You are on page 1of 31

CHAPTER 01

PERIOD 3 ELEMENTS
1.1 Atomic And Physical Properties Of The Period 3 Elements
Atomic Properties
i. Electronic structures
In Period 3 of the Periodic Table, the 3s and 3p orbitals are filling with electrons. Just as a
reminder, the shortened versions of the electronic structures for the eight elements are:
In each case, [Ne] represents the complete electronic structure of a neon atom.
Chemical element
11 Na Sodium
12 Mg Magnesium
13 Al Aluminium
14 Si Silicon
15 P
Phosphorus
16 S
Sulfur
17 Cl Chlorine
18 Ar Argon

Chemical series
Alkali metal
Alkaline earth metal
Post-transition metal
Metalloid
Polyatomic nonmetal
Polyatomic nonmetal
Diatomic nonmetal
Noble gas

Electron configuration
[Ne] 3s1
[Ne] 3s2
[Ne] 3s2 3p1
[Ne] 3s2 3p2
[Ne] 3s2 3p3
[Ne] 3s2 3p4
[Ne] 3s2 3p5
[Ne] 3s2 3p6

ii. First ionisation energy


The first ionisation energy is the energy required to remove the most loosely held electron
from one mole of gaseous atoms to produce 1 mole of gaseous ions each with a charge of
1+.
It is the energy needed to carry out this change per mole of X.
The pattern of first ionisation energies across Period 3

Notice that the general trend is upwards, but this is broken by falls between magnesium
and aluminium, and between phosphorus and sulphur.
Explaining the pattern
First ionisation energy is governed by:
1

the charge on the nucleus;


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

The figures used to construct this diagram are based on:


metallic radii for Na, Mg and Al;
covalent radii for Si, P, S and Cl;
the van der Waals radius for Ar because it doesn't form any strong bonds.
It is fair to compare metallic and covalent radii because they are both being measured in
tightly bonded circumstances. It isn't fair to compare these with a van der Waals radius,
though.
The general trend towards smaller atoms across the period is NOT broken at argon. You
aren't comparing like with like. The only safe thing to do is to ignore argon in the
discussion which follows.
Explaining the trend
2

A metallic or covalent radius is going to be a measure of the distance from the nucleus to
the bonding pair of electrons. If you aren't sure about that, go back and follow the last link.
From sodium to chlorine, the bonding electrons are all in the 3-level, being screened by
the electrons in the first and second levels. The increasing number of protons in the
nucleus as you go across the period pulls the bonding electrons more tightly to it. The
amount of screening is constant for all of these elements.
Note: You might possibly wonder why you don't get extra screening from the 3s
electrons in the cases of the elements from aluminium to chlorine where the bonding
involves the p electrons. In each of these cases, before bonding happens, the existing s and
p orbitals are reorganised (hybridised) into new orbitals of equal energy. When these
atoms are bonded, there aren't any 3s electrons as such.
iv. Electronegativity
Electronegativity is a measure of the tendency of an atom to attract a bonding pair of
electrons.
The Pauling scale is the most commonly used. Fluorine (the most electronegative element)
is assigned a value of 4.0, and values range down to caesium and francium which are the
least electronegative at 0.7.
The trend
The trend across Period 3 looks like this:

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

Sodium, magnesium and aluminium all have metallic structures.


In sodium, only one electron per atom is involved in the metallic bond - the single 3s
electron. In magnesium, both of its outer electrons are involved, and in aluminium all
three.
The other difference you need to be aware of is the way the atoms are packed in the metal
crystal.
Sodium is 8-co-ordinated - each sodium atom is touched by only 8 other atoms.
Both magnesium and aluminium are 12-co-ordinated (although in slightly different ways).
This is a more efficient way to pack atoms, leading to less wasted space in the metal
structures and to stronger bonding in the metal.
A giant covalent structure
Silicon has a giant covalent structure just like diamond. A tiny part of the structure looks
like this:

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

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

ii.

Electrical conductivity
4

Sodium, magnesium and aluminium are all good conductors of electricity.


Conductivity increases as you go from sodium to magnesium to aluminium.
Silicon is a semiconductor.
None of the rest conduct electricity.
The three metals, of course, conduct electricity because the delocalised electrons (the "sea
of electrons") are free to move throughout the solid or the liquid metal.
In the silicon case, explaining how semiconductors conduct electricity is beyond the scope
at this level. With a diamond structure, you mightn't expect it to conduct electricity, but it
does!
The rest don't conduct electricity because they are simple molecular substances. There are
no electrons free to move around.
iii. Melting and boiling points
The chart shows how the melting and boiling points of the elements change as you go
across the period. The figures are plotted in kelvin rather than C to avoid having negative
values.

It is best to think of these changes in terms of the types of structure that we have talked
about further up.
The metallic structures
Melting and boiling points rise across the three metals because of the increasing strength
of the metallic bonds.
The number of electrons which each atom can contribute to the delocalised "sea of
electrons" increases. The atoms also get smaller and have more protons as you go from
sodium to magnesium to aluminium.
The attractions and therefore the melting and boiling points increase because:
The nuclei of the atoms are getting more positively charged.
The "sea" is getting more negatively charged.
The "sea" is getting progressively nearer to the nuclei and so more strongly
attracted.
Silicon
Silicon has high melting and boiling points because it is a giant covalent structure. You
have to break strong covalent bonds before it will melt or boil.
Because you are talking about a different type of bond, it isn't profitable to try to directly
compare silicon's melting and boiling points with aluminium's.
5

The four molecular elements


Phosphorus, sulphur, chlorine and argon are simple molecular substances with only van
der Waals attractions between the molecules. Their melting or boiling points will be lower
than those of the first four members of the period which have giant structures.
The sizes of the melting and boiling points are governed entirely by the sizes of the
molecules. Remember the structures of the molecules:

Phosphorus
Phosphorus contains P4 molecules. To melt phosphorus you don't have to break any
covalent bonds - just the much weaker van der Waals forces between the molecules.
Sulphur
Sulphur consists of S8 rings of atoms. The molecules are bigger than phosphorus
molecules, and so the van der Waals attractions will be stronger, leading to a higher
melting and boiling point.
Chlorine
Chlorine, Cl2, is a much smaller molecule with comparatively weak van der Waals
attractions, and so chlorine will have a lower melting and boiling point than sulphur or
phosphorus.
Argon
Argon molecules are just single argon atoms, Ar. The scope for van der Waals attractions
between these is very limited and so the melting and boiling points of argon are lower
again.
Note: Boiling point is a better guide to the strength of the metallic bonds than
melting point. Metallic bonds still exist in the liquid metals and aren't completely broken
until the metal boils.

1.2 Chemical Reactions Of The Period 3 Elements


Reactions with water
Sodium
Sodium has a very exothermic reaction with cold water producing hydrogen and a
colourless solution of sodium hydroxide.
Magnesium
Magnesium has a very slight reaction with cold water, but burns in steam.
A very clean coil of magnesium dropped into cold water eventually gets covered in small
bubbles of hydrogen which float it to the surface. Magnesium hydroxide is formed as a
very thin layer on the magnesium and this tends to stop the reaction.
6

Magnesium burns in steam with its typical white flame to produce white magnesium oxide
and hydrogen.
Note: If you are heating the magnesium in a glass tube, the magnesium also reacts
with the glass. That leaves dark grey products (including silicon and perhaps boron from
the glass) as well as the white magnesium oxide.Notice also that the oxide is produced on
heating in steam. Hydroxides are only ever produced using liquid water.
Aluminium
Aluminium powder heated in steam produces hydrogen and aluminium oxide. The
reaction is relatively slow because of the existing strong aluminium oxide layer on the
metal, and the build-up of even more oxide during the reaction.
Silicon
There is a fair amount of disagreement in the books and on the web about what silicon
does with water or steam. The truth seems to depend on the precise form of silicon you are
using. The common shiny grey lumps of silicon with a rather metal-like appearance are
fairly unreactive. Most sources suggest that this form of silicon will react with steam at
red heat to produce silicon dioxide and hydrogen.
But it is also possible to make much more reactive forms of silicon which will react with
cold water to give the same products.
Note: These more reactive forms are produced as powders due to a very high
surface area, or perhaps because the silicon exists in a graphite-like structure. When
silicon is cut into slices, the silicon dust formed reacts with water at room temperature producing hydrogen and getting very hot.
Phosphorus and sulphur
These have no reaction with water.
Chlorine
Chlorine dissolves in water to some extent to give a green solution. A reversible reaction
takes place to produce a mixture of hydrochloric acid and chloric(I) acid (hypochlorous
acid).
Note: You may also find the chloric(I) acid written as HClO. The form I have used
more accurately reflects the way the atoms are joined up. It doesn't matter which you use.
In the presence of sunlight, the chloric(I) acid slowly decomposes to produce more
hydrochloric acid, releasing oxygen gas, and you may come across an equation showing
the overall change:
Argon
There is no reaction between argon and water.

Reactions with oxygen


7

Sodium
Sodium burns in oxygen with an orange flame to produce a white solid mixture of sodium
oxide and sodium peroxide.
For the simple oxide:
For the peroxide:
Magnesium
Magnesium burns in oxygen with an intense white flame to give white solid magnesium
oxide.
Note: If magnesium is burns in air rather than in pure oxygen, it also reacts with
the nitrogen in the air. You get a mixture of magnesium oxide and magnesium nitride
formed.
Aluminium
Aluminium will burn in oxygen if it is powdered, otherwise the strong oxide layer on the
aluminium tends to inhibit the reaction. If you sprinkle aluminium powder into a Bunsen
flame, you get white sparkles. White aluminium oxide is formed.
Silicon
Silicon will burn in oxygen if heated strongly enough. Silicon dioxide is produced.
Note: There is disagreement about the temperature needed to ignite the silicon,
varying from 400C to well over 1000C. It depends on what sort of silicon you are
talking about and how finely divided it is. For example, one of the amorphous (noncrystalline powder) forms of silicon even catches fire spontaneously in air at room
temperature. Other forms need higher temperatures and a richer oxygen supply.
Phosphorus
White phosphorus catches fire spontaneously in air, burning with a white flame and
producing clouds of white smoke - a mixture of phosphorus(III) oxide and phosphorus(V)
oxide.
The proportions of these depend on the amount of oxygen available. In an excess of
oxygen, the product will be almost entirely phosphorus(V) oxide.
For the phosphorus(III) oxide:
For the phosphorus(V) oxide:
Note: You may come across these oxides written as P2O3and P2O5. Don't use these
forms! They are as logical as writing, say, ethene as CH2 and ethane as CH3.

Sulphur

Sulphur burns in air or oxygen on gentle heating with a pale blue flame. It produces
colourless sulphur dioxide gas.
Note: Sulphur dioxide can, of course, be converted further into sulphur trioxide in
the presence of oxygen, but it needs the presence of a catalyst and fairly carefully
controlled conditions.
Chlorine and argon
Despite having several oxides, chlorine won't react directly with oxygen. Argon doesn't
react either.
Reactions with chlorine
Sodium
Sodium burns in chlorine with a bright orange flame. White solid sodium chloride is
produced.
Magnesium
Magnesium burns with its usual intense white flame to give white magnesium chloride.
Aluminium
Aluminium is often reacted with chlorine by passing dry chlorine over aluminium foil
heated in a long tube. The aluminium burns in the stream of chlorine to produce very pale
yellow aluminium chloride. This sublimes (turns straight from solid to vapour and back
again) and collects further down the tube where it is cooler.
Note: You may find versions of this equation showing the aluminium chloride as
Al2Cl6. In fact, this exists in the vapour at temperatures not too far above the sublimation
temperature - not in the solid.
Silicon
If chlorine is passed over silicon powder heated in a tube, it reacts to produce silicon
tetrachloride. This is a colourless liquid which vaporises and can be condensed further
along the apparatus.
Phosphorus
White phosphorus burns spontaneously in chlorine to produce a mixture of two chlorides,
phosphorus(III) chloride and phosphorus(V) chloride (phosphorus trichloride and
phosphorus pentachloride).
Phosphorus(III) chloride is a colourless fuming liquid.
Phosphorus(V) chloride is an off-white (going towards yellow) solid.
Note: These equations are often given starting from P rather than P4. It depends
which form of phosphorus you are talking about. If you are talking about white
phosphorus, P4 is the correct version. If you are talking about red phosphorus, then P is
correct. Red phosphorus has a different (polymeric) structure, and P4 would be wrong for
it.
9

Sulphur
If a stream of chlorine is passed over some heated sulphur, it reacts to form an orange,
evil-smelling liquid, disulphur dichloride, S2Cl2.
Chlorine and argon
It obviously doesn't make sense to talk about chlorine reacting with itself, and argon
doesn't react with chlorine.

1.3 Physical Properties Of The Period 3 Oxides


A Quick Summary Of The Trends Of Period 3 Oxides
The oxides we'll be looking at are:
Na2O
MgO
Al2O3
SiO2
P4O10
SO3
Cl2O7
P4O6
SO2
Cl2O
Those oxides in the top row are known as the highest oxides of the various elements.
These are the oxides where the Period 3 elements are in their highest oxidation states. In
these oxides, all the outer electrons in the Period 3 element are being involved in the
bonding - from just the one with sodium, to all seven of chlorine's outer electrons.
The structures
The trend in structure is from the metallic oxides containing giant structures of ions on the
left of the period via a giant covalent oxide (silicon dioxide) in the middle to molecular
oxides on the right.
Melting and boiling points
The giant structures (the metal oxides and silicon dioxide) will have high melting and
boiling points because a lot of energy is needed to break the strong bonds (ionic or
covalent) operating in three dimensions.
The oxides of phosphorus, sulphur and chlorine consist of individual molecules - some
small and simple; others polymeric. The attractive forces between these molecules will be
van der Waals dispersion and dipole-dipole interactions. These vary in size depending on
the size, shape and polarity of the various molecules - but will always be much weaker
than the ionic or covalent bonds you need to break in a giant structure. These oxides tend
to be gases, liquids or low melting point solids.
Electrical conductivity
None of these oxides has any free or mobile electrons. That means that none of them will
conduct electricity when they are solid.
The ionic oxides can, however, undergo electrolysis when they are molten. They can
conduct electricity because of the movement of the ions towards the electrodes and the
discharge of the ions when they get there.

The metallic oxides


The structures
10

Sodium, magnesium and aluminium oxides consist of giant structures containing metal
ions and oxide ions. Magnesium oxide has a structure just like sodium chloride. The other
two have more complicated arrangements of the ions beyond the scope of syllabuses at
this level.
Melting and boiling points
There are strong attractions between the ions in each of these oxides and these attractions
need a lot of heat energy to break. These oxides therefore have high melting and boiling
points.
Problems:
i. You would expect that the greater the charge, the greater the attractions.
Unfortunately, the oxide with the highest melting and boiling point is magnesium
oxide, not aluminium oxide! So that theory bit the dust! The reason for this
probably lies in the increase in electronegativity as you go from sodium to
magnesium to aluminium. That would mean that the electronegativity
difference between the metal and the oxygen is decreasing. The smaller difference
means that the bond won't be so purely ionic. It is also likely that molten aluminium
oxide contains complex ions containing both aluminium and oxygen rather than
simple aluminium and oxide ions.
ii. The other problem lies with sodium oxide. Most sources say that this sublimes
(turns straight from solid to vapour) at 1275C. However, the melting point of
sodium oxide is 1132C followed by a decomposition temperature (before boiling)
of 1950C. Other sources talk about it decomposing (to sodium and sodium
peroxide) above 400C.
Electrical conductivity
None of these conducts electricity in the solid state, but electrolysis is possible if they are
molten. They conduct electricity because of the movement and discharge of the ions
present. The only important example of this is in the electrolysis of aluminium oxide in
the manufacture of aluminium.
Whether you can electrolyse molten sodium oxide depends, of course, on whether it
actually melts instead of subliming or decomposing under ordinary circumstances. If it
sublimes, you won't get any liquid to electrolyse!
Magnesium and aluminium oxides have melting points far too high to be able to
electrolyse them in a simple lab.
Silicon dioxide (silicon(IV) oxide)
The structure
The electronegativity of the elements increases as you go across the period, and by the
time you get to silicon, there isn't enough electronegativity difference between the silicon
and the oxygen to form an ionic bond. Silicon dioxide is a giant covalent structure.
There are three different crystal forms of silicon dioxide. The easiest one to remember and
draw is based on the diamond structure.
Crystalline silicon has the same structure as diamond. To turn it into silicon dioxide, all
you need to do is to modify the silicon structure by including some oxygen atoms.
11

Notice that each silicon atom is bridged to its neighbours by an oxygen atom. Don't forget
that this is just a tiny part of a giant structure extending in all 3 dimensions.
Note: If you want to be fussy, the Si-O-Si bond angles are wrong in this diagram.
In reality the "bridge" from one silicon atom to its neighbour isn't in a straight line, but
via a "V" shape (similar to the shape around the oxygen atom in a water molecule). It's
extremely difficult to draw that convincingly and tidily in a diagram involving this number
of atoms. The simplification is perfectly acceptable.
Melting and boiling points
Silicon dioxide has a high melting point - varying depending on what the particular
structure is (remember that the structure given is only one of three possible structures), but
they are all around 1700C. Very strong silicon-oxygen covalent bonds have to be broken
throughout the structure before melting occurs. Silicon dioxide boils at 2230C.
Because you are talking about a different form of bonding, it doesn't make sense to try to
compare these values directly with the metallic oxides. What you can safely say is that
because the metallic oxides and silicon dioxide have giant structures, the melting and
boiling points are all high.
Electrical conductivity
Silicon dioxide doesn't have any mobile electrons or ions - so it doesn't conduct electricity
either as a solid or a liquid.
The molecular oxides
The structure
Phosphorus, sulphur and chlorine all form oxides which consist of molecules. Some of
these molecules are fairly simple - others are polymeric. We are just going to look at some
of the simple ones.
Melting and boiling points
Melting and boiling points of these oxides will be much lower than those of the metal
oxides or silicon dioxide. The intermolecular forces holding one molecule to its
neighbours will be van der Waals dispersion forces or dipole-dipole interactions. The
strength of these will vary depending on the size of the molecules.
Electrical conductivity
None of these oxides conducts electricity either as solids or as liquids. None of them
contains ions or free electrons.
The phosphorus oxides
Phosphorus has two common oxides:
12

Phosphorus(III) oxide
Phosphorus(III) oxide is a white solid, melting at 24C and boiling at 173C.
The structure of its molecule is best worked out starting from a P4 molecule which is a
little tetrahedron.

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

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

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

13

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

The sulphur oxides


Sulphur has two common oxides, sulphur dioxide (sulphur(IV) oxide), SO2, and sulphur
trioxide (sulphur(VI) oxide), SO3.
Sulphur dioxide
Sulphur dioxide is a colourless gas at room temperature with an easily recognised choking
smell. It consists of simple SO2molecules.

The sulphur uses 4 of its outer electrons to form the double bonds with the oxygen,
leaving the other two as a lone pair on the sulphur. The bent shape of SO2 is due to this
lone pair.
Sulphur trioxide
Pure sulphur trioxide is a white solid with a low melting and boiling point. It reacts very
rapidly with water vapour in the air to form sulphuric acid. That means that if you make
some in the lab, you tend to see it as a white sludge which fumes dramatically in moist air
(forming a fog of sulphuric acid droplets).
Gaseous sulphur trioxide consists of simple SO3 molecules in which all six of the sulphur's
outer electrons are involved in the bonding.

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

14

There are also other polymeric forms in which the SO3molecules join together in long
chains. For example:

Note: It is difficult to draw this convincingly. In fact, on each sulphur atom, one of
the double bonded oxygens is coming out of the diagram towards you, and the other one is
going back in away from you.
The fact that the simple molecules join up in this way to make bigger structures is
what makes the sulphur trioxide a solid rather than a gas.
The chlorine oxides
Chlorine forms several oxides. Here we are just looking at two of them- chlorine(I) oxide,
Cl2O, and chlorine(VII) oxide, Cl2O7.
Chlorine(I) oxide
Chlorine(I) oxide is a yellowish-red gas at room temperature. It consists of simple small
molecules.

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

15

1.4 Acid-Base Behaviour Of The Period 3 Oxides


A quick summary of the trend of The period 3 oxides
The oxides we'll be looking at are:
Na2
O

Mg
O

Al2O SiO P4O1 SO Cl2O


3

P4O6 SO Cl2O
The trend in acidbase behaviour
2
The trend in acidbase behaviour is
shown in various reactions, but as a simple summary:
The trend is from strongly basic oxides on the left-hand side to strongly acidic ones
on the right, via an amphoteric oxide (aluminium oxide) in the middle. An
amphoteric oxide is one which shows both acidic and basic properties.
For this simple trend, you have to be looking only at the highest oxides of the individual
elements. Those are the ones on the top row above, and are where the element is in its
highest possible oxidation state. The pattern isn't so simple if you include the other oxides
as well.
For the non-metal oxides, their acidity is usually thought of in terms of the acidic solutions
formed when they react with water - for example, sulphur trioxide reacting to give
sulphuric acid. They will, however, all react with bases such as sodium hydroxide to form
salts such as sodium sulphate.
These reactions are all explored in detail.
Chemistry of the individual oxides
Sodium oxide
Sodium oxide is a simple strongly basic oxide. It is basic because it contains the oxide ion,
O2-, which is a very strong base with a high tendency to combine with hydrogen ions.
Reaction with water
Sodium oxide reacts exothermically with cold water to produce sodium hydroxide
solution. Depending on its concentration, this will have a pH around 14.
Reaction with acids
As a strong base, sodium oxide also reacts with acids. For example, it would react with
dilute hydrochloric acid to produce sodium chloride solution.
Magnesium oxide
Magnesium oxide is again a simple basic oxide, because it also contains oxide ions.
However, it isn't as strongly basic as sodium oxide because the oxide ions aren't so free.
In the sodium oxide case, the solid is held together by attractions between 1+ and 2- ions.
In the magnesium oxide case, the attractions are between 2+ and 2-. It takes more energy
to break these.
Even allowing for other factors (like the energy released when the positive ions form
attractions with water in the solution formed), the net effect of this is that reactions
involving magnesium oxide will always be less exothermic than those of sodium oxide.
16

Reaction with water


If you shake some white magnesium oxide powder with water, nothing seems to happen it doesn't look as if it reacts. However, if you test the pH of the liquid, you find that it is
somewhere around pH 9 - showing that it is slightly alkaline.
There must have been some slight reaction with the water to produce hydroxide ions in
solution. Some magnesium hydroxide is formed in the reaction, but this is almost
insoluble - and so not many hydroxide ions actually get into solution.
Reaction with acids
Magnesium oxide reacts with acids as you would expect any simple metal oxide to react.
For example, it reacts with warm dilute hydrochloric acid to give magnesium chloride
solution.
Aluminium oxide
Describing the properties of aluminium oxide can be confusing because it exists in a
number of different forms. One of those forms is very unreactive. It is known chemically
as alpha-Al2O3and is produced at high temperatures.
In what follows we are assuming one of the more reactive forms.
Aluminium oxide is amphoteric. It has reactions as both a base and an acid.
Reaction with water
Aluminium oxide doesn't react in a simple way with water in the sense that sodium oxide
and magnesium oxide do, and doesn't dissolve in it. Although it still contains oxide ions,
they are held too strongly in the solid lattice to react with the water.
Reaction with acids
Aluminium oxide contains oxide ions and so reacts with acids in the same way as sodium
or magnesium oxides. That means, for example, that aluminium oxide will react with hot
dilute hydrochloric acid to give aluminium chloride solution.
In this (and similar reactions with other acids), aluminium oxide is showing the basic side
of its amphoteric nature.
Reaction with bases
Aluminium oxide has also got an acidic side to its nature, and it shows this by reacting
with bases such as sodium hydroxide solution.
Various aluminates are formed - compounds where the aluminium is found in the negative
ion. This is possible because aluminium has the ability to form covalent bonds with
oxygen.
In the case of sodium, there is too much electronegativity difference between sodium and
oxygen to form anything other than an ionic bond. But electronegativity increases as you
go across the period - and the electronegativity difference between aluminium and oxygen
is smaller. That allows the formation of covalent bonds between the two.
With hot, concentrated sodium hydroxide solution, aluminium oxide reacts to give a
colourless solution of sodium tetrahydroxoaluminate.

17

Silicon dioxide (silicon(IV) oxide)


By the time you get to silicon as you go across the period, electronegativity has increased
so much that there is no longer enough electronegativity difference between silicon and
oxygen to form ionic bonds.
Silicon dioxide has no basic properties - it doesn't contain oxide ions and it doesn't react
with acids. Instead, it is very weakly acidic, reacting with strong bases.
Reaction with water
Silicon dioxide doesn't react with water, because of the difficulty of breaking up the giant
covalent structure.
Reaction with bases
Silicon dioxide reacts with sodium hydroxide solution, but only if it is hot and
concentrated. A colourless solution of sodium silicate is formed.
You may also be familiar with one of the reactions happening in the Blast Furnace
extraction of iron - in which calcium oxide (from the limestone which is one of the raw
materials) reacts with silicon dioxide to produce a liquid slag, calcium silicate. This is also
an example of the acidic silicon dioxide reacting with a base.
Important! For the remainder of the oxides, we are mainly going to be considering
the results of reacting them with water to give solutions of various acids.
When we talk about the acidity of the oxides increasing as you go from, say,
phosphorus(V) oxide to sulphur trioxide to chlorine(VII) oxide, what we are normally
talking about is the increasing strengths of the acids formed when they react with water.
The phosphorus oxides
We are going to be looking at two phosphorus oxides, phosphorus(III) oxide, P4O6, and
phosphorus(V) oxide, P4O10.
Phosphorus(III) oxide
Phosphorus(III) oxide reacts with cold water to give a solution of the weak acid, H3PO3 known variously as phosphorous acid, orthophosphorous acid or phosphonic acid. Its
reaction with hot water is much more complicated.
The pure un-ionised acid has the structure:

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

18

It is pretty unlikely that you would ever react phosphorus(III) oxide directly with a base,
but you might need to know what happens if you react the phosphorous acid formed with
a base.
In phosphorous acid, the two hydrogen atoms in the -OH groups are acidic, but the other
one isn't. That means that you can get two possible reactions with, for example, sodium
hydroxide solution depending on the proportions used.
In the first case, only one of the acidic hydrogens has reacted with the hydroxide ions
from the base. In the second case (using twice as much sodium hydroxide), both have
reacted.
If you were to react phosphorus(III) oxide directly with sodium hydroxide solution rather
than making the acid first, you would end up with the same possible salts.
Phosphorus(V) oxide
Phosphorus(V) oxide reacts violently with water to give a solution containing a mixture of
acids, the nature of which depends on the conditions. We usually just consider one of
these, phosphoric(V) acid, H3PO4 - also known just as phosphoric acid or as
orthophosphoric acid.
This time the pure un-ionised acid has the structure:

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

Again, if you were to react phosphorus(V) oxide directly with sodium hydroxide solution
rather than making the acid first, you would end up with the same possible salts.
This is getting ridiculous, and so I will only give one example out of the possible
equations:
The sulphur oxides
19

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

Sulphurous acid is also a weak acid with a pKa of around 1.8 - very slightly stronger than
the two phosphorus-containing acids above. A reasonably concentrated solution of
sulphurous acid will again have a pH of about 1.
Sulphur dioxide will also react directly with bases such as sodium hydroxide solution. If
sulphur dioxide is bubbled through sodium hydroxide solution, sodium sulphite solution is
formed first followed by sodium hydrogensulphite solution when the sulphur dioxide is in
excess.
Note: Sodium sulphite is also called sodium sulphate(IV). Sodium
hydrogensulphite is also sodium hydrogensulphate(IV) or sodium bisulphite.
Notice that the equations for these reactions are different from the phosphorus examples.
In this case, we are reacting the oxide directly with the sodium hydroxide, because that's
the way we are most likely to do it. If you want to, it isn't difficult to work out equations
for the reactions between sodium hydroxide and sulphurous acid in exactly the same way
as we did with the phosphorus acids.
Another important reaction of sulphur dioxide is with the base calcium oxide to form
calcium sulphite (calcium sulphate(IV)). This is at the heart of one of the methods of
removing sulphur dioxide from flue gases in power stations.
Sulphur trioxide
Sulphur trioxide reacts violently with water to produce a fog of concentrated sulphuric
acid droplets.
Note: If you know about the Contact Process for the manufacture of sulphuric acid,
you will know that the sulphur trioxide is always converted into sulphuric acid by a
round-about process to avoid the problem of the sulphuric acid fog.
Pure un-ionised sulphuric acid has the structure:

20

Sulphuric acid is a strong acid, and solutions will typically have pH's of around 0.
The acid reacts with water to give a hydroxonium ion (a hydrogen ion in solution, if you
like) and a hydrogensulphate ion. This reaction is virtually 100% complete.
The second hydrogen is more difficult to remove. In fact the hydrogensulphate ion is a
relatively weak acid - similar in strength to the acids we have already discussed. This time
you get an equilibrium:
You probably won't need this, but it is useful if you understand the reason that sulphuric
acid is a stronger acid than sulphurous acid. You can apply the same reasoning to other
acids that you find ahead.
When a hydrogen ion is lost from one of the -OH groups, the negative charge left on the
oxygen is spread out (delocalised) over the ion by interacting with the doubly-bonded
oxygens.
It follows that the more of these you have, the more delocalisation you can get - and the
more delocalisation, the more stable the ion becomes. The more stable the ion, the less
likely it is to recombine with a hydrogen ion and revert to the un-ionised acid.
Sulphurous acid only has one doubly-bonded oxygen, whereas sulphuric acid has two that makes for a much more effective delocalisation, a much more stable ion, and so for a
stronger acid.
Sulphuric acid, of course, has all the reactions of a strong acid that you are familiar with
from introductory chemistry courses. For example, the normal reaction with sodium
hydroxide solution is to form sodium sulphate solution - in which both of the acidic
hydrogens react with hydroxide ions.
In principle, you can also get sodium hydrogensulphate solution by using half as much
sodium hydroxide and just reacting with one of the two acidic hydrogens in the acid. In
practice, I personally have never ever done it - I can't at the moment see much point!
Sulphur trioxide itself will also react directly with bases to form sulphates. For example, it
will react with calcium oxide to form calcium sulphate. This is just like the reaction with
sulphur dioxide described above.
The chlorine oxides
Chlorine forms several oxides but we will discuss only two of them- chlorine(VII) oxide,
Cl2O7, and chlorine(I)oxide, Cl2O. Chlorine(VII) oxide is also known as dichlorine
heptoxide, and chlorine(I) oxide as dichlorine monoxide.Chlorine(VII) oxide
Chlorine(VII) oxide is the highest oxide of chlorine - the chlorine is in its maximum
oxidation state of +7. It continues the trend of the highest oxides of the Period 3 elements
towards being stronger acids.
Chlorine(VII) oxide reacts with water to give the very strong acid, chloric(VII) acid - also
known as perchloric acid. The pH of typical solutions will, like sulphuric acid, be around
0.
21

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

When the chlorate(VII) ion (perchlorate ion) forms by loss of a hydrogen ion (when it
reacts with water, for example), the charge can be delocalised over every oxygen atom in
the ion. That makes it very stable, and means that chloric(VII) acid is very strong.
Chloric(VII) acid reacts with sodium hydroxide solution to form a solution of sodium
chlorate(VII).
Chlorine(VII) oxide itself also reacts with sodium hydroxide solution to give the same
product.
Chlorine(I) oxide
Chlorine(I) oxide is far less acidic than chlorine(VII) oxide. It reacts with water to some
extent to give chloric(I) acid, HOCl - also known as hypochlorous acid.
Note: You may also find the chloric(I) acid written as HClO. The form used above
more accurately reflects the way the atoms are joined up.
The structure of chloric(I) acid is exactly as shown by its formula, HOCl. It has no
doubly-bonded oxygens, and no way of delocalising the charge over the negative ion
formed by loss of the hydrogen.
That means that the negative ion formed isn't very stable, and readily reclaims its
hydrogen to revert to the acid. Chloric(I) acid is very weak (pKa = 7.43).
Chloric(I) acid reacts with sodium hydroxide solution to give a solution of sodium
chlorate(I) (sodium hypochlorite).
Chlorine(I) oxide also reacts directly with sodium hydroxide to give the same product.

1.5 PROPERTIES OF THE PERIOD 3 CHLORIDES


A Quick Summary Of The Trends Of Chlorides Of Period 3
The chlorides we'll be looking at are:
NaCl
MgCl2
AlCl3

SiCl4

PCl5
PCl3

There are three chlorides of sulphur, but we will only discuss is S2Cl2.
As you will see later, aluminium chloride exists in some circumstances as a dimer, Al2Cl6.
The structures
22

S2Cl

Sodium chloride and magnesium chloride are ionic and consist of giant ionic lattices at
room temperature
Aluminium chloride and phosphorus(V) chloride are tricky! They change their structure
from ionic to covalent when the solid turns to a liquid or vapour.
The others are simple covalent molecules.
Melting and boiling points
Sodium and magnesium chlorides are solids with high melting and boiling points because
of the large amount of heat which is needed to break the strong ionic attractions.
The rest are liquids or low melting point solids. Leaving aside the aluminium chloride and
phosphorus(V) chloride cases where the situation is quite complicated, the attractions in
the others will be much weaker intermolecular forces such as van der Waals dispersion
forces. These vary depending on the size and shape of the molecule, but will always be far
weaker than ionic bonds.
Electrical conductivity
Sodium and magnesium chlorides are ionic and so will undergoelectrolysis when they are
molten. Electricity is carried by the movement of the ions and their discharge at the
electrodes.
In the aluminium chloride and phosphorus(V) chloride cases, the solid doesn't conduct
electricity because the ions aren't free to move. In the liquid (where it exists - both of these
sublime at ordinary pressures), they have converted into a covalent form, and so don't
conduct either.
The rest of the chlorides don't conduct electricity either solid or molten because they don't
have any ions or any mobile electrons.
Reactions with water
As an approximation, the simple ionic chlorides (sodium and magnesium chloride) just
dissolve in water.
The other chlorides all react with water in a variety of ways described below for each
individual chloride. The reaction with water is known as hydrolysis.
The individual chlorides
Sodium chloride, NaCl
Sodium chloride is a simple ionic compound consisting of a giant array of sodium and
chloride ions.
A small representative bit of a sodium chloride lattice looks like this:

This is normally drawn in an exploded form as:

23

The strong attractions between the positive and negative ions need a lot of heat energy to
break, and so sodium chloride has high melting and boiling points.
It doesn't conduct electricity in the solid state because it hasn't any mobile electrons and
the ions aren't free to move. However, when it melts it undergoes electrolysis.
Sodium chloride simply dissolves in water to give a neutral solution.
Magnesium chloride, MgCl2
Magnesium chloride is also ionic, but with a more complicated arrangement of the ions to
allow for having twice as many chloride ions as magnesium ions.
Again, lots of heat energy is needed to overcome the attractions between the ions, and so
the melting and boiling points are again high.
Note: There is a problem here, though! You would expect the attractions between
magnesium ions and chloride ions to be greater than those between sodium and chloride
ions due to the extra charge on the magnesium. However, magnesium chloride melts at
a lower temperature than sodium chloride, and the boiling points are almost identical (to
within one degree).
The most likely explanation for this is that magnesium chloride is less purely ionic than
we normally suggest, and shows some small degree of covalency. That means that you
can't make a simple comparison between the melting and boiling points of magnesium
chloride and the more purely ionic sodium chloride.
The most purely ionic of the magnesium halides is magnesium fluoride, because that has
the greatest electronegativity difference between the magnesium and the halogen. In fact,
magnesium fluoride has significantly higher melting and boiling points than sodium
fluoride, which is what you would expect from the greater attractions due to the extra
charge on the magnesium ion.
Solid magnesium chloride is a non-conductor of electricity because the ions aren't free to
move. However, it undergoes electrolysis when the ions become free on melting.
Magnesium chloride dissolves in water to give a faintly acidic solution (pH =
approximately 6).
When magnesium ions are broken off the solid lattice and go into solution, there is enough
attraction between the 2+ ions and the water molecules to get co-ordinate (dative covalent)
bonds formed between the magnesium ions and lone pairs on surrounding water
molecules.
Hexaaquamagnesium ions are formed, [Mg(H2O)6]2+.

24

Ions of this sort are acidic - the degree of acidity depending on how much the electrons in
the water molecules are pulled towards the metal at the centre of the ion. The hydrogens
are made rather more positive than they would otherwise be, and more easily pulled off by
a base.
In the magnesium case, the amount of distortion is quite small, and only a small
proportion of the hydrogen atoms are removed by a base - in this case, by water molecules
in the solution.
Note: The reason for the colour-coding is to try to avoid confusion between the
water molecules attached to the ion and those in the solution.
The presence of the hydroxonium ions in the solution causes it to be acidic. The fact that
there aren't many of them formed (the position of equilibrium lies well to the left), means
that the solution is only weakly acidic.
You may also find the last equation in a simplified form:
Hydrogen ions in solution are hydroxonium ions. If you use this form, it is essential to
include the state symbols.
Aluminium chloride, AlCl3
Electronegativity increases as you go across the period and, by the time you get to
aluminium, there isn't enough electronegativity difference between aluminium and
chlorine for there to be a simple ionic bond.
Aluminium chloride is complicated by the way its structure changes as temperature
increases.
At room temperature, the aluminium in aluminium chloride is 6-coordinated. That means
that each aluminium is surrounded by 6 chlorines. The structure is an ionic lattice although with a lot of covalent character.
At ordinary atmospheric pressure, aluminium chloride sublimes (turns straight from solid
to vapour) at about 180C. If the pressure is raised to just over 2 atmospheres, it melts
instead at a temperature of 192C.
Both of these temperatures, of course, are completely wrong for an ionic compound - they
are much too low. They suggest comparatively weak attractions between molecules - not
strong attractions between ions.
The coordination of the aluminium changes at these temperatures. It becomes 4coordinated - each aluminium now being surrounded by 4 chlorines rather than 6.
What happens is that the original lattice has converted into Al2Cl6 molecules. The structure
of this is:

25

This conversion means, of course, that you have completely lost any ionic character which is why the aluminium chloride vaporises or melts (depending on the pressure).
There is an equilibrium between these dimers and simple AlCl3molecules. As the
temperature increases further, the position of equilibrium shifts more and more to the
right.
Summary
At room temperature, solid aluminium chloride has an ionic lattice with a lot of
covalent character.
At temperatures around 180 - 190C (depending on the pressure), aluminium
chloride coverts to a molecular form, Al2Cl6. This causes it to melt or vaporise
because there are now only comparatively weak intermolecular attractions.
As the temperature increases a bit more, it increasingly breaks up into simple
AlCl3 molecules.
Solid aluminium chloride doesn't conduct electricity at room temperature because the ions
aren't free to move. Molten aluminium chloride (only possible at increased pressures)
doesn't conduct electricity because there aren't any ions any more.
The reaction of aluminium chloride with water is dramatic. If you drop water onto solid
aluminium chloride, you get a violent reaction producing clouds of steamy fumes of
hydrogen chloride gas.
If you add solid aluminium chloride to an excess of water, it still splutters, but instead of
hydrogen chloride gas being given off, you get an acidic solution formed. A solution of
aluminium chloride of ordinary concentrations (around 1 mol dm-3, for example) will have
a pH around 2 - 3. More concentrated solutions will go lower than this.
The aluminium chloride reacts with the water rather than just dissolving in it. In the first
instance, hexaaquaaluminium ions are formed together with chloride ions.
You will see that this is very similar to the magnesium chloride equation given above - the
only real difference is the charge on the ion.
That extra charge pulls electrons from the water molecules quite strongly towards the
aluminium. That makes the hydrogens more positive and so easier to remove from the ion.
In other words, this ion is much more acidic than in the corresponding magnesium case.
These equilibria (whichever you choose to write) lie further to the right, and so the
solution formed is more acidic - there are more hydroxonium ions in it.
or, more simply:
We haven't so far accounted for the burst of hydrogen chloride formed if there isn't much
water present.
All that happens is that because of the heat produced in the reaction and the concentration
of the solution formed, hydrogen ions and chloride ions in the mixture combine together
as hydrogen chloride molecules and are given off as a gas. With a large excess of water,
the temperature never gets high enough for that to happen - the ions just stay in solution.
26

Silicon tetrachloride, SiCl4


Silicon tetrachloride is a simple no-messing-about covalent chloride. There isn't enough
electronegativity difference between the silicon and the chlorine for the two to form ionic
bonds.
Silicon tetrachloride is a colourless liquid at room temperature which fumes in moist air.
The only attractions between the molecules are van der Waals dispersion forces.
It doesn't conduct electricity because of the lack of ions or mobile electrons.
It fumes in moist air because it reacts with water in the air to produce hydrogen chloride.
If you add water to silicon tetrachloride, there is a violent reaction to produce silicon
dioxide and fumes of hydrogen chloride. In a large excess of water, the hydrogen chloride
will, of course, dissolve to give a strongly acidic solution containing hydrochloric acid.
The phosphorus chlorides
There are two phosphorus chlorides - phosphorus(III) chloride, PCl3, and phosphorus(V)
chloride, PCl5.
Phosphorus(III) chloride (phosphorus trichloride), PCl3
This is another simple covalent chloride - again a fuming liquid at room temperature.
It is a liquid because there are only van der Waals dispersion forces and dipole-dipole
attractions between the molecules.
Note: The phosphorus(III) chloride molecule has a permanent dipole, which is why
dipole-dipole attractions are possible. The phosphorus(III) chloride case is rather similar
to CHCl3, except that there is a lone pair of electrons at the top of the molecule rather than
a hydrogen atom.
It doesn't conduct electricity because of the lack of ions or mobile electrons.
Phosphorus(III) chloride reacts violently with water. You get phosphorous acid, H3PO3,
and fumes of hydrogen chloride (or a solution containing hydrochloric acid if lots of water
is used).
Note: Phosphorous acid is also known as orthophosphorous acid or as phosphonic
acid. Notice the "-ous" ending in the first two names. That's not a spelling mistake - it's for
real! It is used to distinguish it from phosphoric acid which is quite different (see below).
Phosphorus(V) chloride (phosphorus pentachloride), PCl5
Unfortunately, phosphorus(V) chloride is structurally more complicated.
Phosphorus(V) chloride is a white solid which sublimes at 163C. The higher the
temperature goes above that, the more the phosphorus(V) chloride dissociates (splits up
reversibly) to give phosphorus(III) chloride and chlorine.
Solid phosphorus(V) chloride contains ions - which is why it is a solid at room
temperature. The formation of the ions involves two molecules of PCl5.
A chloride ion transfers from one of the original molecules to the other, leaving a positive
ion, [PCl4]+, and a negative ion, [PCl6]-.

27

At 163C, the phosphorus(V) chloride converts to a simple molecular form containing


PCl5 molecules. Because there are only van der Waals dispersion forces between these, it
then vaporises.
Solid phosphorus(V) chloride doesn't conduct electricity because the ions aren't free to
move.
Note: Phosphorus(V) chloride does, however, undergo electrolysis in a suitable
solvent which it doesn't react with.
Phosphorus(V) chloride has a violent reaction with water producing fumes of hydrogen
chloride. As with the other covalent chlorides, if there is enough water present, these will
dissolve to give a solution containing hydrochloric acid.
The reaction happens in two stages. In the first, with cold water, phosphorus oxychloride,
POCl3, is produced along with HCl.
If the water is boiling, the phosphorus(V) chloride reacts further to give phosphoric(V)
acid and more HCl. Phosphoric(V) acid is also known just as phosphoric acid or as
orthophosphoric acid.
The overall equation in boiling water is just a combination of these:
Disulphur dichloride, S2Cl2
Disulphur dichloride is just one of three sulphur chlorides. This is possibly because it is
the one which is formed when chlorine reacts with hot sulphur.
Disulphur dichloride is a simple covalent liquid - orange and smelly!
The shape is surprisingly difficult to draw convincingly! The atoms are all joined up in a
line - but twisted:

The reason for drawing the shape is to give a hint about what sort of intermolecular
attractions are possible. There is no plane of symmetry in the molecule and that means that
it will have an overall permanent dipole.
The liquid will have van der Waals dispersion forces and dipole-dipole attractions.
There are no ions in disulphur dichloride and no mobile electrons - so it never conducts
electricity.
Disulphur dichloride reacts slowly with water to produce a complex mixture of things
including hydrochloric acid, sulphur, hydrogen sulphide and various sulphur-containing
28

acids and anions (negative ions). There is no way that you can write a single equation for
this - and one would never be expected in an exam.
1.6 PROPERTIES OF THE PERIOD 3 "HYDROXIDES"
A quick summary of the trends
Sodium and magnesium hydroxides
These contain hydroxide ions, and are simple basic hydroxides.
Aluminium hydroxide
Aluminium hydroxide, like aluminium oxide, is amphoteric - it has both basic and acidic
properties.
The other "hydroxides"
All of these have -OH groups covalently bound to the atom from period 3. These
compounds are all acidic - ranging from the very weakly acidic silicic acids (one of which
is shown below) to the very strong sulphuric or chloric(VII) acids.

There are other acids (also containing -OH groups) formed by these elements, but these
are the ones where the Period 3 element is in its highest oxidation state.
Adding some detail
Sodium and magnesium hydroxides
These are both basic because they contain hydroxide ions - a strong base.
Both react with acids to form salts. For example, with dilute hydrochloric acid, you get
colourless solutions of sodium chloride or magnesium chloride.
Aluminium hydroxide
Aluminium hydroxide is amphoteric.
Like sodium or magnesium hydroxides, it will react with acids. This is showing the basic
side of its nature.
With dilute hydrochloric acid, a colourless solution of aluminium chloride is formed.

29

But aluminium hydroxide also has an acidic side to its nature. It will react with sodium
hydroxide solution to give a colourless solution of sodium tetrahydroxoaluminate.
Note: You may find all sorts of other formulae given for the product from this
reaction. These range from NaAlO2 (which is a dehydrated form of the one in the
equation) to Na3Al(OH)6 (which is a different product altogether).
What you actually get will depend on things like the temperature and the concentration of
the sodium hydroxide solution.
The other "hydroxides"
A quick reminder of what we are talking about here:

None of these contains hydroxide ions. In each case the -OH group is covalently bound to
the Period 3 element, and in each case it is possible for the hydrogens on these -OH
groups to be removed by a base. In other words, all of these compounds are acidic.
But they vary considerably in strength:
Orthosilicic acid is very weak indeed.
Phosphoric(V) acid is a weak acid - although somewhat stronger than simple
organic acids like ethanoic acid.
Sulphuric acid and chloric(VII) acids are both very strong acids.
The main factor in determining the strength of the acid is how stable the anion (the
negative ion) is once the hydrogen has been removed. This in turn depends on how much
the negative charge can be spread around the rest of the ion.
If the negative charge stays entirely on the oxygen atom left behind from the -OH group, it
will be very attractive to hydrogen ions. The lost hydrogen ion will be easily recaptured
and the acid will be weak.
On the other hand, if the charge can be spread out (delocalised) over the whole of the ion,
it will be so "dilute" that it won't attract the hydrogen back very easily. The acid will then
be strong.
30

Wherever possible, the negative charge is delocalised by interacting with doubly-bonded


oxygens.
For example, in chloric(VII) acid, the ion produced is the chlorate(VII) ion (also known as
the perchlorate ion), ClO4-.
The structure of the ion doesn't stay like this:

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

When sulphuric acid loses a hydrogen ion to form the hydrogensulphate ion, HSO4-, the
charge can be spread over three oxygens (the original one with the negative charge, and
the two sulphur-oxygen double bonds. That's still an effective delocalisation, and
sulphuric acid is almost as strong as chloric(VII) acid.
Note: Sulphuric acid can, of course, lose a second hydrogen ion as well from the
other -OH group and form sulphate ions. However, that is a bit more difficult. If you lose
that second hydrogen, you can use all four oxygens to delocalise the charge - but now you
have to delocalise two negative charges rather than just one. The hydrogensulphate ion
isn't a strong acid. It's strength is similar to phosphoric(V) acid.
Phosphoric(V) acid is much weaker than sulphuric acid because it only has one
phosphorus-oxygen double bond which it can use to help delocalise the charge on the ion
formed by losing one hydrogen ion - so the charge on that ion is delocalised less
effectively.
In orthosilicic acid, there aren't any silicon-oxygen double bonds to delocalise the charge.
That means the ion formed by loss of a hydrogen ion isn't at all stable, and easily recovers
its hydrogen.

31

You might also like