity in
Groups 1 and 7.

Explaining the patterns in electronegativity

The attraction that a bonding pair of electrons feels for a particular nucleus depends
on:

the number of protons in the nucleus;

the distance from the nucleus;

the amount of screening by inner electrons.

Note: If you aren't happy about the concept of screening 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 charges on the


nucleus increases. That attracts the bonding pair of electrons more strongly.

Why does electronegativity fall as you go down a group?

Think of hydrogen fluoride and hydrogen chloride.

The bonding pair is shielded from the fluorine's nucleus only by the 1s 2 electrons. In
the chlorine case it is shielded by all the 1s22s22p6 electrons.

In each case there is a net pull from the centre of the fluorine or chlorine of +7. But
fluorine has the bonding pair in the 2-level rather than the 3-level as it is in chlorine.
If it is closer to the nucleus, the attraction is greater.

As you go down a group, electronegativity decreases because the bonding pair of


electrons is increasingly distant from the attraction of the nucleus.

Diagonal relationships in the Periodic Table

What is a diagonal relationship?

At the beginning of periods 2 and 3 of the Periodic Table, there are several cases
where an element at the top of one group has some similarities with an element in
the next group.

Three examples are shown in the diagram below. Notice that the similarities occur in
elements which are diagonal to each other - not side-by-side.

For example, boron is a non-metal with some properties rather like silicon. Unlike
the rest of Group 2, beryllium has some properties resembling aluminium. And
lithium has some properties which differ from the other elements in Group 1, and in
some ways resembles magnesium.

There is said to be a diagonal relationship between these elements.

There are several reasons for this, but each depends on the way atomic properties
like electronegativity vary around the Periodic Table.

So we will have a quick look at this with regard to electronegativity - which is


probably the simplest to explain.

Explaining the diagonal relationship with regard to electronegativity

Electronegativity increases across the Periodic Table. So, for example, the
electronegativities of beryllium and boron are:

Be 1.5
B 2.0

Electronegativity falls as you go down the Periodic Table. So, for example, the
electronegativities of boron and aluminium are:

B 2.0
Al 1.5

So, comparing Be and Al, you find the values are (by chance) exactly the same.
The increase from Group 2 to Group 3 is offset by the fall as you go down Group 3
from boron to aluminium.

Something similar happens from lithium (1.0) to magnesium (1.2), and from boron
(2.0) to silicon (1.8).

In these cases, the electronegativities aren't exactly the same, but are very close.

Similar electronegativities between the members of these diagonal pairs means that
they are likely to form similar types of bonds, and that will affect their chemistry. You
may well come across examples of this later on in your course.

Questions to test your understanding

If this is the first set of questions you have done, please read the introductory page before you start.
You will need to use the BACK BUTTON on your browser to come back here afterwards.

questions on electronegativity

answers

There are no questions on the rest of this page.

Warning! As far as I am aware, none of the UK-based A level (or equivalent)


syllabuses any longer want the next bit. It used to be on the AQA syllabus, but has
been removed from their new syllabus. At the time of writing, it does, however, still
appear on at least one overseas A level syllabus (Malta, but there may be others
that I'm not aware of). If in doubt, check your syllabus.

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

The polarising ability of positive ions

What do we mean by "polarising ability"?


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

Solid aluminium chloride is covalent. Imagine instead that it was ionic. It would
contain Al3+ and Cl- ions.

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

We say that the aluminium ions polarise the chloride ions.

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

Factors affecting polarising ability

Positive ions can have the effect of polarising (electrically distorting) nearby negative
ions. The polarising ability depends on the charge density in the positive ion.

Polarising ability increases as the positive ion gets smaller and the number of
charges gets larger.

As a negative ion gets bigger, it becomes easier to polarise. For example, in an


iodide ion, I-, the outer electrons are in the 5-level - relatively distant from the
nucleus.
A positive ion would be more effective in attracting a pair of electrons from an iodide
ion than the corresponding electrons in, say, a fluoride ion where they are much
closer to the nucleus.

Aluminium iodide is covalent because the electron pair is easily dragged away from
the iodide ion. On the other hand, aluminium fluoride is ionic because the aluminium
ion can't polarise the small fluoride ion sufficiently to form a covalent bond.

Anda mungkin juga menyukai