Anda di halaman 1dari 35

UNIT 1

MODULE 1: Fundamentals in Chemistry

2. FORCES OF ATTRACTION

2.1 – 2.7

View ppt - Bonding

What are intermolecular attractions?

Intermolecular versus intramolecular bonds

Intermolecular attractions are attractions between one molecule and a


neighboring molecule. The forces of attraction which hold an individual
molecule together (for example, the covalent bonds) are known as
intramolecular attractions.

All molecules experience intermolecular attractions, although in some cases


those attractions are very weak. Even in a gas like hydrogen, H 2, if you slow
the molecules down by cooling the gas, the attractions are large enough for
the molecules to stick together eventually to form a liquid and then a solid.

In hydrogen's case the attractions are so weak that the molecules have to be
cooled to 21 K (-252°C) before the attractions are enough to condense the
hydrogen as a liquid. Helium's intermolecular attractions are even weaker -
the molecules won't stick together to form a liquid until the temperature
drops to 4 K (-269°C).

Van der Waals forces: dispersion forces

Dispersion forces are also known as "London forces" (named after Fritz
London who first suggested how they might arise).

The origin of van der Waals dispersion forces

Temporary fluctuating dipoles

Attractions are electrical in nature. In a symmetrical molecule like hydrogen,


however, there doesn't seem to be any electrical distortion to produce
positive or negative parts. But that's only true on average.

40
The lozenge-shaped diagram represents a small symmetrical molecule - H 2,
perhaps, or Br2. The even shading shows that on average there is no
electrical distortion.

But the electrons are mobile, and at any one instant they might find
themselves towards one end of the molecule, making that end -. The other
end will be temporarily short of electrons and so becomes +.

An instant later the electrons may well have moved up to the other end,
reversing the polarity of the molecule.

This constant "sloshing around" of the electrons in the molecule causes


rapidly fluctuating dipoles even in the most symmetrical molecule. It even
happens in monatomic molecules - molecules of noble gases, like helium,
which consist of a single atom.

If both the helium electrons happen to be on one side of the atom at the
same time, the nucleus is no longer properly covered by electrons for that
instant.

How temporary dipoles give rise to intermolecular attractions

I'm going to use the same lozenge-shaped diagram now to represent any
molecule which could, in fact, be a much more complicated shape. Shape
does matter (see below), but keeping the shape simple makes it a lot easier
to both draw the diagrams and understand what is going on.

Imagine a molecule which has a temporary polarity being approached by one


which happens to be entirely non-polar just at that moment. (A pretty unlikely
event, but it makes the diagrams much easier to draw! In reality, one of the
molecules is likely to have a greater polarity than the other at that time - and
so will be the dominant one.)

41
As the right hand molecule approaches, its electrons will tend to be attracted
by the slightly positive end of the left hand one.

This sets up an induced dipole in the approaching molecule, which is


orientated in such a way that the + end of one is attracted to the - end of
the other.

An instant later the electrons in the left hand molecule may well have moved
up the other end. In doing so, they will repel the electrons in the right hand
one.

The polarity of both molecules reverses, but you still have + attracting -. As
long as the molecules stay close to each other the polarities will continue to
fluctuate in synchronization so that the attraction is always maintained.

There is no reason why this has to be restricted to two molecules. As long as


the molecules are close together this synchronized movement of the
electrons can occur over huge numbers of molecules.

This diagram shows how a whole lattice of molecules could be held together
in a solid using van der Waals dispersion forces. An instant later, of course,
you would have to draw a quite different arrangement of the distribution of
the electrons as they shifted around - but always in synchronization.

The strength of dispersion forces

Dispersion forces between molecules are much weaker than the covalent
bonds within molecules. It isn't possible to give any exact value, because the
size of the attraction varies considerably with the size of the molecule and its
shape.

42
How molecular size affects the strength of the dispersion forces

The boiling points of the noble gases are

-
helium
269°C
-
neon
246°C
-
argon
186°C
krypto -
n 152°C
-
xenon
108°C
radon -62°C

All of these elements have monatomic molecules.

The reason that the boiling points increase as you go down the group is that
the number of electrons increases, and so also does the radius of the atom.
The more electrons you have, and the more distance over which they can
move, the bigger the possible temporary dipoles and therefore the bigger the
dispersion forces.

Because of the greater temporary dipoles, xenon molecules are "stickier"


than neon molecules. Neon molecules will break away from each other at
much lower temperatures than xenon molecules - hence neon has the lower
boiling point.

This is the reason that (all other things being equal) bigger molecules have
higher boiling points than small ones. Bigger molecules have more electrons
and more distance over which temporary dipoles can develop - and so the
bigger molecules are "stickier".

How molecular shape affects the strength of the dispersion forces

The shapes of the molecules also matter. Long thin molecules can develop
bigger temporary dipoles due to electron movement than short fat ones
containing the same numbers of electrons.

Long thin molecules can also lie closer together - these attractions are at
their most effective if the molecules are really close.

43
For example, the hydrocarbon molecules butane and 2-methylpropane both
have a molecular formula C4H10, but the atoms are arranged differently. In
butane the carbon atoms are arranged in a single chain, but 2-methylpropane
is a shorter chain with a branch.

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

van der Waals forces: dipole-dipole interactions

A molecule like HCl has a permanent dipole because chlorine is more


electronegative than hydrogen. These permanent, in-built dipoles will cause
the molecules to attract each other rather more than they otherwise would if
they had to rely only on dispersion forces.

It's important to realize that all molecules experience dispersion forces.


Dipole-dipole interactions are not an alternative to dispersion forces - they
occur in addition to them. Molecules which have permanent dipoles will
therefore have boiling points rather higher than molecules which only have
temporary fluctuating dipoles.

Surprisingly dipole-dipole attractions are fairly minor compared with


dispersion forces, and their effect can only really be seen if you compare two
molecules with the same number of electrons and the same size. For
example, the boiling points of ethane, CH 3CH3, and fluoromethane, CH3F, are

Why choose these two molecules to compare? Both have


identical numbers of electrons, and if you made models you would find that
the sizes were similar - as you can see in the diagrams. That means that the
dispersion forces in both molecules should be much the same.

44
The higher boiling point of fluoromethane is due to the large permanent
dipole on the molecule because of the high electronegativity of fluorine.
However, even given the large permanent polarity of the molecule, the
boiling point has only been increased by some 10°.

Here is another example showing the dominance of the dispersion forces.


Trichloromethane, CHCl3, is a highly polar molecule because of the
electronegativity of the three chlorines. There will be quite strong dipole-
dipole attractions between one molecule and its neighbors.

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


molecule is uniformly - in all directions. CCl4 has to rely only on dispersion
forces.

So which has the highest boiling point? CCl4


does, because it is a bigger molecule with more
electrons. The increase in the dispersion forces more
than compensates for the loss of dipole-dipole
interactions.

The boiling points are:

61.2°
CHCl3
C
76.8°
CCl4
C

THE CAPE 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

45
You may have come across some of the following ions in a basic course like
CSEC. 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
[Xe]4f145d106
Pb2+
s2

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

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 aluminum 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.

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 ionization


energy. The more electrons you remove, the greater the total ionization
energy becomes. Eventually the total ionization 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.

46
For example, why is calcium chloride CaCl 2 rather than CaCl or CaCl3?

If one mole of CaCl (containing Ca + ions) is made from its elements, it is


possible to estimate that about 171 kJ of heat is evolved.

However, making 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 ionization 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.

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. 6

Covalent bonding at CAPE

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

If you stick closely to the CAPE syllabus, there is little need to move far from
the simple (CSEC) 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 PCl3, 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.

47
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 BF 3, three bonds is the
maximum possible because boron only has 3 electrons to share.

Phosphorus(V) chloride, PCl5

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

Phosphorus forms two chlorides - PCl 3 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.

48
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 CSEC, but PCl5 would have looked very worrying.

Why does phosphorus sometimes break away from a noble gas structure and
form five bonds? In order to answer that question, we need to explore
territory beyond the limits of current CAPE syllabus. 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

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


methane?

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

There is a serious mis-match between this structure and the modern


electronic structure of carbon, 1s22s22px12py1. The modern structure shows
that there are only 2 unpaired electrons for hydrogens to share with, instead
of the 4 which the simple view requires.

You can see this more readily using the electrons-


in-boxes notation. Only the 2-level electrons are

49
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.

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.

Hybridization

The electrons rearrange themselves again in a process


called hybridization. This reorganizes the electrons into
four identical hybrid orbitals called sp 3 hybrids
(because they are made from one s orbital and three p
orbitals). You should read "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

50
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


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

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:

51
Phosphorus has the electronic structure 1s 22s22p63s23px13py13pz1. 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 hybridize to produce 4 equivalent sp 3 hybrids
just like in carbon - except that one of these hybrid orbitals contains a lone
pair of electrons.

Each of the 3 chlorines then forms a covalent bond by merging the atomic
orbital containing its unpaired electron with one of the phosphorus unpaired
electrons to make 3 molecular orbitals.

You might wonder whether all this is worth the bother! Probably 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.

52
If the phosphorus is going to form PCl 5 it has first to generate 5 unpaired
electrons. It does this by promoting one of the electrons in the 3s orbital to
the next available higher energy orbital.

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

This leaves the phosphorus with this arrangement of its electrons:

The 3-level electrons now rearrange (hybridize) 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.

53
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 NCl 5.

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

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 section, 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 colorless gases are allowed to mix, a thick white smoke of solid
ammonium chloride is formed.

54
Ammonium ions, NH4+, are formed by the transfer of a hydrogen ion from the
hydrogen chloride to the lone pair of electrons on the ammonia molecule.

When the ammonium ion, NH 4+, is formed, the fourth hydrogen is attached by
a dative covalent bond, 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.

55
The H3O+ ion is variously called the hydroxonium ion, the hydronium ion or
the oxonium ion.

In an introductory chemistry course (such as CSEC), 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. BF 3 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.

56
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 aluminum chloride

Aluminum chloride sublimes (turns straight from a solid to a gas) at


about 180°C. 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 aluminum and boron are in the same group of the Periodic Table, as
are fluorine and chlorine.

Measurements of the relative formula mass of aluminum chloride show that


its formula in the vapor 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.

Energy is released when the two co-ordinate bonds are formed, and so the
dimer is more stable than two separate AlCl 3 molecules.

The bonding in hydrated metal ions

57
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 aluminum chloride is covalent, when it dissolves in water, ions are


produced. Six water molecules bond to the aluminum 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 aluminum 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.

Aluminum 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 aluminum re-
organizes (hybridizes) 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 aluminum 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 aluminum and so isn't involved in the bonding. The
resulting ion looks like this:

58
Because of the movement of electrons towards the centre of the ion, the 3+
charge is no longer located entirely on the aluminum, but is now spread over
the whole of the ion.

Two more molecules

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

59
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
delocalization.

The evidence for hydrogen bonding

Many elements form compounds with hydrogen - referred to as "hydrides". If


you plot the boiling points of the hydrides of the Group 4 elements, you find
that the boiling points increase as you go down the group.

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

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

Although for the most part the trend is exactly the same as in group 4 (for
exactly the same reasons), the boiling point of the hydride of the first
element in each group is abnormally high.

In the cases of NH3, H2O and HF there must be some additional intermolecular
forces of attraction, requiring significantly more heat energy to break. These
relatively powerful intermolecular forces are described as hydrogen bonds.

60
The origin of hydrogen bonding

The molecules which have this extra bonding are:

Notice that in each of these molecules:

 The hydrogen is attached directly to one of the most electronegative


elements, causing the hydrogen to acquire a significant amount of
positive charge.
 Each of the elements to which the hydrogen is attached is not only
significantly negative, but also has at least one "active" lone pair.

Lone pairs at the 2-level have the electrons contained in a relatively


small volume of space which therefore has a high density of negative
charge. Lone pairs at higher levels are more diffuse and not so
attractive to positive things.

Consider two water molecules coming close together.

The + hydrogen is so strongly attracted to the lone pair that it is almost as if


you were beginning to form a co-ordinate (dative covalent) bond. It doesn't
go that far, but the attraction is significantly stronger than an ordinary dipole-
dipole interaction.

61
Hydrogen bonds have about a tenth of the strength of an average covalent
bond, and are being constantly broken and reformed in liquid water. If you
liken the covalent bond between the oxygen and hydrogen to a stable
marriage, the hydrogen bond has "just good friends" status. On the same
scale, van der Waals attractions represent mere passing acquaintances!

Water as a "perfect" example of hydrogen bonding

Notice that each water molecule can potentially form four hydrogen bonds
with surrounding water molecules. There are exactly the right numbers of +
hydrogens and lone pairs so that every one of them can be involved in
hydrogen bonding.

This is why the boiling point of water is higher than that of ammonia or
hydrogen fluoride. In the case of ammonia, the amount of hydrogen bonding
is limited by the fact that each nitrogen only has one lone pair. In a group of
ammonia molecules, there aren't enough lone pairs to go around to satisfy all
the hydrogens.

In hydrogen fluoride, the problem is a shortage of hydrogens. In water, there


are exactly the right number of each. Water could be considered as the
"perfect" hydrogen bonded system.

More complex examples of hydrogen bonding

The hydration of negative ions

When an ionic substance dissolves in water, water molecules cluster around


the separated ions. This process is called hydration.

Water frequently attaches to positive ions by co-ordinate (dative covalent)


bonds. It bonds to negative ions using hydrogen bonds.

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

62
However complicated the negative ion, there will always be lone pairs that
the hydrogen atoms from the water molecules can hydrogen bond to.

Hydrogen bonding in alcohols

An alcohol is an organic molecule containing an -O-H group.

Any molecule which has a hydrogen atom attached directly to an oxygen or a


nitrogen is capable of hydrogen bonding. Such molecules will always have
higher boiling points than similarly sized molecules which don't have an -O-H
or an -N-H group. The hydrogen bonding makes the molecules "stickier", and
more heat is necessary to separate them.

Ethanol, CH3CH2-O-H, and methoxymethane, CH3-O-CH3, both have the same


molecular formula, C2H6O.

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

However, ethanol has a hydrogen atom attached directly to an oxygen - and


that oxygen still has exactly the same two lone pairs as in a water molecule.
Hydrogen bonding can occur between ethanol molecules, although not as

63
effectively as in water. The hydrogen bonding is limited by the fact that there
is only one hydrogen in each ethanol molecule with sufficient + charge.

In methoxymethane, the lone pairs on the oxygen are still there, but the
hydrogens aren't sufficiently + for hydrogen bonds to form. Except in some
rather unusual cases, the hydrogen atom has to be attached directly to the
very electronegative element for hydrogen bonding to occur.

The boiling points of ethanol and methoxymethane show the dramatic effect
that the hydrogen bonding has on the stickiness of the ethanol molecules:

ethanol (with hydrogen bonding) 78.5°C


methoxymethane (without hydrogen -
bonding) 24.8°C

The hydrogen bonding in the ethanol has lifted its boiling point about 100°C.

It is important to realize that hydrogen bonding exists in addition to van der


Waals attractions. For example, all the following molecules contain the same
number of electrons, and the first two are much the same length. The higher
boiling point of the butan-1-ol is due to the additional hydrogen bonding.

Comparing the two alcohols (containing -OH groups), both boiling points are
high because of the additional hydrogen bonding due to the hydrogen
attached directly to the oxygen - but they aren't the same.

The boiling point of the 2-methylpropan-1-ol isn't as high as the butan-1-ol


because the branching in the molecule makes the van der Waals attractions
less effective than in the longer butan-1-ol.

Hydrogen bonding in organic molecules containing nitrogen

Hydrogen bonding also occurs in organic molecules containing N-H groups - in


the same sort of way that it occurs in ammonia. Examples range from simple
molecules like CH3NH2 (methylamine) to large molecules like proteins and
DNA.

The two strands of the famous double helix in DNA are held together by
hydrogen bonds between hydrogen atoms attached to nitrogen on one
strand, and lone pairs on another nitrogen or an oxygen on the other one.

64
What is a metallic bond?

Metallic bonding in sodium

Metals tend to have high melting points and boiling points suggesting strong
bonds between the atoms. Even a metal like sodium (melting point 97.8°C)
melts at a considerably higher temperature than the element (neon) which
precedes it in the Periodic Table.

Sodium has the electronic structure 1s 22s22p63s1. When sodium atoms come
together, the electron in the 3s atomic orbital of one sodium atom shares
space with the corresponding electron on a neighboring atom to form a
molecular orbital - in much the same sort of way that a covalent bond is
formed.

The difference, however, is that each sodium atom is being touched by eight
other sodium atoms - and the sharing occurs between the central atom and
the 3s orbitals on all of the eight other atoms. And each of these eight is in
turn being touched by eight sodium atoms, which in turn are touched by
eight atoms - and so on and so on, until you have taken in all the atoms in
that lump of sodium.

All of the 3s orbitals on all of the atoms overlap to give a vast number of
molecular orbitals which extend over the whole piece of metal. There have to
be huge numbers of molecular orbitals, of course, because any orbital can
only hold two electrons.

The electrons can move freely within these molecular orbitals, and so each
electron becomes detached from its parent atom. The electrons are said to be
delocalized. The metal is held together by the strong forces of attraction
between the positive nuclei and the delocalized electrons.

This is sometimes described as "an array of positive ions in a sea of


electrons".

If you are going to use this view, beware! Is a metal made up of atoms or
ions? It is made of atoms.

65
Each positive centre in the diagram represents all the rest of the atom apart
from the outer electron, but that electron hasn't been lost - it may no longer
have an attachment to a particular atom, but it's still there in the structure.
Sodium metal is therefore written as Na - not Na+.

Metallic bonding in magnesium

If you work through the same argument with magnesium, you end up with
stronger bonds and so a higher melting point.

Magnesium has the outer electronic structure 3s 2. Both of these electrons


become delocalized, so the "sea" has twice the electron density as it does in
sodium. The remaining "ions" also have twice the charge (if you are going to
use this particular view of the metal bond) and so there will be more
attraction between "ions" and "sea".

More realistically, each magnesium atom has one more proton in the nucleus
than a sodium atom has, and so not only will there be a greater number of
delocalized electrons, but there will also be a greater attraction for them.

Magnesium atoms have a slightly smaller radius than sodium atoms, and so
the delocalized electrons are closer to the nuclei. Each magnesium atom also
has twelve near neighbors rather than sodium's eight. Both of these factors
increase the strength of the bond still further.

Metallic bonding in transition elements

Transition metals tend to have particularly high melting points and boiling
points. The reason is that they can involve the 3d electrons in the
delocalization as well as the 4s. The more electrons you can involve, the
stronger the attractions tend to be.

The metallic bond in molten metals

In a molten metal, the metallic bond is still present, although the ordered
structure has been broken down. The metallic bond isn't fully broken until the
metal boils. That means that boiling point is actually a better guide to the
strength of the metallic bond than melting point is. On melting, the bond is
loosened, not broken.7

See PPT - Bonding


2.8 – 2.11

The electron pair repulsion theory

The shape of a molecule or ion is governed by the arrangement of the


electron pairs around the central atom. All you need to do is to work out how
many electron pairs there are at the bonding level, and then arrange them to

66
produce the minimum amount of repulsion between them. You have to
include both bonding pairs and lone pairs.

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

 Write down the number of electrons in the outer level of the central
atom. That will be the same as the Periodic Table group number,
except in the case of the noble gases which form compounds, when it
will be 8.
 Add one electron for each bond being formed. (This allows for the
electrons coming from the other atoms.)

 Allow for any ion charge. For example, if the ion has a 1- charge, add
one more electron. For a 1+ charge, deduct an electron.

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

 Divide by 2 to find the total number of electron pairs around the


central atom.
 Work out how many of these are bonding pairs, and how many are lone
pairs. You know how many bonding pairs there are because you know
how many other atoms are joined to the central atom (assuming that
only single bonds are formed).

For example, if you have 4 pairs of electrons but only 3 bonds, there
must be 1 lone pair as well as the 3 bonding pairs.

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

 Arrange these electron pairs in space to minimize repulsions. How this


is done will become clear in the examples which follow.

Two electron pairs around the central atom

The only simple case of this is beryllium chloride, BeCl 2. The electronegativity
difference between beryllium and chlorine isn't enough to allow the formation
of ions.

Beryllium has 2 outer electrons because it is in group 2. It forms bonds to two


chlorines, each of which adds another electron to the outer level of the
beryllium. There is no ionic charge to worry about, so there are 4 electrons
altogether - 2 pairs.

It is forming 2 bonds so there are no lone pairs. The two bonding pairs
arrange themselves at 180° to each other, because that's as far apart as they
can get. The molecule is described as being linear.

67
Three electron pairs around the central atom

The simple cases of this would be BF 3 or BCl3.

Boron is in group 3, so starts off with 3 electrons. It is forming 3 bonds,


adding another 3 electrons. There is no charge, so the total is 6 electrons - in
3 pairs.

Because it is forming 3 bonds there can be no lone pairs. The 3 pairs arrange
themselves as far apart as possible. They all lie in one plane at 120° to each
other. The arrangement is called original planar.

In the diagram, the other electrons on the fluorines have been left out
because they are irrelevant.

Four electron pairs around the central atom

There are lots of examples of this. The simplest is methane, CH4.

Carbon is in group 4, and so has 4 outer electrons. It is forming 4 bonds to


hydrogens, adding another 4 electrons - 8 altogether, in 4 pairs. Because it is
forming 4 bonds, these must all be bonding pairs.

Four electron pairs arrange themselves in space in what is called a


tetrahedral arrangement. A tetrahedron is a regular triangularly-based
pyramid. The carbon atom would be at the centre and the hydrogens at the
four corners. All the bond angles are 109.5°.

68
Other examples with four electron pairs around the central atom

Ammonia, NH3

Nitrogen is in group 5 and so has 5 outer electrons. Each of the 3 hydrogens


is adding another electron to the nitrogen's outer level, making a total of 8
electrons in 4 pairs. Because the nitrogen is only forming 3 bonds, one of the
pairs must be a lone pair. The electron pairs arrange themselves in a
tetrahedral fashion as in methane.

In this case, an additional factor comes into play. Lone pairs are in orbitals
that are shorter and rounder than the orbitals that the bonding pairs occupy.
Because of this, there is more repulsion between a lone pair and a bonding
pair than there is between two bonding pairs.

That forces the bonding pairs together slightly - reducing the bond angle from
109.5° to 107°. It's not much, but the examiners will expect you to know it.

Remember this:

Greatest lone pair - lone


repulsion pair
lone pair - bond
pair
Least bond pair - bond
repulsion pair

Be very careful when you describe the shape of ammonia. Although the
electron pair arrangement is tetrahedral, when you describe the shape, you
only take notice of the atoms. Ammonia is pyramidal - like a pyramid with
the three hydrogens at the base and the nitrogen at the top.

Water, H2O

69
Following the same logic as before, you will find that the oxygen has four
pairs of electrons, two of which are lone pairs. These will again take up a
tetrahedral arrangement. This time the bond angle closes slightly more to
104°, because of the repulsion of the two lone pairs.

The shape isn't described as tetrahedral, because we only "see" the oxygen
and the hydrogens - not the lone pairs. Water is described as bent or V-
shaped.

The ammonium ion, NH4+

The nitrogen has 5 outer electrons, plus another 4 from the four hydrogens -
making a total of 9.

But take care! This is a positive ion. It has a 1+ charge because it has lost 1
electron. That leaves a total of 8 electrons in the outer level of the nitrogen.
There are therefore 4 pairs, all of which are bonding because of the four
hydrogens.

The ammonium ion has exactly the same shape as methane, because it has
exactly the same electronic arrangement. NH4+ is tetrahedral.

Methane and the ammonium ion are said to be is electronic. Two species
(atoms, molecules or ions) are is electronic if they have exactly the same
number and arrangement of electrons (including the distinction between
bonding pairs and lone pairs).

The hydroxonium ion, H3O+

Oxygen is in group 6 - so has 6 outer electrons. Add 1 for each hydrogen,


giving 9. Take one off for the +1 ion, leaving 8. This gives 4 pairs, 3 of which
are bond pairs. The hydroxonium ion is electronic with ammonia, and has an
identical shape - pyramidal.

70
Five electron pairs around the central atom

A simple example: phosphorus(V) fluoride, PF5

(The argument for phosphorus (V) chloride, PCl 5, would be identical.)

Phosphorus (in group 5) contributes 5 electrons, and the five fluorines 5


more, giving 10 electrons in 5 pairs around the central atom. Since the
phosphorus is forming five bonds, there can't be any lone pairs.

The 5 electron pairs take up a shape described as a trigonal bipyramid -


three of the fluorines are in a plane at 120° to each other; the other two are
at right angles to this plane. The trigonal bipyramid therefore has two
different bond angles - 120° and 90°.

A tricky example, ClF3

Chlorine is in group 7 and so has 7 outer electrons. The three fluorines


contribute one electron each, making a total of 10 - in 5 pairs. The chlorine is
forming three bonds - leaving you with 3 bonding pairs and 2 lone pairs,
which will arrange themselves into a trigonal bipyramid.

But don't jump to conclusions. There are actually three different ways in
which you could arrange 3 bonding pairs and 2 lone pairs into a trigonal
bipyramid. The right arrangement will be the one with the minimum amount
of repulsion - and you can't decide that without first drawing all the
possibilities.

71
These are the only possible arrangements. Anything else you might think of is
simply one of these rotated in space.

We need to work out which of these arrangements has the minimum amount
of repulsion between the various electron pairs.

A new rule applies in cases like this:

If you have more than four electron pairs arranged around the
central atom, you can ignore repulsions at angles of greater than
90°.

One of these structures has a fairly obvious large amount of repulsion.

In this diagram, two lone pairs are at 90° to each other, whereas in the other
two cases they are at more than 90°, and so their repulsions can be ignored.
ClF3 certainly won't take up this shape because of the strong lone pair-lone
pair repulsion.

To choose between the other two, you need to count up each sort of
repulsion.

In the next structure, each lone pair is at 90° to 3 bond pairs, and so each
lone pair is responsible for 3 lone pair-bond pair repulsions.

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

72
Now consider the final structure.

Each lone pair is at 90° to 2 bond pairs - the ones above and below the plane.
That makes a total of 4 lone pair-bond pair repulsions - compared with 6 of
these relatively strong repulsions in the last structure. The other fluorine (the
one in the plane) is 120° away, and feels negligible repulsion from the lone
pairs.

The bond to the fluorine in the plane is at 90° to the bonds above and below
the plane, so there are a total of 2 bond pair-bond pair repulsions.

The structure with the minimum amount of repulsion is therefore this last
one, because bond pair-bond pair repulsion is less than lone pair-bond pair
repulsion. ClF3 is described as T-shaped.

Six electron pairs around the central atom

A simple example: SF6

6 electrons in the outer level of the sulphur, plus 1 each from the six
fluorines, make a total of 12 - in 6 pairs. Because the sulphur is forming 6
bonds, these are all bond pairs. They arrange themselves entirely at 90°, in a
shape described as octahedral.

Two slightly more difficult examples

XeF4

73
Xenon forms a range of compounds, mainly with fluorine or oxygen, and this
is a typical one. Xenon has 8 outer electrons, plus 1 from each fluorine -
making 12 altogether, in 6 pairs. There will be 4 bonding pairs (because of
the four fluorines) and 2 lone pairs.

There are two possible structures, but in one of them the lone pairs would be
at 90°. Instead, they go opposite each other. XeF 4 is described as square
planar.

ClF4-

Chlorine is in group 7 and so has 7 outer electrons. Plus the 4 from the four
fluorines. Plus one because it has a 1- charge. That gives a total of 12
electrons in 6 pairs - 4 bond pairs and 2 lone pairs. The shape will be identical
with that of XeF4.

74

Anda mungkin juga menyukai