0% found this document useful (0 votes)
17 views307 pages

GTTNOC Complete Notes 120425

The document discusses fundamental particles, including protons, neutrons, and electrons, and their roles in determining atomic properties. It explains the use of the periodic table to identify elements and their isotopes, as well as the process of time of flight mass spectrometry for analyzing atomic and molecular masses. Additionally, it covers electron configurations, including sub-levels and the unique cases of chromium and copper, along with the concept of ionization energies.

Uploaded by

cidego5155
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
17 views307 pages

GTTNOC Complete Notes 120425

The document discusses fundamental particles, including protons, neutrons, and electrons, and their roles in determining atomic properties. It explains the use of the periodic table to identify elements and their isotopes, as well as the process of time of flight mass spectrometry for analyzing atomic and molecular masses. Additionally, it covers electron configurations, including sub-levels and the unique cases of chromium and copper, along with the concept of ionization energies.

Uploaded by

cidego5155
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 307

3.1.1.

1 Fundamental Particles
Key point: Atoms comprise the sub-atomic particles protons, neutrons and electrons. These determine
both its physical and chemical properties.

How do protons, neutrons and electrons compare in terms of charge and mass?

Proton Neutron Electron


*
Relative charge +1 0 -1
1
Relative mass 1 1
1840

*
when we say relative, we mean ‘compared with’. So, if we set the mass of proton as 1, then mass of a
neutron is near enough the same as, and an electron’s 1/1840th of, that of a proton.

How do we know how many protons and electrons a species contains using the periodic table? The
periodic table organises the 118 elements that have ever existed (only 92 occur ‘naturally’ in the
universe). Information on each element is presented as follows:

The smaller number (e.g. 12) is the atomic number.

• The atomic number tells us how many protons the element has.
• In atoms, the number of electrons equals the number of protons (or atomic number).
• In ions, the no of electrons = the number of protons – the ions charge.
• The number of electrons in O2- = 8 – (–2) = 10

What does the larger number (e.g. 24.3) represent and why is it presented to 1 decimal place? This tell
us the element’s relative atomic mass (Ar). The decimal place is because the atoms of most elements
exist in more than one form called isotopes.

And isotopes are… atoms of an element that have the same number of protons and electrons but
different numbers of neutrons. For example, there are three isotopes of carbon: 12C (carbon-12), 13C
(carbon-13) and 14C (carbon-14).

The number of neutrons = mass number – atomic number

Isotopes have the same chemical properties, because they have the same electron configuration, but
different physical properties e.g. density.

Why is it a ‘relative’ atomic mass? In this instance the relative means ‘when compared to 1/12th the
mass of a carbon-12 atom’.
How do we know all this stuff? This is bound up with the discovery of each of the sub-atomic particles.
The word ‘atom’ comes from the ancient Greek word for ‘uncuttable’.

18th Century – John Dalton imagines atoms to be very small, solid spheres.
1897 – J.J. Thomson discovers that atoms contain a negatively charged particle, which he names the
‘electron’. This leads him to develop the ‘plum pudding model’ of the atom. Thomson knows the atom is
neutral overall and so he imagines it as a ball of positive charge with electrons randomly embedded in it.

1909 – Under Ernest Rutherford’s guidance, Geiger and Marsden perform an experiment where they fire
alpha particles (He2+) at a thin gold foil (just a few atoms thick). If Thomson’s model is correct, the alpha
particles should pass straight through.

However, a small number of the positively charge particles were deflected and even smaller number
(about 1 in 8000) bounced back.

This led Rutherford to realise that the atom was mostly empty space and contained a tiny, dense,
positively charged centre – the nucleus. The atom was now imagined to be like a solar system with
electrons orbiting the nucleus.
1913 – According to the physics of the day, Rutherford’s atom wasn’t possible (the electrons should
eventually spiral into the centre of the atom – they clearly didn’t). Nascent quantum theory had already
begun to show that that energy was quantised i.e. that only certain values of energy are allowed. As a
result, Niels Bohr proposed that the electrons in an atom could only possess specific values of energy,
leading to the idea of energy levels or electron shells. Further experimental evidence backed this up.
1932 – James Chadwick discovered the neutron. It was the last sub-atomic particle to be found chiefly
because it has no charge and so is difficult to detect. We now have the modern nuclear model of the
atom.
3.1.1.2 Time of Flight (TOF) Mass Spectrometry
Key point: Using mass spectroscopy data produced by a time of flight mass spectrometer, we can
identify elements, calculate relative atomic mass and find the relative formula mass of compounds.

What is mass spectrometry? It’s a method for determining the mass of atoms or molecules. It can be
used to find relative isotopic abundance, molecular mass and help determine the structure of a
compound.

How does it a TOF mass spectrometer work? It comprises four stages and any sample being analysed
must firstly be vaporised.

Stage 1: IONISATION
• If the sample is an element or small molecule, it is ionised by firing high energy electrons from an
‘electron gun’ (a hot wire filament). These knock an electron of each particle of the sample
resulting in an ion with a 1+ charge.
• If the sample is a large molecule, electrospray ionisation is used. The sample is dissolved in a
volatile solvent (e.g. water or methanol) and injected through a fine hypodermic needle to give a
fine mist (aerosol). The tip of the needle is attached to the positive terminal of a high-voltage
power supply. The particles are ionised by gaining a proton (i.e. H+) from the solvent.

Stage 2: ACCELERATION
• The positive ions are accelerated using an electric field (negatively charged plate). This results in
ALL the particles, regardless of their mass, having the SAME KINETIC ENERGY (½mv2).

Stage 3: FLIGHT TUBE


• This is what the phrase ‘time of flight’ refers to. Given that all the particles have the same kinetic
energy, the velocity of each particle depends on its mass. This, in turn, affects the time of flight
of each particle.

Stage 4: DETECTION
• The positively charged ions hit a negatively charged electric plate. When they do, they gain an
electron. This results in a flow of charge and an electric current is measured. The size of the
current gives a measure of the number of ions hitting the plate.

What sort of data is recorded then? A computer will plot a graph of ‘relative abundance’ against ‘mass
to charge ratio (m/z)’. The abundance is often given as a percentage but not always. Regarding elements,
this allows us to work their relative atomic mass by working out an average.
The recorded mass spectrum gets more complicated when we consider diatomic molecules such as Br2
and Cl2 This is because we record not only the individual atomic isotopes but also the molecules made up
of these isotopes. Regarding the latter, we will get specific ratios of the m/z values based on the
probabilities of which isotopes make up the molecule.
Chlorine is particularly interesting when it comes to mass spectrometry and has featured in a few past
exam questions, including the organic chemistry side of the course (organic molecules, such as
halogenoalkanes, will feature isotopes in their relative abundance proportions e.g. 75% of CH3CH2Cl will
be CH3CH235Cl and 25% will be CH3CH237Cl.
Thinking again about organic molecules, CH2Cl2 can exist as CH235Cl2, CH237Cl2 and CH235Cl37Cl. So, again,
there will be three peaks, at 84, 86 and 88, with the relative intensities of 9:6:1.

It’s very important that you never assume the abundance given as a % and adds up to 100. The
following example makes this clear.

Sometimes you’re given the Ar of the element along with the mass and abundance of one of the isotopes
and are asked to determine the mass and abundance of any remaining isotopes. The key here it to make
sure you have only one unknown.
In nearly every AS and A Level Paper 1, there has been a question based around the kinetic energy of the
ions accelerated in a TOF mass spectrometer. These calculations revolve around the following two
equations and units really matter.

𝟏 𝒅
𝑲. 𝑬. = 𝟐
𝒎𝒗𝟐 and 𝒗= 𝒕

Quantity Unit
Kinetic Energy (K.E.) Joules (J)
Mass (m) kg (kilograms)
Velocity (v) m/s (metres per second)
Distance (d) m (metres)
Time (t) s (seconds)

You need to be confident in rearranging, and substituting into, the Kinetic Energy equation.

Rearranging Substitution and Rearranging

𝟏 𝟏
𝑲. 𝑬. = 𝒎𝒗𝟐 𝑲. 𝑬. = 𝒎𝒗𝟐
𝟐 𝟐

𝟐𝑲.𝑬.
= 𝒗𝟐 𝟏 𝒅𝟐
𝒎 𝑲. 𝑬. = 𝒎 𝟐
𝟐 𝒕
𝟐𝑲.𝑬.
. =𝒗 𝒎
𝒎 𝒕 = 𝒅.
𝟐 𝑲. 𝑬.

The first step in many questions will often involve you having to work out the mass of a single ion. A
massive clue is if they give you Avogadro’s number 6.022 x 1023 (this is the number of particles in a
mole). If they ever give you a value, you will definitely need to use it.
Here are some example questions:
You need to practice a variety of questions, but this is probably as hard as it gets…

The time of flights for two isotopes of the same element will be very similar. The isotope with the
greater will take a little longer.
3.1.1.3 Electron Configuration (and most of 3.2.1 – Periodicity)
Key points: Our understanding of electron configuration presses from what we learned at GCSE. Now we
learn that, in addition to there being energy levels, there are now sub-levels/sub-shells and orbitals.
Evidence for their existence comes, in part, from experimental data on ionisation energies.

I hear that everything I learned at GCSE about electron configurations is wrong. That’s not strictly true.
Sure, compared with what we’re about to do, it was a crude representation. But it worked as the level
we needed it to.

So, what’s the deal at A level then? Electron shells, or energy levels, are in fact made up of sub-shells (s,
p, d and f) and these, in turn, consist of orbitals.

Does each sub-shell have the same number of orbitals?

Sub-level Maximum number of e- Number of orbitals


s 2 1
p 6 3
d 10 5
f 14 7

Tell me more about the orbitals. These are regions of space in which up to a pair of electrons are found.
As such, they have very specific shapes. It’s only really necessary to know about the orbitals found in in
the s and p sub-shells. These look like this:

The s orbital is spherical in shape and the three p orbitals, which lie on the x, y and z axes respectively,
are described as being dumb-bell in shape.

Okay, I’m struggling to visualise how I would now represent the electron configuration in an atom. I
think the following image, and the rules that follow, should help.

• Electrons enter the lowest energy orbital available (each circle represents an orbital).
• The following diagram helps you to work out the order in which orbitals fill:

• Electrons prefer to occupy orbitals on their own and only pair up when no empty orbitals of the
same energy are available.
• Paired electrons will have opposite spin as this reduces the mutual repulsion between the paired
e-. This shown using up and down arrows

Here’s the example of Nitrogen.

We could also write this as 1s22s22p3. Here are some more examples:

1s22s22p4

1s22s22p63s23p1
The elements in Group 1 of the Periodic Table all have an outer electronic structure of ns1 (where n is a
number between 2 and 7). All Group 2 elements have an outer electronic structure of ns2. Elements in
Groups 1 and 2 are described as s-block elements.

Elements from Group 3 (the boron group) across to the noble gases all have their outer electrons in p
orbitals. These are then described as p-block elements. d-block elements are elements in which the last
electron to be added to the atom is in a d orbital.

So, all I have to do is follow the pattern then? That’s easy! Hold on… not so fast! Chromium and copper
have anomalous electronic structures.

Cr - 1s22s22p63s23p64s13d5
Cu - 1s22s22p63s23p64s13d10

With chromium, the half-filled 4s and 3d sub-shells provide extra stability through the equal distribution
of charge around the atom. With copper, the conclusion must be made that a felled d sub-shell and a
half-filled 4s sub-shell is more stable that the configuration we expected.

Exam questions regarding electron configurations feature chromium and copper more often than not.

Okay, so I think I’ve got the hand of working out the electron configuration of atoms. What about
ions? As regards elements in groups 1 – 7 becoming ions, the principle we learned at GCSE of losing or
gaining electrons to achieve a noble gas structure still applies.

Positive Ions – formed by the loss of e-


Na Atom Na+ ion
1s 2s2 2p6 3s1
2
1s2 2s2 2p6
Negative Ions – formed by the gain of e-
O Atom O2- ion
1s2 2s2 2p4 1s2 2s2 2p6

Species with the same electronic structure are isoelectronic


I take it it’s the same with the transition metals? No, I’m afraid not. With the transition metals it is the
4s electrons that are lost first when ions are formed.

• Ti 1s2 2s2 2p63s2 3p6 3d24s2


• Ti2+ 1s2 2s2 2p63s2 3p6 3d2
• Cr 1s2 2s2 2p63s2 3p6 3d54s1
• Cr3+ 1s2 2s2 2p63s2 3p6 3d3

Okay, okay… I get the picture (I think). How do I know you’re not just feeding me a load of old baloney
about these energy levels, sub-shells and orbitals? Ah, I’m glad you asked. There’s actually some rather
compelling experimental evidence, mainly regarding ionisation energies.

Ionisation what-y? Ionisation energies. The first ionisation energy refers to the energy required to
remove the outermost electron from an atom. The energy needed to do this to one atom is so small that
energies are given for 1 mole of atoms in kJ mol-1. In fact, we need to know the definition of first
ionisation energy.

‘The first ionisation energy is energy required to remove one electron from each atom in a mole of
gaseous atoms producing one moles of gaseous ions with one positive charge.’

X(g) àX+(g) + e- e.g. Na(g) àNa+(g) + e-

Needless to say, there are second ionisation energies, third ionisation energies and so on until all the
electrons surrounding the nucleus have been removed.

2nd ionisation energy X+ (g) à X2+(g) + e- Na+ (g) à Na2+(g) + e-


3rd ionisation energy X2+ (g) à X3+(g) + e- Na2+ (g) à Na3+(g) + e-

The second ionisation energy is energy required to remove one electron from each ion in a mole of 1+
gaseous ions producing one moles of gaseous ions with a 2+ charge.’

Al(g) à Al(g) + 3e- represents the 1st, 2nd and 3rd ionisation energies combined.

I take it the ionisation energies of the various elements aren’t all the same so what affects the value?
There are three factors we need to consider:

_
e
1) distance from nucleus

+ 3) shielding (repulsion) by electrons


in inner shells between nucleus
2) nuclear charge
and outer electron
How can I judge the first factor, the distance from the nucleus of the electron being removed? The
atomic radius of atoms increases as you go down a group in the periodic table. This means the
electrostatic force of attraction between the nucleus and out electrons decreases.

How do I know the atomic radius of atoms increases as you go down a group? Think back to the
electron configuration diagrams you drew at GCSE. As you go down a group in the periodic table, the
number of shells increases… the atoms get bigger.

What about if I go across a period? What happens to atomic radius here? The atoms get smaller. This is
because the number of shells stays the same but the nuclear charge increases. So, the outer shell is
pulled in a little closer. It’s a bit like tightening a belt. Here’s an image showing the pattern.

Okay, so nuclear charge is straight-forward as the more protons there are the greater this is. But
what’s ‘shielding’? Let’s imagine sodium, which has a GCSE electron configuration of 2,8,1. So, between
the outer electron and the nucleus, there are two full shells of electrons. These are obviously of the
same charge as the outer electron(s) and so will repel it/them. We have an attractive force from the
nucleus and a repulsive force from full shells. To put it another way, the outer electron is being shielded
from the full attractive force of the nucleus by the two electron filled energy levels before it. As you go
down a group, the number of full shells between the outer electron and nucleus increases so the
shielding increases.
Right, so can you help me put this all together? Of course. It’s easiest if we consider the trends in
ionisation energy, starting with going down the group.

1st ionisation energy decreases down a group


because:

• Atomic radius increases


• There is greater shielding
• So, there’s a decreasing attraction
between the nucleus and the outermost
electron being removed.

And across the period? Ah, so things get a little interesting here. Let’s start with the general trend in
ionisation energy across the period.

1st ionisation energy increases across the period


because:

• There is an increased in nuclear charge


(i.e. more protons)
• The atomic radius decreases
• So, there’s an increasing attraction
between the nucleus and the outermost
electron being removed.

Great, only the graph doesn’t look as straight-forward as that. There are two points along the period
where I can see general trend is bucked and the 1st ionisation energy actually decreases. Well-spotted!
In the graph above, you can see that there is a decrease when we go from magnesium to aluminium and
also when we go from phosphorus to sulfur. This decrease doesn’t just apply when we go from these
elements – it applies whenever we go from group 2 to 3 and group 5 to 6. This is because we’re always
seeing the same changes in electron configuration at these explain the experimental data.

Let’s start with Group 2 à 3 then. No problem. Compare how the electron configurations of a group 2
and group 3 element end… we’re going from s2 to s2p1. The electron being removed from the Group 3
element is from a p orbital while the one lost from the Group 2 element is from s orbital. The 3p orbital
has a higher energy than the 3s orbital and is, on average, to be found further from the nucleus. This
makes it easier to lose
And when we go from Group 5 à 6? Here, we have to remember our electrons in boxes. When we go
from group 5 to group 6, our electrons in boxes for the final p subshell change as follows:

Group 5 to Group 6

Group 6 is the first occasion in the period that there is a pair of electrons in the same p orbital. The
resulting extra electron-electron repulsions make it easier to lose electron from p4 than p3. So, we could
phrase our answer as follows: the e- removed from paired p orbital, which is easier to remove due to
repulsion between electrons

And these trends are always seen? Yes. Have a look at the images below.
2nd Ionisation energy

Second ionisation energies are much larger than first because you are trying to remove an electron
from an ion that is already positive. In the case of the group 1 metal, second ionisation involves
removing an electron from a lower energy level, which is that much closer to the nucleus so a LOT
more energy is needed to remove this.
Ah, I see what we’re really saying is that the idea of sub-shells and orbitals fit the data rather than the
other way around. What about energy levels. Does experimental data fit with that too? If we consider
successive ionisation energies (1st, 2nd, 3rd etc.), yes. Here’s potassium (1s22s22p63s23p64s1):

After the 1st electron is removed…

• subsequent ionisations remove e- from


a cation (+ve ion)
• the cation gets progressively smaller
• so, we get an increasing force of
attraction between e- and nucleus

When we drop down an energy level to


remove the next electron e.g. when we remove
the second electron from potassium
(1s22s22p63s23p64s1 vs [1s22s22p63s23p6]+, the
electron being removed is much closer to the
nucleus and we get a big jump in ionisation
energy.

Does it tell us anything else? Yes. The point a jump in ionisation energy takes place can tell us what
group the element is in. Have look at the following:

Group 4 (there’s a big jump after the 4th electron Group 6 (there’s a big jump after the 6th electron
is removed) is removed)
GETTING TO THE NUB OF CHEMISTRY
Amount of Substance 3.1.2
3.1.2.1 Relative atomic mass and relative molecular mass

Every measurement we make is relative to something. In chemistry, atomic masses are measured
relative to 12C (carbon-12).

More specifically, the relative atomic mass of an element is the average mass of an atom of the element
relative to one-twelfth the mass of an atom of 12C (carbon-12).

𝑎𝑣𝑒𝑟𝑎𝑔𝑒 𝑚𝑎𝑠𝑠 𝑜𝑓 𝑎𝑛 𝑎𝑡𝑜𝑚 𝑜𝑓 𝑡ℎ𝑒 𝑒𝑙𝑒𝑚𝑒𝑛𝑡


𝑅𝑒𝑙𝑎𝑡𝑖𝑣𝑒 𝑎𝑡𝑜𝑚𝑖𝑐 𝑚𝑎𝑠𝑠 =
1
12 𝑥 𝑡ℎ𝑒 𝑚𝑎𝑠𝑠 𝑜𝑓 𝑐𝑎𝑟𝑏𝑜𝑛 − 12

In chemistry, there are essentially two types of formulae: empirical and molecular.

Empirical formulae show the simplest whole number ratio of particles in a substance. Formulae of ionic
compounds are empirical (the ions are the particles). For example, we can draw a very small 2D portion
of the giant lattice structure of NaCl:

The lattice contains many, many Na+ and Cl- ions and there is no single NaCl (in the 3D structure, every
Na+ ion is surrounded by six Cl- ions and every Cl- ion is surrounded by six Na+ ions). However, the ratio of
Na+ to Cl- is 1:1, which is why its formula is NaCl.

Molecular formulae show the actual number of each atom is present in one molecule (a single particle).

Molecular substance can also have empirical formulae but these are a special case that we’ll deal with
later in these notes.
GETTING TO THE NUB OF CHEMISTRY
Understanding these two types of formulae helps us understand the terms ‘relative formula mass’ (RFM)
and ‘relative molecular mass’ (Mr). How you determine each is the same i.e. you add up all the relative
atomic masses that make up that formula.

The term relative formula mass is used when talking about non-molecular substances, such as ionic
compounds, metals and giant covalent structures. Here are some examples:

• The RFM of NaCl = 23.0 + 35.5 = 58.5


• The RFM of CuSO4 = 63.5 + 32.1 + (4 x 16.0) = 159.6
• The RFM of SiO2 = 28.1 + (2 x 16.0) = 60.1

The term molecular mass is used when talking about molecular substances. For example:

• The Mr of H2O = (2 x 1.0) + 16.0 = 18.0


• The Mr of CH3CH2OH = (2 x 12.0) + (6 x 1.0) + 16.0 = 46.0
• The Mr of C6H12O6 = (6 x 12.0) + (12 x 1.0) + (6 x 16.0) =180.0

3.1.2.2 The mole and the Avogadro constant

Chemists measure the amount of a substance in moles (from the Latin for ‘pile’ or ‘heap’). One mole of a
substance is the RFM/Mr of that substance in grams. This is also called the molar mass and has the units
g mol-1.

1 mole of NaCl has a mass of 58.5g.


1 mole of CH3CH2OH has a mass of 46.0g.
2 moles of H2O has a mass of 32.0g

So, there is a relationship between moles, mass and RFM/Mr.

The Avogadro constant is 6.022 x 1023 and is the number of units of a substance’s formula in one mole.

Substance Molar mass Mass of 1 mole Avogadro constant


(g mol-1) (g)
Fe 55.8 55.8 6.022 x 1023 units of Fe
H2O 18.0 18.0 6.022 x 1023 units of H2O
CaCO3 100.1 100.1 6.022 x 1023 units of CaCO3
H2SO4 98.1 98.1 6.022 x 1023 units of H2SO4
CO2 44.0 44.0 6.022 x 1023 units of CO2
AlCl3 133.5 133.5 6.022 x 1023 units of AlCl3
GETTING TO THE NUB OF CHEMISTRY
In each of the compounds above, the formulae contain more than one particle, which means we can
take things a little further.

• 1 mole of H2O contains: 2 x 6.022 x 1023 H atoms and 6.022 x 1023 O atoms = 1.807 x 1024 atoms.
• 1 mole of CaCO3 contains: 6.022 x 1023 Ca2+ ions and 6.022 x 1023 CO32- ions = 1.204 x 1024 ions.
• 1 mole of AlCl3 contains: 6.022 x 1023 Al3+ ions and 3 x 6.022 x 1023 Cl- ions = 2.409 x 1024 ions.
• 5 moles of AlCl3 contains: 5 x 6.022 x 1023 Al3+ ions and 15 x 6.022 x 1023 Cl- ions = 1.204 x 1025 ions.

The Avogadro constant can also be used to determine the mass of an atom or ion. For example, if we
need to know the mass of an Mg+ ion to answer a Time of Flight mass spectrometer question:

If using the kinetic energy equation, this mass still needs to be converted to kg by dividing by 1000

One of the reasons for using standard form, besides making it easier to use very large numbers such as
the Avogadro constant, is to make it clear how many significant figures a value is to.

When performing multiplication or division calculations, numbers should be kept in a calculator until the
very end (of course, values need to be written down to show your working – these should be to 3 or 4
s.f.). When giving a final answer, this should be to the same number of significant figures as the least
precise value in your calculation.

Q. Calculate the average speed of a car that travels 1557 m in 95.0 seconds.
A. Average speed = 1557/95.0 = 16.4 m s-1 (answer given to 3s.f. as lowest sig figs in data is 3 sig fig for
time)

When performing addition or subtraction calculations, the number of significant figures is irrelevant – it
is about the place value of the data. For example:

Q. Calculate the total energy released when 263 kJ and 1282 kJ are released.
A. Energy released = 263 + 1282 = 1545 kJ (answer is to nearest unit as both values are to nearest unit)

Q. Calculate the total mass of calcium carbonate when 0.154g and 0.01234g are mixed.
A. Mass = 0.154 + 0.01234 = 0.166 g (answer is to nearest 0.001g as least precise number is to nearest
0.001 g)
GETTING TO THE NUB OF CHEMISTRY
3.1.2.3 The ideal gas equation

Before we get to the ideal equation itself, it’s important to know that Avogadro’s Law says that equal
volumes of gases at the same temperature and pressure contain the same number of molecules.

1 mole of any gas at room temperature and pressure (RTP) occupies a volume of 24 dm3 (24,000 cm3).
This is the molar volume.

Questions involving Avogadro’s Law are rare at A level and mostly appear in multiple choice questions
like the one below.

A 30 cm3 sample of nitrogen was reacted with a 60 cm3 sample of fluorine according to the equation:

What is the volume of the gas mixture after the reaction, at constant temperature and pressure?

So, the volume of gas after the reaction is 10 + 40 cm3 = 50 cm3


GETTING TO THE NUB OF CHEMISTRY
The ideal gas equation is:

Quite often, you’re given pressure in kPa (1 kPa = 1,000 Pa) and volume in cm3 (1m3 = 1,000,000cm3 or
1cm3 = 1 x 10-6 m3).

The common errors students make when answering questions involving the ideal gas equations are:

1. Rearranging the equation incorrectly


2. Forgetting convert units or doing so incorrectly.

When answering questions involving pV = nRT, there’s almost always 1 mark for simply rearranging the
ideal gas equation.
𝒑𝑽 = 𝒏𝑹𝑻

𝒏𝑹𝑻 𝒑𝑽 𝒏𝑹𝑻
𝑽 = = 𝒏 𝒑 =
𝒑 𝑹𝑻 𝑽

Q. How many moles of gas occupy 19400 cm3 at 27.0oC and 101 kPa?
GETTING TO THE NUB OF CHEMISTRY
Q. Find the volume of ethyne (C2H2) that can be prepared from 10.0 g of calcium carbide at 20.0oC and
100 kPa (3sf).
GETTING TO THE NUB OF CHEMISTRY
3.1.2.4 Empirical and molecular formula

To remind ourselves of empirical and molecular formulae:

• Empirical formulae show the simplest whole number ratio of particles in a substance.
• Molecular formulae show the actual number of each atom is present in one molecule (a single
particle).

Molecular substances can also have an empirical formula. Empirical formulae are worked out using
either percentage or mass compositions.

Q. A compound is found to contain, by mass, iron 72.4% and oxygen 27.6%.

Q. A compound contains 40.0 g of carbon, 6.7 g of hydrogen and 53.5 g of oxygen. It has a relative
molecular formula of 60. Find both the empirical and the molecular formula of the compound.
GETTING TO THE NUB OF CHEMISTRY
3.1.2.5 Balanced equations and associated calculations

Ionic equations are usually written for reactions that occur in solution. They provide a much clearer
picture of how the ions present are reacting.

While reaction equations show all the substances present in solution, ionic equations show only those
ions that are changed in some way during the reaction.

State symbols are VERY important.

In the two examples above, SO42-, Na+ and Cl- are examples of spectator ions. These take no part in the
actual chemical reaction and so are ignored in an ionic equation.

Precipitation reactions can also be represented using ionic equations. For example:
GETTING TO THE NUB OF CHEMISTRY
It’s in everyone’s interests that the chemical industry converts as much of the raw materials they use
into useful products. Atom economy is the calculation used to determine the efficiency of a reaction.

𝑡𝑜𝑡𝑎𝑙 𝑚𝑎𝑠𝑠 𝑜𝑓 𝑡ℎ𝑒 𝑑𝑒𝑠𝑖𝑟𝑒𝑑 𝑝𝑟𝑜𝑑𝑢𝑐𝑡


𝑃𝑒𝑟𝑐𝑒𝑛𝑡𝑎𝑔𝑒 𝑎𝑡𝑜𝑚 𝑒𝑐𝑜𝑛𝑜𝑚𝑦 = 𝑥 100%
𝑡𝑜𝑡𝑎𝑙 𝑚𝑎𝑠𝑠 𝑜𝑓 𝑡ℎ𝑒 𝑝𝑟𝑜𝑑𝑢𝑐𝑡𝑠

Let’s consider the atom economy of the fermentation of glucose to produce ethanol.

Reacting Mass Calculations

Q. What mass of carbon dioxide is produced when 5.60 g of butene is burnt?

Q. 4.92 g of hydrated magnesium sulphate crystals (MgSO4.nH2O) gave 2.40 g of anhydrous magnesium
sulphate on heating to constant mass. Work out the formula mass of the hydrated magnesium sulphate
and so the value of n.
GETTING TO THE NUB OF CHEMISTRY
Limiting Reagents

Reactants are very rarely added in their exact molar ratios. Nearly always, there is too much of one and
too little of another. When there is too large a quantity of a reagent, we say it is in excess. When there’s
too little, we describe it as the limiting reagent or reactant. It’s the amount of limiting reagent that
determines how much product is made.

Q. What’s the greatest amount of NH3 that can be made with 3.2 moles of N2 and 5.4 moles of H2?
Which is the limiting reactant? Which is in excess and how many moles are left over?
GETTING TO THE NUB OF CHEMISTRY
Q. Hydrazine (N2H4) was used as the rocket fuel for the Apollo missions to the moon. It is by reaction of
ammonia with sodium chlorate. What mass of hydrazine is made by reaction of 100 g of ammonia with
100 g of sodium chlorate (I)?

When you make a new substance by a chemical reaction, you may not get all the expected amount of
product. For example, if you reacted 4 g of hydrogen with 32 g of oxygen, you may get less than 36 g of
water.

Reasons include:

• the reaction may be reversible (both the forwards and backwards reaction can take place)
• some of the product may be lost when it is separated from the reaction mixture
• some of the reactants may react in other reactions.

As a result, we talk about % yield (The more stages there are in producing the intended product, the
lower yield).

Q. Nitrogen reacts with hydrogen to make ammonia. Calculate the maximum theoretical mass of
ammonia that can be made by reacting 90.0 g of hydrogen with an excess of nitrogen. In the reaction,
only 153 g of ammonia was produced. Calculate the percentage yield.
GETTING TO THE NUB OF CHEMISTRY
3.1.2.5 Balanced equations and associated calculations (solutions)

1 dm = 10 cm
1 dm = 10 cm x 10 cm x 10 cm = 1000cm3
3

To work out the concentration of a solution in g dm-3…

Q. 25.0 cm3 of 0.020 mol dm-3 sulfuric acid neutralises 18.6 cm3 of barium hydroxide solution. Find the
concentration of the barium hydroxide solution in mol dm-3 and g dm-3.

.
GETTING TO THE NUB OF CHEMISTRY
Titrations
Many solution calculations in A level chemistry centre around titrations. A major feature of these is the
making of a standard solution i.e. fixed volumes of solutions of accurately known concentrations.
Titrations are then performed against a portion of this volumetric solution e.g. 25.0 cm3 from a 250cm3
volumetric flask.

Making a volumetric solution…


GETTING TO THE NUB OF CHEMISTRY
Performing the titration…
GETTING TO THE NUB OF CHEMISTRY
Titrations are repeated until 2 concordant values are obtained (within 0.10 cm3 of each other).
GETTING TO THE NUB OF CHEMISTRY
There can be several errors in a titration’s practical method that can affect the accuracy of the results.

1. The final rinsing of the pipette should be using the solution that will be pipetted into the conical
flask and not water. Rinsing with water will dilute the added solution, decreasing its
concentration.

2. The funnel must be removed from the burette before starting the titration. This is because
droplets could drip down during a titration making the measured titre smaller than it actually is.

3. The presence of an air bubble in the jet will result less being added from the burette than is
recorded using the final and initial volumes – the burette will record the air bubble’s volume as
having been added but it wasn’t – it was just air let out). You should ensure there are no bubbles
before recording your initial burette reading.

4. The inside of the conical flask should be rinsed with water during a titration. This improves the
accuracy by ensuring ALL the reactants are mixed.

Adding water doesn’t affect the total number of moles in the flask.

Experimental data is never exact. There is always an element of uncertainty in the data due to the
precision of the equipment. This usually expressed as a percentage error. Uncertainty is an estimate
attached to a measurement that gives the range of values within which the true value lies. For
example, the measuring cylinder below will measure a volume with an uncertainty of ±0.5cm3.

Percentage uncertainty can be calculated using the expression:

The lower the percentage error, the more accurate the piece of equipment is.
GETTING TO THE NUB OF CHEMISTRY
Be aware that some experimental results involve two readings on a piece of equipment.

Situations involving one readings Situations involving two readings


Using a measuring cylinder to measure Using burette to measure the volume added
volume. (final reading – initial reading)
Using a volumetric flask to measure a Using a mass balance to measure mass adde
specific volume (final reading – initial reading)
Using a pipette to measure a specific Using a thermometer to measure a
volume temperature change (final reading – initial
reading).
Using a thermometer to record a
temperature

If 3.5g is measured using a mass balance with an uncertainty of 0.005g


a. What is the percentage error in the measurement?
b. If 0.15g is measured using the same balance, what is the percentage error?
c. Comment on the difference in % uncertainty calculated. Why is this the case?

How might the above look when it comes to titrations.

How could we reduce the uncertainty in this titration? In other words, how can we make the titre
bigger?

• Increase the number of moles in the conical flask by either pipetting a greater volume or
dissolving a greater mass when making the standard solution.
• Decrease the concentration of the solution in the burette so a greater volume must be added.
GETTING TO THE NUB OF CHEMISTRY
Sometimes, a titration can be used to determine the metal in a carbonate or hydogencarbonate.
GETTING TO THE NUB OF CHEMISTRY
Q. A 1.575 g sample of ethanedioic acid crystals, H2C2O4.nH2O, was dissolved in water and made up to
250 cm3. One mole of the acid reacts with two moles of NaOH. In a titration, 25.0 cm3 of this solution of
acid reacted with exactly 15.6 cm3 of 0.160 mol dm-3 NaOH. Calculate the value of n.
GETTING TO THE NUB OF CHEMISTRY
A back titration is done to analyse a base (or acid) that does not react easily or quickly with an acid (or
base). Instead, the base (or acid) is treated with an excess of acid (or base), and then the left over acid
(or base) titrated. You can then work back to find out about the original base (or acid).
GETTING TO THE NUB OF CHEMISTRY
Here’s a harder example.
GETTING TO THE NUB OF CHEMISTRY
3.1.3.1-3.1.3.4 Structure and Bonding
Key point: There are three varieties of bonding (covalent, ionic and metallic) and four types of crystal
structure involving this bonding (ionic, giant covalent, metallic and molecular).

Before we look at the three forms of bonding in detail, do they have anything in common? Yes, they do
– they’re all a consequence of electrostatic forces of attraction. However, they differ in how this
attraction comes about.

Let’s start with ionic bonding then. Ions are elements, or compounds, that have more, or fewer,
electrons than protons. For example, Na+ has 11 protons and 10 electrons; SO42- has 48 protons and 50
electrons. When you have positively and negatively charged ions present, there is a strong electrostatic
attraction between these oppositely charged ions. This attraction is in all directions and the ions come
together to form a 3D giant lattice* structure (they pack as closely together as they can).

* a lattice is a regular, repeating pattern – something the above has.

Hmm, the above seems to remind me of something I remember doing years ago. I’m guessing you’re
thinking of the solid particle diagram you did in year 7.

Really, every solid form of a substance has a lattice structure. This includes molecular substance when in
their solid stated e.g. I2

What about metallic bonding then? You may have realised from your studies that metals are generally
very good at losing electrons. In metals, the outer electron(s) of each metal atom become delocalised i.e.
they no longer have a fixed position. As each metal atom has ‘lost’ this electron, or electrons, they’re
now positively charged ions (cations). This results in a lattice structure of cations throughout which are
the delocalised electrons. As a consequence, there are strong electrostatic attractions between the two.
GETTING TO THE NUB OF CHEMISTRY
And covalent bonding… is when adjacent atoms share a pair, or pairs, of electrons. The electrostatic
attraction is between the shared electrons and the nuclei of the atoms either side of them. Covalent
bonds are represented in displayed formulae with an ‘–‘ e.g H–H.

And dative / coordinate covalent bonding? Typically, each of the two electrons in a covalent bond came
from different atoms. For example, the covalent bond in H–H is the result of each hydrogen atom’s single
electron being shared. Dative / coordinate covalent bonding is a term that explains a particular covalent
bond came about when both shared electrons came from just one of the bonded atoms.

You said ‘pair, or pairs, of electrons’. Could you explain this further please? There can be single, double
and triple bonds between atoms e.g. C–C, C=C and CºC. As the electrostatic attraction between the
nuclei and the electrons is increasing (because there are more electrons), bond lengths get shorter and
bond enthalpies (the energy required to break them) increase.

Bond Bond length (x 10-10 m) Bond enthalpy (kJmol-1)


C–C 1.54 346
C=C 1.34 602
CºC 1.20 835
GETTING TO THE NUB OF CHEMISTRY
Bond lengths are also affected by the size of the atoms involved. The bigger atoms, the larger the bond
length and the weaker the bond. This is because a greater distance between the nuclei and shared
electrons leads to a weaker electrostatic attraction.

Bond Bond length (x 10-10 m) Bond enthalpy (kJmol-1)


C–F 1.37 485
C–Cl 1.76 339
C–Br 1.94 285
C–I 2.14 213

Other than their strength and length, is there anything different about double and triple bonds? Well,
single bonds are known as s (sigma) bonds. These are formed by head-on overlapping of atomic orbitals:

The ‘second’ bond in a double bond is formed by sideways or lateral overlapping of atomic orbitals. This
is called a π (pi) bond.

So, a double bond such as C=C in C2H4 is made up of a s and π bond.

In organic chemistry, it’s important to remember that the π bond restricts rotation and so geometric
isomerism, a type of stereoisomerism (when molecules have the same structural formula but a different
spatial arrangement of atoms), can arise. This is signified using the letters E and Z at the beginning of the
molecules name e.g. E-but-2-ene.
GETTING TO THE NUB OF CHEMISTRY
You don’t think you spent too long on covalent bonds? Considering their importance in organic
chemistry, no. Oh, okay. The title of this document is structure and bonding. Explain, please. Easy peasy
(!?!), the physical properties of a substance (e.g. its melting and boiling points, conductivity, solubility
etc.) can be explained by considering the structure of, and bonding present in, a substance.
GETTING TO THE NUB OF CHEMISTRY
Melting and boiling points are common exam question themes aren’t they? Can I have some context
please? Okay, let’s compare the energy required to overcome the forces of attraction to change state in
NaCl, diamond and H2O.

• NaCl – Its lattice dissociation energy is 787 kJmol-1


• Diamond – there are four moles of C-C bonds for every mole of C(diamond) = 4 x 346 kJmol-1 = 1384
kJ per mole of diamond
• H2O – In ice, there are two moles of Hydrogen bonds (an intermolecular force of attraction) for
every mole of H2O = 2 x 23.3 kJmol-1 = 46.6 kJ per mole of ice.

This makes sense as diamond has a very high melting point, NaCl a high melting point and H2O the lowest
of the three

What about if we compare substances with the same structure and bonding? Okay, let’s compare two
ionic compounds first NaF and MgF2.

Mg2+ has a greater charge and smaller radius than Na+ so we would expect a higher melting point for
MgCl2 compared with NaCl, which is indeed the case (1536 K vs 1266 K). Similarly, we would expect NaF
to have a higher melting point than NaCl because F- is a smaller ion than Cl- (melting points: NaF (1226
K); NaCl (1074 K).

And with metals? One mole of Na has one mole of Na+ ions and one mole of delocalised electrons (e-).
One mole of magnesium has one mole of Mg2+ ions and two moles of delocalised electrons. So, we would
expect a stronger electrostatic force of attraction in magnesium compared with sodium and therefore a
higher melting point. This is indeed the case (melting points: magnesium (923 K); sodium (371 K).

I think I’m getting the hang of this. So, diamond has a higher melting point than graphite because we
need to break FOUR covalent bonds for every carbon atom compared with THREE in graphite. You got
it kid! Woo hoo!

So, can I have some example questions now please?

1. Use your knowledge of structure and bonding to explain why sodium bromide has a melting point
that is higher than that of sodium, and higher than that of sodium iodide.

Sodium has a giant lattice structure consisting of Na+ ions surrounded by a sea of delocalised electrons. In
this metallic bonding, there is a strong electrostatic attraction between the metal ions and delocalised
electrons. NaBr and NaI are both ionic compounds in which there is a strong electrostatic attraction
between oppositely charged ions. In both cases, this is stronger than the attraction found in sodium
GETTING TO THE NUB OF CHEMISTRY
metal, hence their higher melting points. The difference between NaBr and NaI can be explained by the
size of the Br- and I- ions. The former is smaller, therefore it has a greater charge density and there is a
stronger attraction with the Na+ ions.

2. Name the type of bond formed between N and Al in H3NAlCl3 and explain how this bond is formed.

It is a dative / co-ordinate covalent bond in which both electrons have come from the nitrogen atom.

3. Explain why strontium has a higher melting point than barium.

The metal cations in strontium have a smaller ionic radius than those in barium (1s2 2s2 2p6 3s2 3p6 4s2
3d10 4p6]2+ vs 1s2 2s2 2p6 3s2 3p6 4s2 3d10 4p6 5s2 4d10 5p6]2+ ). Both ions have a 2+ charge and this charge is
spread over a smaller volume in strontium than barium i.e. the strontium ions have a greater charge
density (charge/volume). This results in a greater electrostatic attraction between the Sr2+ ions and
delocalised electrons so more energy is required to overcome them.

Interesting fact… You can’t actually melt diamond under standard conditions. If you heat diamond to
very high temperatures at atmospheric pressure it becomes graphite and, if you continue heating this,
the graphite will sublimate.
GETTING TO THE NUB OF CHEMISTRY
A Brief Interlude – Beyond the Octet
When electrons are shared, bonded atoms often obtain a noble gas configuration in which they now
have 8 electrons in their outer shell. This is called the Octet Rule and is what you learned at GCSE.
However, this isn’t always the case and you need to be prepared for this.

The Incomplete Octet

While most elements below atomic number 20 follow the octet rule, several exceptions exist, including
compounds of the Group 3 elements boron and aluminium. These two elements readily form
compounds that have six valence electrons, rather than the usual eight predicted.

While molecules exist that contain atoms with fewer than eight valence electrons, these compounds are
often reactive and can react to form species with eight valence electrons. For example, BF3 will readily
bind a fluoride anion to form the BF4- anion, in which boron follows the octet rule. Another example is
that NH3 will form a dative / coordinate bond with AlCl3, after which aluminium follows the octet rule.

BF3 AlCl3

The Expanded Octet

Many group elements in the third period and below form compounds that deviate from the octet rule by
having more than 8 valence electrons.
The octet rule can be ‘expanded’ by some elements by utilizing the d- orbitals found in the third principal
energy level and beyond. Sulfur, phosphorus, silicon, and chlorine are common examples of elements
that form an expanded octet.
Phosphorus pentachloride (PCl5) and sulfur hexafluoride (SF6) are examples of molecules that deviate
from the octet rule by having more than 8 electrons around the central atom.

PCl5 SF6
GETTING TO THE NUB OF CHEMISTRY
3.1.3.5 Shapes of simple molecules and ions
Key point: Simple molecules and polyatomic ions have 3D shapes and we can work these out using
VSEPR (Valence Shell Electron Pair Repulsion).
Can you explain what VSEPR is please? VSEPR works on the principle that the electron pairs around an
atom repel each other (they are both negatively charged after all). They will therefore arrange
themselves to be as far apart as possible. Depending on how many pairs of electrons there are, and
whether they’re bonding or lone pairs, we can work out the shape of a molecule.
GETTING TO THE NUB OF CHEMISTRY
Do all electron pairs repel each other equally? Not quite. Lone pairs will repel more than bonding pairs.
This is because the lone pair’s negative charge is more concentrated around the atom compared to the
bonding pair. The difference between lone and bond pairs does not alter the basic shape of a molecule
BUT it does distort it and so alter bond angles (i.e. the angle between two adjacent bond pairs). In
addition, a double bond is essentially treated in VSEPR as if it’s normal bonding pair.

Okay, can you tell me how to work out a simple molecule’s shape please? We start by working out
how many electron domains there are and then how many of these are bonding and lone pairs. This is
how it’s done whenever a molecule contains single bonds only:

("#. #& #'()* )+),(*#"- #" (.) ,)"(*/+ /(#0 ) 2 ("# .#& #(.)* /(#0-)
Step 1: 3
= 𝑛𝑜. 𝑜𝑓 𝑝𝑎𝑖𝑟𝑠 𝑜𝑓 𝑒𝑙𝑒𝑐𝑡𝑟𝑜𝑛𝑠

Step 2: no. of atoms around central atom = no. of bonding pairs

Step 3: Total no. of pairs - bonding pairs from step 2 = no. of lone pairs

Okay. Please show me how it works to explain the shape of ClF3 please?

Step 1: Chlorine has 7 atoms on its outer shell and there are 3 fluorine atoms attached to it. Therefore,
𝟕"𝟑
the no. of pairs of electrons = = 5 𝑝𝑎𝑖𝑟𝑠 𝑜𝑓 𝑒𝑙𝑒𝑐𝑡𝑟𝑜𝑛𝑠
$

Step 2: There are 3 fluorine atoms attached to the chlorine atom so there are 3 bonding pairs.

Step 3: 5 electron pairs – 3 bonding pairs = 2 lone pairs

So, ClF3 has 3 bonding pairs and 2 lone pairs surrounding the central chlorine atom.

Is there any difference if the molecule is an ion, such as PBr4+? Yes. If the ion is positively charged, you
subtract an electron in step 1 for every positive charge. If the ion is negatively charged, you add an
electron in step 1 for every negative charge. Steps 2 and 3 are exactly as before. Here are two examples:

PBr4+

Step 1: Phosphorous has 5 electrons on its outer shell and there are 4 bromine atoms attached to it. The
%"𝟒'𝟏
ion has a 1+ charge. So, the no. of pairs of electrons = = 4 𝑝𝑎𝑖𝑟𝑠 𝑜𝑓 𝑒𝑙𝑒𝑐𝑡𝑟𝑜𝑛𝑠
$

Step 2: There are 4 bromine atoms attached to the phosphorus atom so there are 4 bonding pairs.

Step 3: 4 electron pairs – 4 bonding pairs = 0 lone pairs


GETTING TO THE NUB OF CHEMISTRY
So, PBr4+ has 4 bonding pairs and 0 lone pairs surrounding the phosphorus atom. So, the molecule is
tetrahedral.

TlBr32-

Step 1: Thallium has 3 electrons on its outer shell and there are 3 bromine atoms attached to it. The ion
𝟑"𝟑"𝟐
has a 2- charge. So, the no. of pairs of electrons = $
= 4 𝑝𝑎𝑖𝑟𝑠 𝑜𝑓 𝑒𝑙𝑒𝑐𝑡𝑟𝑜𝑛𝑠

Step 2: There are 3 bromine atoms attached to the thallium atom so there are 3 bonding pairs.

Step 3: 4 electron pairs – 3 bonding pairs = 1 lone pair

So, TlBr32- has 3 bonding pairs and 1 lone pair surrounding the central thallium atom. So, the molecule is
trigonal pyramidal.

Can I have some example questions please?

1. There are two lone pairs of electrons on the oxygen atom in a molecule of oxygen difluoride (OF2).
Explain how the lone pairs of electrons on the oxygen atom influence the bond angle in oxygen
difluoride.

Lone pairs repel more than bond pairs so the molecule’s bond angle will be 104.5°, lower than the
regular tetrahedral angle

2. Explain why CF4 has a bond angle of 109.5°.

Around carbon there are 4 bonding pairs of electrons and no lone pairs. These repel equally and spread
as far apart as possible

3. Explain how the electron pair repulsion theory can be used to deduce the shape of, and the bond
angle in, PF3

Phosphorus has 5 electrons in its outer shell and 3 fluorine atoms attached. So, PF3 has 3 bonding pairs
and 1 lone pair. Electron pair repulsion theory tells us that electron pairs repel as far as possible and lone
pairs repel more than bonding pairs. Therefore, PF3 has atrigonal pyramidal shape
with an angle of to 107°.
To get a better idea of the actual shapes, go to https://phet.colorado.edu/en/simulation/molecule-
shapes
GETTING TO THE NUB OF CHEMISTRY
3.1.3.6 Electronegativity and Bond Polarity
Definition: the power of an atom to attract the pair of electrons in a covalent bond.

The strength of this 'power' is reflected by a value on the Pauling scale from 0.7 (caesium) to 4.0
(fluorine). The higher the number, the greater the power.

What factors affect this power? The answer is… the same factors that affect ionisation energy: nuclear
charge, atomic radius and shielding.

Why is that? Well, both ionisation energy and electronegativity are a measure of the strength of
electrostatic attraction between an atom’s nucleus and its outer electrons.

Nuclear charge – more protons means a stronger attraction between nucleus and bonding pair of electrons.
Atomic radius – being closer to the nucleus means a stronger attraction between nucleus and bonding pair of
electrons.
Shielding – fewer shells of electrons between the nucleus and the pair of electrons means there's less shielding (less
repulsion) and a stronger attraction between nucleus and bonding pair of electrons.

So, there must be a trend on the periodic table? Of course, and it's the same as the general trend for
ionisation energy, except we ignore the noble gases. Electronegativity increases going up a group and
across a period.

What’s the point of the electronegativity scale? It allows us to understand better the bonding in a
compound.

How? At A level, we begin to understand that ionic and covalent bonding isn't simply the either/or we
may have understood it to be at GCSE. Instead, we have a spectrum of bonding, with 100% covalent at
one end and 100% (perfect) ionic at the other. In between, there is polar covalent and polar ionic. The
size of the difference in two elements’ electronegativities gives us an idea of where on the spectrum the
bonding in a substance lies.
GETTING TO THE NUB OF CHEMISTRY
What? If two elements have differing electronegativity values, then their ability to attract a pair of
electrons is different. The element with the greater electronegativity is more powerful. You can think of
it as a sort of tug-of-war. If element X is only a little more powerful than Y, then X will pull the electrons a
little toward it but not enough to be able to take them. This results in partial charges on each atom, d+
and d-. This means one atom is slightly positive and the other slightly negative. So, we get a polar
covalent substance. If element X is a lot more powerful than Y, then X will, in effect, possess both
electrons and we now have a X- ion and a Y+ ion.

I think it would be useful to see a table of actual values, don’t you? Okay, here you go…

Now I need some concrete examples. You never have to remember the electronegativity values of
elements, nor do you need to remember the table below. But it might be helpful in understanding the
area.

Type of bond Electronegativity difference Examples


Covalent (non-polar) < 0.5 H-H, Br-Br, C-H
Polar covalent Between 0.4 and 1.6 C-Cl, H-F, O-H, Al-Cl*
Ionic > 1.6 NaCl, CaF2, MgO

*at GCSE, a useful rule of thumb for deciding whether a compound had ionic or covalent bonding was whether it consisted of a metal and non-
metal or two non-metals. This still generally works as it mainly reflects differences in electronegativity. However, there are exceptions such as
AlCl3 having covalent bonding. As it happens, aluminium is at the heart of many exceptions in A level chemistry.

Can I have an example question, please? d– d+


A bond between nitrogen and hydrogen can be represented as: N–H
(i) In this representation, what is the meaning of the symbol d+?
It means that the atom is partially positive (or electron deficient)

(ii) From this bond representation, what can be deduced about the electronegativity of hydrogen relative
to that of nitrogen?
Nitrogen is more electronegative than hydrogen.
GETTING TO THE NUB OF CHEMISTRY
3.1.3.7 Forces between molecules
Key point: The melting and boiling points of molecular substances are influenced by the strength of the
forces between molecules.

There are three types of forces between molecules: van der Waals forces, permanent dipole-dipole
forces and hydrogen bonding

Tell me about van der Waals (vdW) forces. Also known London dispersion forces, every molecular
substance, including noble gases, has these between every particle. They exist because the random
movement of electrons in these substances can lead to temporary dipoles i.e. regions of these
substances can become temporarily d+ and d- (these symbols represent partial charges). While these
exist (they are continually appearing and disappearing), they can induce (cause) d+ and d- regions in
adjacent molecules leading to weak electrostatic attraction between them.

Does anything affect the strength of vdW forces? Okay, so the above says that the resulting
electrostatic forces are weak, and they are. But if you have a LOT of them, then they can be strong. It
hopefully isn't a surprise to learn that the more electrons there are, the more induced dipoles there can
be, and therefore, the stronger the attractions overall. So, the bigger the Mr of the molecule, the
stronger these forces. This helps explain why Iodine (I2) is a solid at room temperature.

Does anything else affect the strength? Yes, the contact surface area of the molecule. This only really
comes into play if we’re comparing the melting or boiling points of two molecules with the same, or
similar, Mr. The more branched a molecule is, the smaller its contact surface area.
GETTING TO THE NUB OF CHEMISTRY
That’s van der Waals forces dealt with. What about permanent dipole-dipole forces? When we
looked at electronegativity, we saw that if the difference in electronegativity between two bonded
elements was between 0.4 and 1.8, their bonding was polar covalent. This means the resulting bond is
always d+ and d- i.e. it has a permanent dipole. Also, this dipole will be more significant than the
temporary ones responsible for vdW forces.

So, are you saying you just need to determine whether a substance has one or more polar bonds to
know if it possesses permanent dipole-dipole forces or not? Hold on, not so fast. This where things get
a little tricky. Ultimately, we need to determine if a molecule has a net (overall) dipole or not. Here’s a
quick, imperfect checklist:

1. If there's only one polar bond, then the molecule must have a net dipole and will, therefore,
possesses permanent dipole-dipole forces. E.g. HBr, CH3CH2Cl, CH3CH2CHO

2. If there’s more than polar bond AND there’s alone pair on the centrally bonding atom, there’s a
net dipole.

3. A molecule has not net dipole and is therefore non-polar if ALL the bonds form the central
atom are the same polar bond AND there are no lone pairs on the central atom.

Pound for pound (really, Mr for Mr), permanent dipole-dipole forces are stronger than vdW forces.
GETTING TO THE NUB OF CHEMISTRY
And hydrogen bonding? This is a particular type of permanent dipole-dipole force that has its own
name. For hydrogen bonds to exist, a hydrogen atom must be bonded to Nitrogen, Oxygen or Fluorine.
This results in the hydrogen atom having a significant partial positive charge. This then forms an
electrostatic attraction with a lone pair of electrons on the nitrogen, oxygen or fluorine of an adjacent
molecule.

Hydrogen bond strengths range from 4 kJ to 50 kJ per mole of hydrogen bonds (by comparison, the
covalent bonds H-H and O-H are 432 and 459 kJmol-1 respectively. So, hydrogen bonding is roughly a
tenth as strong as a covalent bond.

Mr for Mr, hydrogen bonding > permanent dipole-dipole forces > van der Waals forces
Hydrogen bonds are pretty special then? You bet they are. For a start, they explain some of the
‘strange’ properties of water, such as its melting point of 273K, which is remarkably high for a molecule
with an Mr of only 18.

Tell me more… Well, H2O is very unusual in that its solid form, ice, is less dense than its liquid form,
water. The reason for this is that as the temperature of H2O gets close to its freezing point, the
individual molecules begin to line up so that the maximum number of hydrogen bonds can be made (as
a liquid, they're continually breaking and reforming because of the motion of the molecules). In the
final solid structure, the H2O molecules in this giant lattice are, on average, actually further apart than
when liquid.
GETTING TO THE NUB OF CHEMISTRY
Anything else? Yes, they're an essential feature of DNA too. The two strands of DNA’s double helix are
held together by hydrogen bonds between the complementary bases: guanine with cytosine and
adenine with thymine.

Is it possible to draw everything together? I’m struggling in particular to get my head around van der
Waals forces and the fact that they’re temporary and yet always there! No problem. To help us
understand the latter a little better, I’m going to use an analogy with disco lights.
Type Key thoughts Relevant to… Disco lights analogy
van der Waals Temporary / instantaneous dipoles are Every molecule! The lights are constantly
formed due to constantly moving Alkanes, alkenes and flashing and hence on
electrons and these, in turn, induce aromatic compounds with temporarily. However, they
temporary dipoles in adjacent only alkyl substituents flash at different times so the
molecules. feature these forces ONLY. dance floor is always
Greater Mr à greater number of illuminated, albeit dimly.
electrons à stronger vdW forces. A greater number of flashing
Straight-chain > branched due to there lights (a bigger Mr) means the
being a greater contact surface area. floor is more illuminated.
Permanent dipole-dipole These arise when atoms with different • Acyl chlorides Some of the lights are
electronegativities are covalently • Aldehydes and ketones permanently switched on
bonded resulting in a polar bond AND • Halogenoalkanes while the others continue to
there is a net dipole. The shape of the • 3o amines flash. As a result, the dance
molecule is typically very important • Polyesters floor is more brightly lit.
when determining if there’s the latter.
Hydrogen-bonding These arise when hydrogen is bonded to • Water and ammonia As with PDD, some of the
fluorine, oxygen or nitrogen (the three • Alcohols lights are permanently
most electronegative elements). This is • Ammonia switched on. However,
really a type of permanent dipole-dipole • 1o and 2o amines brighter bulbs are being used
interaction but its greater strength • Amides and so the dance floor is even
results in it having its own classification. • Carboxylic acids more brightly lit.
• Hydrogen fluoride
• Polyamides and DNA

Can I have some example questions please?

1. Silicon tetrafluoride (SiF4) is a tetrahedral molecule. Deduce the type of intermolecular forces in
SiF4. Explain how this type of intermolecular force arises and why no other type of intermolecular
force exists in a sample of SiF4.

Van der Waals forces are present SiF4. The random movement of electrons means any individual
molecule of SiF4 will have an uneven distribution of electron density. This induces a temporary dipole in
a neighbouring molecule resulting in a weak electrostatic attraction between the two. Even though Si-F
is a polar bond, SiF4 is a symmetrical molecule, and the four polar bonds cancel each other out – there
is no net dipole and therefore no permanent dipole-dipole forces. There is no hydrogen bonding as
there are no hydrogens bonded to F, N or O.

2. Explain, with reference to intermolecular forces, why distillation allows propanal to be separated
from the other organic compounds in this reaction mixture.

Both propan-1-ol and propanoic acid possess hydrogen bonding between their molecules. Propanal has
permanent dipole-dipole forces between its molecules. The forces between the molecules in aldehyde
are weaker (than those in alcohol and acid so it will evaporate first.)
GETTING TO THE NUB OF CHEMISTRY
3. Use your knowledge of structure and bonding to explain why the melting point of iodine is low
(113.5 °C) and why that of hydrogen iodide is very low (–50.8 °C).

Both I2 and HI are molecular substances, and so their melting points are related to the strength of the
forces between their molecules. Neither has a high melting point, so these forces are relatively weak. I2
possesses only van der Waals forces whereas HI contains permanent dipole-dipole attractions in
addition to van der Waals forces. However, their respective melting points tells us that the van der
Waals forces in I2 are stronger than the forces already stated in HI. This is because I2 has almost twice as
many electrons as HI (Mr of I2 = 253.8; Mr of HI = 127.9)

Hmm, that last question was particularly interesting… I agree. What ultimately reveals the strength of
the forces between particles are the melting and boiling points of a substance. Our knowledge of these
forces hopefully allows us to explain these relative strengths. Van der Waals forces may be the
weakest, Mr for Mr, but they can still be very strong if there’s a lot of them. Just ask geckos.

Geckos? What have they got to do with all this? They’re the lizards that can ‘stick to walls’. Well, it
turns out they can do this because of van der Waals forces.

No way! Yes, way! The soles of Tokay geckos (Gekko gecko) contain half a million hairs, and the end of
each hair splits into between 100 and 1,000 tiny spatulas, visible only with an electron microscope. The
ultra-close contact of the split ends with a wall or floor creates billions of weak molecular attractions
(van der Waals forces). Fascinatingly, geckos are incapable of sticking to Teflon because the latter
doesn’t support van der Waals forces. I suppose a fried egg could have told us this if we’d asked.
Source: https://www.nature.com/news/2002/020826/full/news020826-2.html
GETTING TO THE NUB OF CHEMISTRY
3.1.4.1 Enthalpy Change
I take it you’re going to start these notes by explaining the word ‘enthalpy’? Errm, yes… of course I was
although I don’t think you’re going to like it. Try me. Okay, here goes. Enthalpy, represented by the
symbol H, is the heat content of a substance. You can’t measure this directly, but you can measure its
change and it’s this that we’re interested in. We represent the change in enthalpy with DH. In science,
the Greek letter delta represents change. As a capital (D), it represents a sizeable change. In lower-case
form (d), it represents a very small change. So, we use DH to signify a change in enthalpy (usually in
kJmol-1)

Tell me more about this change in enthalpy or DH. It’s actually something you have considered before
when speaking about exothermic and endothermic reactions. Here, you were describing whether DH
was a transfer of energy to or from the surroundings.

Oh yeah, I remember the reaction profiles we drew to illustrate this. That’s right. I think things make
more sense if we think of reactants and products being chemical stores of energy.

• If the reactants have a greater chemical store than the products then the chemical store
decreases during the reaction and the difference is transferred to the surroundings (remember,
energy cannot be created or destroyed… only transferred from one store to another).
• If the products have a greater chemical store than the reactants, then the chemical store
increases during the reaction and the difference must have been taken in from the surroundings.

However, we still use the term enthalpy and so that’s what we’ll continue to use.

Exothermic Endothermic

Thanks for that. I seem to recall deciding whether a reaction was exothermic or endothermic based on
the sign (+ or -) when I did bond energies at GCSE. That’s right. Typically, whenever you spoke about
energy transfer at GCSE, you were speaking qualitatively. However, when you did bond energies (we’ll be
looking at these short but will be referring to them as bond enthalpies), you considered energy change
quantitatively. At A level, we do much more of the latter. However, we need to consider some
definitions of particular enthalpy changes first.
GETTING TO THE NUB OF CHEMISTRY
Groan, I hate definitions. B...b…but they’re important. Are they though? Yes, of course. You can’t
possibly get specific examples correct if you don’t know the definition. Does it help if we think of them as
classifications? Dunno, maybe. Shall we get this over and done with though? Hmm, I think we will do
better with a more positive attitude but here goes.

In order to make a fair comparison between our definitions of enthalpy changes, all take place under
standard conditions. These are:

• A temperature of 298K (25oC)


• A pressure of 1 atmosphere (» 100 kPa)
• Substances are in their standard state (e.g. O2 is a gas and Na is a solid)
• Solutions have a concentration of 1.0 mol dm-3

We show this using a superscript q e.g. DHq (delta H standard)

Right, got it. Great. So, here are the definitions we need to know for the AS part of the course.

Enthalpy change of formation (DHqf): The change in enthalpy when 1 mole of a substance is formed from
its elements in their standard states. Hopefully, you won’t be surprised to hear that state symbols are
essential.

E.g. DHqf (Na2O):


2Na(s) + ½O2(g) à Na2O(s)

E.g. DHqf (CH3CH2OH):


2C(s) + 3H2(g) + ½O2(g) à CH3CH2OH(l)

By definition, The standard enthalpy of formation of an element in its standard state is zero.

Enthalpy change of combustion (DHqc): The change in enthalpy when 1 mole of a substance is
completely burnt in oxygen under standard conditions.

E.g. DHqc (Na):


Na(s) + ¼O2(g) à ½Na2O(s)

E.g. DHqc (CH3CH2OH):


CH3CH2OH(l) + 3O2(g) à 2CO2(g) + 3H2O(l)

Finally, many questions will ask you to calculate DHqr . This is simply the enthalpy change of a given
reaction when equation quantities of materials react under standard conditions, and with everything in
its standard state. This isn’t a definition to learn.

E.g Zn(s) + CuSO4(aq) à ZnSO4(aq) + Cu(s)


GETTING TO THE NUB OF CHEMISTRY
3.1.4.2 Calorimetry
You mentioned earlier that at A level we often look at ‘enthalpy changes’ quantitatively. How? Before I
answer that, I need to show you two equations we’re going to be using. Here’s the first:

And here’s the second:

So, we’re going to be taking measurements and then plugging these values into the two equations to
work out the DHq? Absolutely. Let’s consider how we could determine DHqc. We’re going to do this by
combusting a known mass of whatever the substance is (at A level, this is usually an alcohol) and
measuring the change in temperature of a known volume of water. The experiment would look
something like this.
GETTING TO THE NUB OF CHEMISTRY
We want to ensure that as much of the energy as possible that’s released in the combustion reaction is
used to heat the water. The wind shield is there to reduce draughts and help ensure the flame is acting
directly on the can. This is made of copper to ensure energy is transferred to the water efficiently.

You may remember from GCSE Physics that water has a density of 1g cm-3. This means that 100cm3 of
water has a mass of 100g. This is the m in the equation Q = mcDT. We also know the specific heat
capacity (the c of the equation) of water. This is 4.18 J g-1 K-1 (they sometimes round it up to 4.2). You will
always be given this.

So, you measure the change in temperature (final temp. – start temp) and multiply this by the mass
(volume) of water and 4.18. This gives you the value of the energy transferred (Q) in the experiment in
Joules (J).

You said a known mass of whatever is being combusted. How is this worked out? Okay, so you’re not
saying in advance of doing the experiment that you’ll combust, say, 0.500g of the fuel. In fact, you
determine the mass retrospectively.

How? You measure the mass of the burner containing the fuel before you start the practical. After
lighting the wick, you heat the water so that there is a sizeable change in the temperature (e.g. a
minimum of 10oC) and then snuff out the wick. Measure the mass of the burner again and the difference
between this and the initial measurement is the mass of fuel combusted. You can now work out the
moles of fuel combusted.

𝑚𝑎𝑠𝑠
𝑚𝑜𝑙𝑒𝑠 =
𝑀𝑟

And if I know the energy transferred (Q) and the moles of fuel combusted, I can work out DHqc That’s
right. Just remember that the unit of Q is J and the unit of DHqc is kJmol-1 so divide Q by 1000 first. And
regarding choosing between a ‘+’ and ‘–‘ to put before our final calculated value… if the temperature
being measured increased, then it was an exothermic reaction and we put ‘–‘… if it decreased then it was
endothermic and we put a ‘+’. Combustion reactions are clearly exothermic!
GETTING TO THE NUB OF CHEMISTRY
I’ve seen a lot of exam questions ask why our calculation is less than that found in a data book. Why is
that? Data book values are accurate. So, ours mustn’t be. Why could that be? Let’s consider again the
definition of DHqc : The change in enthalpy when 1 mole of a substance is completely burnt in oxygen
under standard conditions. In all likelihood, there was a degree of incomplete combustion and this has a
significant impact on the energy released.

Could you maybe show me an example calculation? Sure.

Am I only going to be asked about DHqc? No, the specification says:

Students could be asked to find ΔH for a reaction by calorimetry. Examples of reactions could include:
• dissolution (dissolving) of potassium chloride
• dissolution of sodium carbonate
• neutralising NaOH with HCl
• displacement reaction between CuSO4 + Zn
• combustion of alcohols.

You mean I’ve got to learn four more ways of working out DHq? No. The calculations you do are exactly
the same as you did with DHqc. The experimental procedure and analysis of results is slightly different
though and you need to know this because it’s CORE PRACTICAL 2.

Woah!!! I’m all ears. Tell me, all I need to know? For a start, you’re going to use a polystyrene cup
instead of a copper can. Let’s consider one of the examples in the spec: the displacement reaction
between CuSO4 + Zn. Firstly, saying it’s a displacement reaction is a misnomer. It’s a redox reaction:

Cu2+(aq) + Zn(s) à Cu(s) + Zn2+(aq)


GETTING TO THE NUB OF CHEMISTRY
A polystyrene cup… now you’re talking! Indeed. Here’s how things set up followed by the method:

1. Measure the temperature of the solution(s) every minute e.g. CuSO4(aq).


2. At the 4th minute, add the other reactant e.g. Zn(s)
3. Again, record the temperature every minute up to 15 minutes (you haven’t measured the
temperature at 4th minute – you restart recording it at the 5th).

Is that it? Yes. But we haven’t finished, we have to now consider how we use these results so that we
have the most accurate information we can get to put into our earlier equations.

4. Plot a temperature vs time graph (we always describe graphs as y vs x) and extrapolate to the 4th
minute the two lines of best fit.
5. Measure the difference between the two lines at the 4th minute – this is DT (the change in
temperature).

If it’s an exothermic reaction If it’s an endothermic reaction

Okay, I can see I’ve now got a value for DT that I can input into q = mcDT but what about ‘m’ and ‘c’.
Think back to the apparatus diagram. What are you actually measuring the temperature of? The
solution. Right, and what is the solution mostly? Erm, water I think. That’s right. We make the
assumption that we’re measuring the temperature of water. So, m is still 4.18 J g-1 K-1 and the mass is still
the total volume in grams (remember, the density of water is 1 g cm-3). This is very important – the mass
is always the volume in grams.

Is there anything I need to watch out for? This is a classic for working out which reactant is in excess and
therefore which is the limiting reactant. You need to then use the moles of the latter in your calculations.
Also, when questions ask why you value is less than the data book value, it’s because of heat loss to the
surroundings i.e. not all the energy released stayed in the polystyrene cup. If they didn’t use a lid, doing
so would make your calculation more accurate.
GETTING TO THE NUB OF CHEMISTRY
3.1.4.3 Applications of Hess’s Law
What is Hess’s Law? The enthalpy change accompanying a chemical change is independent of the route
by which the chemical change occurs.

Eh? Let’s used an analogy to help us. Imagine I know the change in altitude going to I went from Archway
to Camden and I also know the change in altitude is I went from Archway to Highgate. But what I want to
know is the change in altitude from Camden to Highgate (don’t ask). Using the information I already
have, I could work this out without leaving the doghouse I live in.

∆Alt (Cam to High) = - DAlt (Arch to Cam) + DAlt (Arch to High)

Hess’s law is just this… only about chemical reactions. Some enthalpy changes can’t be measured directly
in the lab but Hess’s law allows us to calculate them anyway.

What sort of information do I get so that I can used Hess’s law? Let’s quote the AQA specification:

Students should be able to use Hess’s law to perform calculations, including calculation of enthalpy
changes for reactions from enthalpies of combustion or from enthalpies of formation.

The information you get could be: DH for a reaction, DHc and/or DHf

Our understanding of the definitions of the latter is therefore very important and this is mainly to do
with the arrows and their direction.

In enthalpies of formation, the arrows go towards the thing being formed.

E.g. DHqf (CH3CH2OH):


2C(s) + 3H2(g) + ½O2(g) à CH3CH2OH(l)

In enthalpies of combustion, the arrows go away from what is being combusted.

E.g. DHqc (CH3CH2OH):


CH3CH2OH(l) + 3O2(g) à 2CO2(g) + 3H2O(l)
GETTING TO THE NUB OF CHEMISTRY
I think seeing an example would help. I agree. Here’s one where the information given is the enthalpies
of formation/

Our understanding of the definition of DHqf is very important.

DHqf CH4(g) C(s) + 2H2(g) à CH4(g)


DHqf CO2(g) C(s) + O2(g) à CO2(g)
DHqf H2O(l) H2(g) + ½O2(g) à H2O(l)

We can now construct our Hess Cycle using this information.

And enthalpy of combustion? Here you go


GETTING TO THE NUB OF CHEMISTRY
Any others to be aware of? Sometimes we’re given other data.

Why do calculations like this? Why not do just measure the enthalpy change? Because it’s not
possible? What do you mean? It’s not possible to measure directly measure the enthalpy change for the
conversion of the blue hydrated copper (II) sulfate crystals into the white anhydrous crystals. This is for a
variety of reasons:

• The dehydration reaction cannot be controlled


• It’s very difficult to measure the temperature rise of a solid
• CuSO4.5H2O would need heating (so the temperature change cannot be measured)
• It’s impossible to add the exact amount of water (to obtain the value by reverse process)
• You cannot mix the solid with water to obtain perfect crystals

Similarly, it’s not possible to measure DHf (CO) i.e. C(s) + ½O2(g) à CO(g) because it would be impossible to
prevent the C reacting with O2 to form some CO2
GETTING TO THE NUB OF CHEMISTRY
3.1.4.4 Bond Enthalpies
This title rings a bell. Yep. At GCSE, you looked at bond energies when you did the Energy Transfer topic.
What’s the difference? Not a lot. The questions might sometimes be a little more complicated than
you’ve previously been used to but essentially there’s not much more to it.

Okay, so would you mind refreshing my memory – what is a bond energy or enthalpy? So, they’re
going to be referred to as bond enthalpies at A level and you’ll often be given information on mean bond
enthalpies. The mean bond enthalpy is…

‘…the average energy needed to break one mole of the bond to give separated atoms - everything being
in the gas state.’

Here are some example values.

Bond C-C C-H C=C C-O C=O O-H


Mean bond enthalpy / kJ mol-1 348 412 612 360 743 463

The higher the mean bond enthalpy, the stronger the bond.

Why ‘mean’? The strength of a bond is affected by the environment it’s in. The energy required to break
a C-H bond in, say, CH4 will be slightly different to that for the same type of bond in CH3F. So, the mean
bond enthalpy is…

‘the energy required to break 1 mole of a specific bond averaged across a range of compounds’.

Not all bond enthalpies you’re given are an average, though? Eh? If you’re given the bond enthalpy for,
say, H-H or Cl-Cl then these can’t be means. Why not? Because these bonds only exist in one molecule
each! The H-H bond only exists in H2 and the Cl-Cl bond only in Cl2

What can we use them for? They provide us with another way of calculating the enthalpy change in a
reaction. At GCSE, you learned that…

The change in energy = ∑ 𝑏𝑜𝑛𝑑𝑠 𝑏𝑟𝑜𝑘𝑒𝑛 − ∑ 𝑏𝑜𝑛𝑑𝑠 𝑚𝑎𝑑𝑒

It’s exactly the same at A level. We can understand why the above is the case if we approach bond
enthalpies as a Hess diagram.
GETTING TO THE NUB OF CHEMISTRY
Can you show me an example? Here’s a straight-forward one.

Example 1 Calculate the enthalpy change for the following reaction given the following bond enthalpies.
CH3-CH3(g) + Cl2(g) → CH3-CH2Cl(g) + HCl(g)

Bond enthalpies: C-C 348, C-H 412, Cl-Cl 242, C-Cl 338, H-Cl 431 kJ/mol

………………………………………………………………………………………………………………………………………………

………………………………………………………………………………………………………………………………………………

………………………………………………………………………………………………………………………………………………

………………………………………………………………………………………………………………………………………………

………………………………………………………………………………………………………………………………………………

………………………………………………………………………………………………………………………………………………

Example 2 Hydrazine has the formula N2H4 and is used as a rocket fuel (e.g. for the Apollo moon rockets). It burns
following reaction for which the enthalpy change is -583 kJ/mol.
N2H4(g) + O2(g) → N2(g) + 2 H2O(g)

Calculate the N-N bond enthalpy in hydrazine given the following bond enthalpies.
Looks
Bondstraight-forward.
enthalpies: N-HJust388,
watch out,498,
O=O they can
N≡N mix things
944, O-Hup!463 kJ/mol

………………………………………………………………………………………………………………………………………………

………………………………………………………………………………………………………………………………………………

………………………………………………………………………………………………………………………………………………

………………………………………………………………………………………………………………………………………………

………………………………………………………………………………………………………………………………………………

………………………………………………………………………………………………………………………………………………

………………………………………………………………………………………………………………………………………………

………………………………………………………………………………………………………………………………………………

Example 3 Ethanol has the formula C2H5OH and is used as a fuel (e.g. for cars in Brazil). It burns in the following rea
for which the enthalpy change is -1015 kJ/mol.
C2H5OH(l) + 3 O2(g) → 2 CO2(g) + 3 H2O(g)

Calculate the C-C bond enthalpy in ethanol given the following bond enthalpies and enthalpy of vaporisat
ethanol.

Bond enthalpies: C-H 412, O-H 463, C-O 360, C=O 743, O=O 498 kJ/mol
Enthalpy of vaporisation of ethanol, C2H5OH(l) = 44 kJ/mol

………………………………………………………………………………………………………………………………………………
GETTING TO THE NUB OF CHEMISTRY
3.1.5.1 Collision theory
Didn’t I do this at GCSE? Yep! So why am I doing it again? Because it’s important you need to make sure
you completely understand it.

If a reaction is to take place, the particles involved must collide. When they do, bonds may be broken,
and new ones formed. Collision is necessary but it isn’t sufficient. Not only must the right particles
collide, but they must also do so with enough energy to cause any relevant bonds to be broken. (They
must also collide in an orientation that favours the reaction, but this isn’t covered by our A level).

The minimum energy for a reaction to occur between colliding particles is called the activation energy,
EA. Activation energy is closely related to the rates of reaction. The greater the activation energy, the
slower the rate of reaction.

You covered EA (activation energy) in your GCSE and identified it when drawing reaction profiles
diagrams for exothermic and endothermic reactions. These illustrate the role of activation energy as an
energy barrier that must be overcome by reactants before they may form products.

Exothermic Endothermic

The fact that a certain minimum energy is needed to initiate most reactions is well-illustrated by fuels
and explosives. Once these reactions have started, they supply the energy needed to keep them going.
GETTING TO THE NUB OF CHEMISTRY
3.1.5.2 Maxwell-Boltzmann distribution
Using the kinetic theory of gases and probability theory, James Clerk Maxwell and Ludwig Boltzmann
derived equations for the distribution of kinetic energies amongst the molecules of a gas at a particular
temperature. The Maxwell-Boltzmann distribution looks like this:

This shows us that:


• Only small numbers of molecules have either extremely high or extremely low energies.
• The curve is not symmetrical.

The graph shows us further things too:


GETTING TO THE NUB OF CHEMISTRY
3.1.5.3 Effect on temperature on reaction rate
If you raise the temperature, a reaction takes place at by 10K, calculations show that you increase the
rate of collisions by approximately 6%. However, experiments have shown that, for many reactions, a
10K temperature increase roughly doubles the reaction rate. How the shape of the Maxwell-Boltzmann
distribution curve changes with temperature helps us understand the reason for this.

Note: The area under the curve for all three cases is the same because the total number of molecules is
the same. However, we can see that the proportion of molecules with E ≥ EA is much greater when the
system is at a higher temperature. A change in temperature has a much larger effect than a change in
concentration.

Sometimes you’re given a curve and asked to sketch a second that represents the same system as a
higher or lower temperature.

• If you have to sketch the curve at a lower temperature, its peak should be higher and to the left
of the first. Your line should only cross the original’s once and must finish below it.
• If you have to sketch the curve at a higher temperature, its peak should be lower and to the right
of the first. Your line should only cross the original’s once and must finish above it.
GETTING TO THE NUB OF CHEMISTRY
3.1.5.4 Effect of concentration and pressure
Increasing the concentration of the reactants increases the rate of a reaction. At greater concentration,
there are more particles in a given volume.

Distances between particles are reduced and there is an increase in the frequency of collisions, which
typically results in an increase in the rate of the reaction (I say ‘typically’ because this doesn’t always
happen – we’ll discover why when we look at 3.1.9 Rate Equations).

For reactions involving gases, increasing the pressure increases the concentration, and so there’s an
increase in the frequency of collisions and, again, typically, an increase in the rate of reaction.

Concentration inevitably changes during the course of a reaction.


GETTING TO THE NUB OF CHEMISTRY

3.1.5.5 Catalysts
A catalyst is a substance that increases the rate of a chemical reaction without being changed in chemical
composition or amount. Consequently, they never appear in the overall reaction equation. They work by
providing an alternative reaction route of lower activation energy.
GETTING TO THE NUB OF CHEMISTRY
3.1.6.1 Chemical Equilibria and Le Chatelier’s Principle
Equilibria lies at the heart of many areas of chemistry. At GCSE, you explored the topic in a qualitative
sense centred around Le Chatelier’s principle. This is also how we begin the topic at A level. Before we
do, it’s important to make sure we understand a few key ideas and terms.

1. Many chemical reactions are reversible. At GCSE showed this by using ⇌ instead of à in a
reaction equation.
reactants ⇌ products

An example of this, which you learned at GCSE, is:

CuSO4.5H2O(s) ⇌ CuSO4(s) + 5H2O(l)

2. For A level and beyond, the symbol ⇌ will tell us that the reaction in question is an equilibria
reaction.
3. Equilibrium is used to denote balance. Two people balanced on a see-saw is an example of static
equilibrium. Running on a treadmill, where the combination of the person is running forwards
and the treadmill moving backwards results in position of the runner not changing, is an example
of dynamic equilibrium.

Static Equilibrium Dynamic Equilibrium

In chemistry, equilibrium is dynamic. This means the forwards and backwards reactions are taking place
at the same time such that the amount of each substance stays the same and it appears as if nothing is
changing. This can only be achieved in a closed system i.e. one in which there is no loss or gain of
materials to or from the surroundings.
GETTING TO THE NUB OF CHEMISTRY
Dynamic equilibrium in chemistry is not like balancing a see-saw. There need not be equal concentration
of reactants and products in the equilibrium mixture. At equilibrium, there could be a much greater
concentration of reactants than products, a similar concentration, or a greater concentration of products
compared with reactants. As a consequence, we refer to the position of equilibrium.

There are four key features of dynamic equilibria.

1. Equilibrium can be attained from either direction.


2. Equilibrium can only be achieved in a closed system.
3. The rate of the forward reaction = the rate of the reverse reaction
4. The position of equilibrium can lie anywhere between the 100% reactants and 100% products.
GETTING TO THE NUB OF CHEMISTRY
We can affect the position of equilibrium by changing temperature, concentration or pressure (gases
only).

Le Chatelier’s principle says:

If a system at equilibrium is subjected to a small change, the equilibrium tends to shift so as to minimise
the effect of the change.

Or, in other words,

The position of equilibrium shifts to try to cancel out any changes you introduce.

In simple terms, this means:

Change Action by Equilibrium


Make it hotter Cool it back down
Make it cooler Heat it back up
Add more chemical Use it up to get rid of it
Remove some chemical Make more of it
Increase the pressure Reduce the pressure
Decrease the pressure Increase pressure

Let’s look at these changes more closely.

A+B⇌C+D ΔH = +ve (this means the forward reaction is endothermic)


Increase Decrease
If you increase the temperature of the If you decrease the temperature of the
system, then you’re increasing the system, then you’re decreasing the
energy in the system. The position of energy in the system. The position of
equilibrium will shift to try to cancel this equilibrium will shift to try to cancel this
Temperature
increase in energy i.e. shift in the decrease in energy i.e. shift in the
endothermic direction. So, in the exothermic direction. So, in the reaction
reaction above, the position of above, the position of equilibrium will
equilibrium will shift to the right. shift to the left.
If you increase the concentration of one If you decrease the concentration of one
of the substances, the position of the substances, the position
equilibrium will shift in an attempt to equilibrium will shift in an attempt to
Concentration cancel this increase i.e. to the other side. cancel this decrease i.e. to the same side.
For example, if you increase the For example, if you decrease the
concentration of A, the position of concentration of A, the position of
equilibrium will shift to the right. equilibrium will shift to the left.
If you increase the pressure, the position If you decrease the pressure, the position
of equilibrium will shift to the side with of equilibrium will shift to the side with
Pressure
the fewest molecules (this counteracts the most molecules (this counteracts the
the change by reducing pressure). change by increasing pressure).
GETTING TO THE NUB OF CHEMISTRY
Sulfur trioxide is made by the following reaction. The forward reaction is exothermic.

2SO2(g) + O2(g) ⇌ 2SO3(g)

If the temperature of this equilibrium was increased, what would happen to the equilibrium yield of
sulfur trioxide? Explain your reasoning.

2 SO2(g) + O2(g) ⇌ 2 SO3(g)

An increase in temperature favours the endothermic reaction, which is the backwards one in this
reaction. So, the position of equilibrium will shift to the left resulting in a decrease in the yield of SO3.

If the pressure of this equilibrium was increased, what would happen to the equilibrium yield of sulfur
trioxide? Explain your reasoning

2SO2(g) + O2(g) ⇌ 2SO3(g)

Increasing the pressure favours the side with the fewest moles in order to lower the pressure. So, the
position of equilibrium will shift to the right and there will be an increase in the yield of SO3.

Catalysts do not change the position of equilibrium because the rates of both the forward and
backwards reactions are sped up. So, while the position isn’t changed, the time it takes to reach
equilibrium is decreased.

Many industrial reactions are reversible. Conditions have to be chosen that will produce as much of the
desired product as possible in the shortest time while not letting costs be too high.

Two significant examples are the Haber Process, which makes ammonia, and the Contact Process, which
produces sulfuric acid. We’ll look closely at the Haber Process.

N2(g) + 3H2(g) ⇌ 2NH3(g) ΔH = -ve

Thinking about Le Chatelier’s principle, to increase the yield of ammonia, we would increase the pressure
and decrease the temperature. However, a lower temperature means a slower rate of reactions and, in
business, time is money. Of course, using a catalyst helps. In the Haber Process, this is iron.
GETTING TO THE NUB OF CHEMISTRY
When everything is considered (yield, time, cost and safety), the following compromise conditions are
used.

Temperature: 450oC
Pressure: 200 atm
Catalyst: Iron

A lower temperature gives a greater yield of NH3 because the forward reaction is exothermic. However, a
higher temperature gives a faster rate. So, 450oC is a compromise between yield and rate.

A higher pressure gives greater yield of NH3 because there are fewer gas molecules on the right-hand
side of the equation. However, high pressure is very expensive because especially strong pipes and a
great deal of energy are needed. So, 200 atm is a compromise between yield and cost.

An iron catalyst is used because it increases the rate of reaction. However, it has no effect on the
position of equilibrium or yield.
GETTING TO THE NUB OF CHEMISTRY
3.1.6.2 Equilibrium constant Kc for homogeneous systems
In a homogenous system, all reactants and products are in the same phase e.g. they’re all aqueous or
gaseous with no boundary separating them. Countless experiments have shown that there is a
relationship between the concentrations of reactants and products in equilibrium mixtures.

For the equilibrium mixture:

mA + nB ⇋ yC + zD

It is found that, with respect to the forward reaction:

[𝑪]𝒚 [𝑫]𝒛
𝑲𝒄 =
[𝑨]𝒎 [𝑩]𝒏

Square brackets round a species mean concentration (mol dm-3) of the species and here we’re
considering the concentrations at equilibrium.

Kc is the equilibrium constant (the ‘c’ refers to ‘concentrations’) and is constant as a given temperature.
The whole formula is called the equilibrium expression. The value of the equilibrium constant is not
affected either by changes in concentration or addition of a catalyst. It is only affected by temperature.

Reactants always feature on the denominator of the expression and products on the top. So, if we look
at the reaction from the other direction:

yC + zD ⇋ mA + nB

[𝑨]𝒎 [𝑩]𝒏
𝑲&𝒄 =
[𝑪]𝒚 [𝑫]𝒛

Such that:

𝟏
𝑲&𝒄 =
𝑲𝒄

Example Question

2HI ⇋ H2 + I2

At 827oC, concentrations at equilibrium for HI, H2 and I2 are 0.05, 0.01 and 0.01 respectively.

Calculate Kc

[𝐻( ][𝐼( ] (0.01)(0.01)


𝐾' = (
= = 0.04 (𝑢𝑛𝑖𝑡𝑠? )
[𝐻𝐼] (0.05)(
GETTING TO THE NUB OF CHEMISTRY
The units of Kc depend on the expression. For example, Kc has no units when the total moles on both
sides are equal. This is because the concentration units cancel out. Consider the equilibrium:

CH3OH + HCOOH ⇋ HCOOCH3 + H2O

[𝐇𝐂𝐎𝐎𝐂𝐇𝟑 ][𝐇𝟐 𝐎] [mol dm−3 ][mol dm−3 ] [mol dm−3 ][mol dm−3 ]
𝑲𝒄 = = [mol dm−3 ][mol dm−3 ] = [mol dm−3 ][mol dm−3 ] = no units
[𝐂𝐇𝟑 𝐎𝐇][𝐇𝐂𝐎𝐎𝐇]

Let’s consider another:

CO2(g) + 2H2(g) ⇋ CH3OH(g)

[𝑪𝑯𝟑 𝑶𝑯(𝒈)]
𝑲𝒄 = [𝑪𝑶𝟐 (𝒈)] ×[𝑯𝟐 (𝒈)]𝟐

𝑚𝑜𝑙 𝑑𝑚#$ 𝑚𝑜𝑙 𝑑𝑚#$ 1


= = = = 𝒎𝒐𝒍#𝟐 𝒅𝒎𝟔
#$ #$
𝑚𝑜𝑙 𝑑𝑚 (𝑚𝑜𝑙 𝑑𝑚 ) % (𝑚𝑜𝑙 𝑑𝑚 ) × (𝑚𝑜𝑙 𝑑𝑚#$ ) × (𝑚𝑜𝑙 𝑑𝑚#$ ) (𝑚𝑜𝑙 𝑑𝑚#$ ) × (𝑚𝑜𝑙 𝑑𝑚#$ )
#$

The size of the equilibrium constant can tell us about the composition of the equilibrium mixture.

Kc < 1 Kc > 1
Reactants predominate over products in the equilibrium Product predominate over reactants in the equilibrium
mixture. Equilibrium position lies towards to the reactants. mixture. Equilibrium position lies towards to the products.

Kc < 1x10 -10 Kc = 1 Kc > 1x10 10


Reaction regarded as Reaction is not product or reactant favoured. Reaction regarded as
not taking place at all. Equilibrium mixture comprised of both reactants going to completion if
and products. above this value.

However, Kc gives no information about the rate of reaction. It tells us how far but not how fast the
reaction goes.

The value of an equilibrium constant can be determined by experiment if the following steps are
followed:

1. Write the balanced equation


2. Mix known molar quantities of either the reactants or the products or both.
3. Allow the mixture to reach equilibrium in a closed system.
4. Determine the equilibrium concentration of at least one substance in the equilibrium mixture.
5. Deduce the equilibrium concentrations of the other substances in the mixture.
6. Substitute the calculated equilibrium concentrations in the expression for Kc
7. Repeat the determination of Kc using different initial concentrations at the same temperature.
GETTING TO THE NUB OF CHEMISTRY
ICE (Initial, Change, Equilibrium) acts as a useful mnemonic when performing calculations based on the
above.

When 5.0 moles of A is mixed with 5.0 moles of B in a container of volume 12 dm3, an equilibrium is
established (A + 2B ⇌ 2C) that contains 3.0 moles of C. Find Kc at this temperature.

1. Write the know initial and equilibrium moles below the relevant species.

A + 2B ⇌ 2C
Initial 5.0 5.0
Change
Equilibrium 3.0

2. In this reaction, there were 0 moles of C before the reactants were mixed and started reacting.
Given we know the moles of C at equilibrium, we therefore know the change in moles.

A + 2B ⇌ 2C
Initial 5.0 5.0 0
Change +3.0
Equilibrium 3.0

3. Once we know the change in moles of one species, we can use the equation’s molar ratios to
determine the change in moles of the remaining species.

A + 2B ⇌ 2C
Initial 5.0 5.0 0
Change -1.5 -3.0 +3.0
Equilibrium 3.0

4. Now the remaining equilibrium quantities can be determined.

A + 2B ⇌ 2C
Initial 5.0 5.0 0
Change -1.5 -3.0 +3.0
Equilibrium 3.5 2.0 3.0

Next, we have to work out their respective concentrations. The volume of the container is 12 dm3. So,

0.2
[A] = = 0.292 mol dm-3
3(

(.4
[B] = = 0.167 mol dm-3
3(

0.4
[C] = 3( = 0.25 mol dm-3
GETTING TO THE NUB OF CHEMISTRY
The equilibrium expression for this reaction is:

[𝑪]𝟐
𝑲𝒄 =
[𝐀][𝐁]𝟐

Insert the concentrations:

[4.(2]( (;<= >;)* )( (;<= >;)* )( 3


𝐾' = = 7.7 (units = ;<= >;)* (;<= >;)* )( = ;<= >;)* (;<= >;)* )( = ;<= >;)*
[4.(7(][4.389](

Kc = 7.7 mol-1 dm3

Sometimes questions don’t give you the volume, which means you can’t work out the equilibrium
concentrations. Whenever this happens, check the total number of moles on each side of the equation.
You should find they’re the same (this will also mean Kc will have no units). Let’s consider the following
reaction:

CH3OH + HCOOH ⇋ HCOOCH3 + H2O

𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 HCOOCH0 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 H( O


[HCOOCH0 ][H( O] 𝑣𝑜𝑙𝑢𝑚𝑒 𝑥 𝑣𝑜𝑙𝑢𝑚𝑒
𝐾' = =
[CH0 OH][HCOOH] 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 CH0 OH 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 HCOOH
𝑥
𝑣𝑜𝑙𝑢𝑚𝑒 𝑣𝑜𝑙𝑢𝑚𝑒

Notice that volume appears twice both above and below the fraction line and can be cancelled out.

𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 HCOOCH0 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 H( O


𝑣𝑜𝑙𝑢𝑚𝑒 𝑥 𝑣𝑜𝑙𝑢𝑚𝑒 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 HCOOCH0 𝑥 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 H( O
=
𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 CH0 OH 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 HCOOH 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 CH0 OH x 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 HCOOH
𝑣𝑜𝑙𝑢𝑚𝑒 𝑥 𝑣𝑜𝑙𝑢𝑚𝑒

So, in this case, volume isn’t needed to work out Kc

Another type of question is one that involves our introducing an algebraic term. In these, you’re typically
given Kc and have to work out a concentration. Here’s an example.

At 440oC, the reaction:

2HI ⇋ H2 + I2

…has a Kc of 0.02

Suppose 1 mole of hydrogen and 1 mole of iodine are mixed together in a 1.0 dm3 vessel at 440oC. What
are the concentrations of HI, H2 and I2 at equilibrium?
GETTING TO THE NUB OF CHEMISTRY
Let’s begin by laying out our ICE table

2HI ⇋ H2 + I2
I 0 1 1
C
E

Let the change in the moles of H2 equal ‘y’. This will be a decrease in moles. Because of the molar ratios
of the equation, we will see the same decrease of I2. This gives us the equilibrium moles of H2 and I2 in
algebraic form.
2HI ⇋ H2 + I2
I 0 1 1
C -y -y
E 1-y 1-y

We can use the molar ratios of the equation to determine the change in HI. This will be an increase in
moles.
2HI ⇋ H2 + I2
I 0 1 1
C +2y -y -y
E 2y 1-y 1-y

We’ve rather fortuitously been given a volume of 1 dm3 so the concentrations will be identical to the
mole values above. However, it’s also the case that the total number of moles on both sides of the
equation are the same so, as we saw previously, the volumes will cancel out (always look out for this).

[𝐻( ][𝐼( ]
𝐾' =
[𝐻𝐼](

Let’s insert the values (remember, Kc = 0.02):

(1 − 𝑦)(1 − 𝑦) (1 − 𝑦)(
0.02 = 𝒐𝒓 0.02 =
(2𝑦)( (2𝑦)(

The version on the right makes the next step much easier to see – square root everything!

1−𝑦
0.14 =
2𝑦

0.28y = 1-y

1.28y = 1

3
y = 3.(@ = 0.78 𝑚𝑜𝑙 𝑑𝑚A0
GETTING TO THE NUB OF CHEMISTRY
Compounds A and B react together to form an equilibrium mixture containing
compounds C and D according to the equation

2A + B ⇌ 3C + D

a) A beaker contained 40 cm3 of a 0.16 mol dm–3 aqueous solution of A. 9.5 × 10–3 mol of B and
2.8 × 10–3 mol of C were added to the beaker and the mixture was left to reach equilibrium. The
equilibrium mixture formed contained 3.9 × 10–3 mol of A. Calculate the amounts, in moles,
of B, C and D in the equilibrium mixture.

We’re given a volume and concentration of A which means we can work out the moles of A.

40
𝑀𝑜𝑙𝑒𝑠 𝑜𝑓 𝐴 = 𝑥 0.16 = 6.4 𝑥 10A0 𝑚𝑜𝑙 𝑑𝑚A0
1000

We can now construct our ICE table and fill the initial moles of A, B, C and D and the equilibrium moles of
A.

2A + B ⇌ 3C + D
–3 –3 –3
I 6.4 × 10 9.5 × 10 2.8 × 10 0
C
E 3.9 × 10–3

Initial moles of A – equilibrium moles of A = change in moles of A.

2A + B ⇌ 3C + D
–3 –3 –3
I 6.4 × 10 9.5 × 10 2.8 × 10 0
C -2.5 × 10–3
E 3.9 × 0–3

Using the molar ratios given by the reaction equation, we can work out the change of the other
substances in the equation.

2A + B ⇌ 3C + D
–3 –3 –3
I 6.4 × 10 9.5 × 10 2.8 × 10 0
C -2.5 × 10–3 -1.25 × 10–3 +3.75 × 10–3 +1.25 × 10–3
E 3.9 × 0–3

This now allows us to calculate the final moles.

2A + B ⇌ 3C + D
–3 –3 –3
I 6.4 × 10 9.5 × 10 2.8 × 10 0
C -2.5 × 10–3 -1.25 × 10–3 +3.75 × 10–3 -1.25 × 10–3
E 3.9 × 0–3 8.25 × 10–3 6.75 × 10–3 1.25 × 10–3
GETTING TO THE NUB OF CHEMISTRY
So,

The amount of B = 8.25 × 10–3 mol


The amount of C = 6.75 × 10–3 mol
The amount of D = 1.25 × 10–3 mol

b) Give the expression for the equilibrium constant (Kc) for this equilibrium and its units.

[𝑪]𝟑 [𝑫]
𝑲𝒄 =
[𝐀]𝟐 [𝐁]

(;<= >;)* )* (;<= >;)* ) (;<= >;)* )* (;<= >;)* )


Units = (;<= >;)* )( (;<= >;)* ) = (;<= >;)* )( (;<= >;)* )
= 𝑚𝑜𝑙 𝑑𝑚A0

c) A different equilibrium mixture of these four compounds, at a different temperature, contained 0.21
mol of B, 1.05 mol of C and 0.076 mol of D in a total volume of 5.00 × 102 cm3 of solution. At this
temperature the numerical value of Kc was 116.

Calculate the concentration of A, in mol dm–3, in this equilibrium mixture. Give your answer to the
appropriate number of significant figures.

[𝑪]𝟑 [𝑫]
𝑲𝒄 =
[𝐀]𝟐 [𝐁]

Rearrange so that [A] is the subject of the equation (this is a mark and where most people go wrong with
this question):

[𝑪]𝟑 [𝑫]
[𝑨]𝟐 =
𝑲𝒄 [𝐁]

Now insert values (moles, volume and Kc)

𝟏. 𝟎𝟓 𝟎. 𝟎𝟕𝟔
( 𝟎. 𝟓 )𝟑 ( 𝟎. 𝟓 )
𝟐
[𝑨] = = 𝟎. 𝟐𝟖𝟗
𝟎. 𝟐𝟏
𝟏𝟏𝟔 ( 𝟎. 𝟓 )

[𝑨] = √𝟎. 𝟐𝟖𝟗 = 𝟎. 𝟏𝟕 𝒎𝒐𝒍 𝒅𝒎A𝟑


GETTING TO THE NUB OF CHEMISTRY
3.1.7 Oxidation, Reduction and Redox Equations
Isn’t Redox something you add to the bath? Are you taking the mick? No, mum and dad get gallons of
the stuff every Christmas. That’s RADOX! Oh, yeah – so what’s redox then? Let’s consider
‘displacement’ reactions, which you’ve done for a few years now. Here’s a classic example:

Fe(S) + CuSO4(aq) → FeSO4(aq) + Cu(s)

Let’s pay attention to the state symbols. If an ionic compound is aqueous, that means the lattice has
broken apart and the ions are separated from each other. So, a more accurate equation would be:

Fe(S) + Cu2+(aq) + SO42-(aq) → Fe2+(aq) + SO42-(aq) + Cu(s)

SO42-(aq) appears on both sides and is a spectator ion. This means it’s playing no part in the reaction,
which is actually:

Fe(S) + Cu2+(aq) → Fe2+(aq) + Cu(s)

We can break this down further into two half equations (NNOOOO! Please no!):

1. Fe(s)→ Fe2+(aq) + 2e-


2. Cu2+(aq) + 2e- → Cu(s)

Recall the mnemonic OIL RIG: Oxidation Is the Loss of Electrons and Reduction Is the Gain of Electrons.
Half-equation 1. involves oxidation and 2. involves reduction. This means the overall reaction involves
both oxidation and reduction and so is a REDOX reaction. This is the case with all ‘displacement’
reactions, even the ones you did at GSCE involving the halogens:

Cl2(aq) + 2Br-(aq) à 2Cl-(aq) + Br2(aq)

1. Cl2(aq) + 2e- → 2Cl-(aq) Reduction


2. 2Br-(aq) → Br2(aq) + 2e- Oxidation

Do we have to do half-equations? Oh yes, we absolutely do. Groan! A redox reaction is one that involves
both oxidation and reduction. In reactions involving ionic compounds, we simply (simply you say) have to
look at the changes in charge of the ions to determine what is oxidised and what is reduced). However,
what about a redox reaction like this:

H2(g) + ½O2 (g) → H2O(l)

Here it’s difficult to pick out the separate oxidation and reduction processes because there isn’t a clear
transfer of electrons. So, we use oxidation numbers (or oxidation states) to help us work out what has
been oxidised and reduced.
GETTING TO THE NUB OF CHEMISTRY
Rules for Assigning Oxidation Numbers

1. The oxidation number of an uncombined element is zero e.g. Na and I2


2. In ionic compounds, the oxidation number of a monatomic ion equal is to its charge.
3. The sum of the oxidation numbers of the atoms in a neutral compound (e.g. NaCl) is zero
4. The sum of the oxidation numbers in a polyatomic ion (e.g. NO3-) equals the charge on the ion
5. Some elements nearly always have the same oxidation number in their compounds and act as
reference points in assigning oxidation numbers to other elements.

• RangeK of oxidation
Na states
+1 – thisHis limited
+1 by the
Except in metal
position of an hydrides e.g.
element in theLiAlH 4
Periodic Table as the diagram below
shows.
Mg Ca +2 F -1
• Al example, N is in+3group 5Cl
For and could lose Except in compounds
up to 5 electrons withstate
(oxidation O and
+5) Fore.g.
gainNaClO
3 electrons (oxidation state
-3) before another shell of electrons is affected.

O -2 Except in peroxides, superoxides and fluoride e.g. H2O2, OF2
• With the transition metals, they usually have positive oxidation states as they are metals, losing electrons until they
reach a Group 0 noble gas electron configuration, but not usually being higher than +7 in practice
Other elements can have a range of oxidation numbers as the below shows:
+4 to –4 +5 to –3

max +5 max +7
C N +6 to –2

S
Ti V Cr Mn Fe Br
+7 to –1

max +4 max +6 highest


seen is +6

• In molecules and polyatomic ions (e.g. SO42-), the more electronegative element is
3)
assumed to be the negative ion.
Give the likely range of oxidation state for each of the following elements.

a)I need to see some


phosphorus examples. Of course! Let’s keep is simple to start with. What’s the oxidation number
PPPPPPPPPPPPPPPPPP

b)ofsilicon
N in ammonia?PPPPPPPPPPPPPPPPPP
NH3
c) iodine PPPPPPPPPPPPPPPPPP

• H is +1 unless
d) gallium PPPPPPPPPPPPPPPPPP
we have a metal hydride (we don’t).
• The sum of the oxidation numbers in a neutral compound is 0. I
4) • whether
State Let theeach
ox. no. of N oxidation
of these be represented
states is by y: to occur in stable compounds.
likely

a) Sc +3 PPPPP.
y + 3(+1) = 0
b) Ni +2 PPPPP. ∴ y = -3
c) K +2 PPPPP.
What about S in sulfur trioxide?
d) W +6 PPPPP.
SO3
e) Se -3 PPPPP.

f) Mo•+7The ox. no. of O is -2 unless it’s bonded to fluorine or is in a peroxide or superoxide (it isn’t).
PPPPP.
• The sum of the oxidation numbers in a neutral compound is 0.
g) Sb +5 PPPPP.
y + 3(-2) = 0
∴ y = +6
GETTING TO THE NUB OF CHEMISTRY
How about S in sulfuric acid?

H2SO4

• H is +1 unless we have a metal hydride (we don’t).


• The ox. no. of O is -2 unless it’s bonded to fluorine or is in a peroxide or superoxide (it isn’t).
• The sum of the oxidation numbers in a neutral compound is 0.

2(+1) + y + 4(-2) = 0
∴ y = +6

Now let’s consider Cr in potassium dichromate:

K2Cr2O7

• K is always +1.
• The ox. no. of O is -2 unless it’s bonded to fluorine or is in a peroxide or superoxide (it isn’t).
• The sum of the oxidation numbers in a neutral compound is 0.
• Let the ox. no. of Cr be represented by y:

2(+1) + 2y + 7(-2) = 0
∴ y = +6

What about ions? Let’s consider P in the phosphate ion.

PO43-

• The ox. no. of O is -2 unless it’s bonded to fluorine or is in a peroxide or superoxide (it isn’t).
• The sum of the oxidation numbers in a polyatomic ion equals the charge on the ion.

y + 4(-2) = -5
∴ y = +5

And a final example, what about S in the sulfite ion.

SO32-

• The ox. no. of O is -2 unless it’s bonded to fluorine or is in a peroxide or superoxide (it isn’t).
• The sum of the oxidation numbers in a polyatomic ion equals the charge on the ion.

y + 3(-2) = -2
∴ y = +4
GETTING TO THE NUB OF CHEMISTRY
Has this got anything to do with the roman numerals I sometimes see in formulae? Yes. We use roman
numerals in compounds’ names whenever there is the possibility of ambiguity. For example, iron
chloride could denote FeCl2 or FeCl3 depending on Fe’s charge / oxidation state. We would make it clear
which we were talking about by including a roman numeral i.e. iron (II) chloride and iron (III) chloride.
Similarly, sodium chlorate could be NaClO or NaClO3. So, we’d make it clear which we were talking about
by naming them sodium chlorate (I) and sodium chlorate (V). Notice the numeral goes after the bit of the
compound’s name containing the ambiguous element. Here are a few more examples:

• copper (I) oxide – Cu2O


• copper (II) sulfate – CuSO4
• potassium dichromate (VI) – K2Cr2O7
• potassium manganate (VII) – KMnO4

Okay… You said we could use these to track the elements being oxidised and reduced in a chemical
reaction. Shall we look at that now? Absolutely. Let’s consider the displacement reaction between
chlorine and bromide ions.

• If the oxidation number of an element increases, then it has been oxidised.


• If it decreases, it has been reduced.

This prompts our need to understand a couple of definitions.

• An oxidising agent is itself reduced


• An reducing agent is itself oxidised

In the reaction above, chlorine is the oxidising agent and bromide is the reducing agent.
GETTING TO THE NUB OF CHEMISTRY
I thought you said we were going to be doing lots of half equations. You were just teasing me, weren’t
you? Oh dear… my answer is…

A half equation shows what has been oxidised and reduced. Here is a reminder of the ones we hsowed
on the first page:

1. Cl2(aq) + 2e- → 2Cl-(aq) Reduction


2. 2Br-(aq) → Br2(aq) + 2e- Oxidation

I’m afraid further half-equations we consider are going mostly going to be more complicated than these
examples. So, we need to know a series of rules on how to construct them:

1. Write down formulae of reactants and products


2. Balance any atoms that are not Hydrogen or Oxygen
3. Add H2O to balance Oxygen atoms
4. Add H+ ions to balance Hydrogen atoms
5. Balance charges by adding electrons
6. (Add state symbols)

Write a ½ equation for the conversion of nitric acid into nitrogen dioxide gas

Write a ½ equation for the conversion of sulphur dioxide gas into sulphuric acid

Write a ½ equation for the conversion of chlorine gas into chlorate (V) ions

Write a ½ equation for the conversion of dichromate (VI) ion into chromate (III) ions

Write a ½ equation for the conversion of manganate (VII) ion into manganese (II) ions
GETTING TO THE NUB OF CHEMISTRY
So, what’s the point of these half equations? We can combine them to get the overall reaction
equation. Again, there is a set of rules we can follow:

1. Write down the two half-equations


2. If necessary, multiply the half-equations to balance electrons
3. Add together the two half-equations
4. Cancel e-, H2O and H+ where necessary
5. Check atoms and charges balance
6. Add state symbols

Use the two ½ equations to work out a balanced equation for the reaction between magnesium and
silver ions in solution:

First, we write down the two half-equations.

When two half-equations are added together, the number of electrons much cancel out. As things stand,
this won’t currently happen. So, we multiply the silver ions half equation by two before adding the half-
equations together.

Use the two ½ equations to work out a balanced equation for the oxidation by chlorine of sulfur
dioxide to sulfuric acid:
GETTING TO THE NUB OF CHEMISTRY
Use the two ½ equations to work out a balanced equation for the oxidation of hydrogen iodide to
iodine using concentrated sulfuric acid

Here are the two half-equations.

We need to multiply the second half-equation by three so that the electrons cancel out.

When manganate (VII) ions react with Fe (II) ions, manganese (II) ions and Fe (III) ions are produced.
Write ½ equations then deduce the full equation

Here are the two half-equations.

We need to multiply the second half-equation by five so that the electrons cancel out.
GETTING TO THE NUB OF CHEMISTRY
Questions can work in a different way as the following shows.
GETTING TO THE NUB OF CHEMISTRY
Thermodynamics 3.1.8
3.1.8.1 Born–Haber cycles

This topic builds on Energetics 3.1.4. We begin by exploring Hess Cycles that focus specifically on the
lattice enthalpy of ionic compounds. According to AQA, Lattice enthalpy can be defined as either
enthalpy of lattice dissociation (HLED) or enthalpy of lattice formation (HLEF).

We begin by understanding the definitions of these two versions of lattice enthalpy, starting with the
enthalpy of lattice dissociation.

HLED refers to the enthalpy change when one mole of an ionic compound is vaporised to form its
gaseous ions under standard conditions. As this involves overcoming the strong electrostatic forces
attraction between oppositely charged ions, energy is required so HLED is always endothermic.

HLED (NaCl(s)) NaCl(s) → Na+(g) + Cl-(g) (H = +787 kJ mol-1)

HLEF refers to the enthalpy change when one mole of an ionic compound is formed from its gaseous
ions under standard conditions. As this involves the creation of strong electrostatic forces of attraction
between oppositely charged ions, energy is released so HLED is always exothermic.

HLEF (NaCl(s)) Na+(g) + Cl-(g) → NaCl(s) (H = -787 kJ mol-1)


GETTING TO THE NUB OF CHEMISTRY
Lattice enthalpies cannot be measured directly so they’re calculated using a Hess cycle, which have the
special name of Born-Haber cycle (this tells us that lattice enthalpies are involved).

The simplest Born-Haber cycle in AQA’s A level involves two new enthalpies, enthalpy of hydration
(Hhyd) and enthalpy of solution (Hsol).

Both Hhyd and enthalpy of solution HSOL result in separated aqueous ions e.g. Na+(aq) and Cl-(aq). When
ions are in solution, water molecules surround them due to electrostatic attractions between the ions
and water’s partial charges. Energy is released when these are formed.

The enthalpy of solution (Hsol) is the enthalpy change accompanying dissolving 1 mole of an ionic solid
in a large excess of water. This is a two-step process: the ionic lattice is broken up, which requires energy
and so is endothermic, and then the resulting ions form electrostatic attractions with water molecules,
which is exothermic. The overall enthalpy change may therefore be endothermic or exothermic.

Hsol (NaCl(s)) NaCl(s) → Na+(aq) + Cl-(aq) (H = +3.9 kJ mol-1)

The enthalpy of hydration (HHYD) is the enthalpy change when one mole of gaseous ions become
hydrated (dissolved in water). As ions form electrostatic attractions with water molecules, this enthalpy
change is always exothermic.

Hhyd (Cl-(g)) Cl-(g) + aq → Cl-(aq) (H = -363 kJ mol-1)


GETTING TO THE NUB OF CHEMISTRY
Here’s an example question:

Using the following data, calculate the standard enthalpy of solution of magnesium chloride.

HLED (MgCl2(s)) +2493 kJ mol-1 (endothermic as energy it needed to break up the lattice)
Hhyd (Mg2+(g)) -1920 kJ mol-1 (exothermic as electrostatic attractions are formed with H2O)
Hhyd (Cl-(g)) -364 kJ mol-1 (exothermic as electrostatic attractions are formed with H2O)

WATCH OUT! AQA can give you, or ask for, lattice enthalpies of formation OR dissociation so take care to
write the arrow in the correct direction.

The lattice enthalpy of a compound is an indication of the strength of the ionic bonding – the greater the
magnitude of the lattice enthalpy, the stronger the bonding. If we remember from GTTNOC – 3.1.3
Bonding:

𝒕𝒉𝒆 𝒔𝒊𝒛𝒆 𝒐𝒇 𝒄𝒉𝒂𝒓𝒈𝒆𝒔 𝒐𝒏 𝒕𝒉𝒆 𝒊𝒐𝒏𝒔


The strength of the electrostatic attraction in an ionic compound is  𝒕𝒉𝒆 𝒓𝒂𝒅𝒊𝒊 𝒐𝒇 𝒕𝒉𝒆 𝒊𝒐𝒏𝒔

So, the factors that affect lattice energies are:

1. The sizes of the ions.


2. The charges on the ions

Combined together, these determine the ion’s charge density. A greater charge density = stronger ionic
bonding.

Mg2+ has a greater charge and smaller radius than Na+ so we would expect a higher lattice enthalpy for
MgCl2 compared with NaCl, which is indeed the case (2493 kJmol-1 vs 787 kJmol-1). Similarly, we would
expect NaF to have a higher lattice enthalpy than NaCl because F- is a smaller ion than Cl- (NaF (904
kJmol-1); NaCl (787 kJmol-1)). Notice, charge has the greatest effect.

Here’s another example: Al2O3 > MgO > CaO. Al3+ has the highest charge and is the smallest ion =
greatest charge density. Mg2+ [1s22s22p6]+ is a smaller ion that Ca2+ [1s22s22p63s23p6]+.
GETTING TO THE NUB OF CHEMISTRY
There is a second Born-Haber cycle that appears in AQA’s specification that involves several enthalpy
changes, some of which are new to this topic.

Enthalpy of formation (Hf). The enthalpy change when one mole of a substance is formed from its
constituent elements with all substances in their standard states. This is exothermic for most substances.

Mg(s) + ½O2(g) → MgO(s)

Ionisation energy (Hie). First ionisation energy = the enthalpy change when each atom in one mole of
gaseous atoms loses one electron to form one mole of gaseous 1+ ions. This is an endothermic process as
electrostatic forces of attraction must be overcome.

Mg(g) → Mg+(g) + e–

Second ionisation energy = the enthalpy change when each ion in one mole of gaseous 1+ ions loses one
electron to form one mole of gaseous 2+ ions. This is an endothermic process as electrostatic forces of
attraction must be overcome.
Mg+(g) → Mg2+(g) + e–

Enthalpy of atomisation (Hat). Enthalpy change when one mole of gaseous atoms is produced from an
element in its standard state. This is an endothermic process as electrostatic forces of attraction must be
overcome.
Mg(s) → Mg(g)

½ Cl2(g) → Cl(g)

Bond dissociation. The enthalpy change when one mole of covalent bonds is broken in the gaseous
state. With diatomic molecules such as O2 and Cl2, this is the equivalent of 2 x enthalpy of atomisation
when the molecule is gaseous. Bond enthalpy is an endothermic process as electrostatic forces of
attraction must be overcome.

O2(g) → 2O(g)

Electron affinity (Hea). First electron affinity = the enthalpy change when each atom in one mole of
gaseous atoms gains one electron to form one mole of gaseous 1– ions. This is an exothermic process for
many non-metals as the resulting electrostatic force of attraction between the nucleus and gained
electron releases energy.
O(g) + e– → O– (g)

Second electron affinity = the enthalpy change when each ion in one mole of gaseous 1– ions gains one
electron to form one mole of gaseous 2– ions. This is an endothermic process as a negatively charged
electron is being added to a negatively charged ion (repulsion).

O– (g) + e– → O2– (g)


GETTING TO THE NUB OF CHEMISTRY
Putting all of these together, we are able to construct the following Born-Haber cycle for, say, sodium
chloride:

Here’s another example:

WATCH OUT FOR:


• If you’re given the bond enthalpy of a diatomic molecule instead of Hat, remember that the
definition for the former refers to 1 mole of covalent bonds whereas the latter refers to 1 mole
of gaseous atoms i.e. bond enthalpy = 2Hat (O2(g) → 2O(g)).
• Is it HLED or HLEF you’re given/asked for?
• In a compound with two halide ions (e.g. MgCl2), don’t forget it will be 2 x Hea
GETTING TO THE NUB OF CHEMISTRY
Here’s an example question:
GETTING TO THE NUB OF CHEMISTRY
Here’s another (you won’t be asked to construct a Born-Haber cycle from scratch in an exam but knowing
how to do so makes your knowledge and understanding much stronger).

Calculate the first electron affinity of iodine given that the lattice enthalpy of dissociation of calcium
iodide is +2054 kJ mol-1 and its enthalpy of formation is –535 kJ mol-1.
GETTING TO THE NUB OF CHEMISTRY

Calculate the lattice enthalpy formation of aluminium oxide given that the enthalpy of formation
aluminium oxide is –1669 kJ mol-1.
GETTING TO THE NUB OF CHEMISTRY
The values used in Born-Haber cycles are based on experimental data. However, it is possible to calculate
theoretical lattice energies using the Born-Landé equation:

Born-Landé equation Extremely simplified version


𝑐ℎ𝑎𝑟𝑔𝑒
𝐿𝑎𝑡𝑡𝑖𝑐𝑒 𝑒𝑛𝑡ℎ𝑎𝑙𝑝𝑦 
𝑟𝑎𝑑𝑖𝑢𝑠

Don’t worry, you don’t need to learn this equation. You just need to appreciate that it can be used to
calculate lattice energies. E.g.

The Born-Landé equation assumes a purely ionic model. Earlier in our A level, we began to understand
that ionic and covalent bonding isn’t simply the either/or we may have understood it to be at GCSE.
Instead, we have a bonding spectrum, with 100% covalent at one end and 100% (perfect) ionic at the
other. In between, there is polar covalent and polar ionic. The size of the difference in two elements’
electronegativities gives us an idea of where the bonding in a substance lies on the spectrum.

0 Electronegativity Difference 4

d-
d+

X Y X Y X- Y+ X- Y+

Pure covalent Polar covalent Polar ionic Pure ionic


Electrons not Distorted ions
equally shared

When we compare theoretical lattice energies with experimental (Born-Haber cycle) values, we discover
some are in very close agreement while others are less so. Which it is depends on just how purely ionic a
compound is.

Compound Theoretical Experimental value % Difference Difference in


formation value (Born-Haber cycle) (kJ mol-1) electronegativity
(kJ mol-1) (kJ mol-1)
NaCl -766 -771 0.6 2.1
LiCl -833 -846 1.5 2.0
MgCl2 -2326 -2526 7.9 1.8
AgCl -770 -905 14.9 1.1

1. The experimental formation value is always more negative / more exothermic than the theoretical
value.
2. The smaller the difference in electronegativity, the greater the % difference between the theoretical
and experimental values.

When asked to compare theoretical and experimental lattice energies, little difference = spherical ions
matching a purely ionic model. A large difference means it’s not purely ionic and has covalent character.
GETTING TO THE NUB OF CHEMISTRY
3.1.8.2 Gibbs free-energy change, ΔG, and entropy change, ΔS

Why do particular chemical reactions occur?

Many chemical processes tend to proceed in a direction in which they are said to be spontaneous. This
raises an obvious question: What makes a reaction spontaneous? What drives the reaction in one
direction and not the other? The answer is a brand-new concept for us: entropy.

The first law of thermodynamics states:

Energy cannot be created or destroyed, only transferred from one store to another.

Here are two versions of the second law:

• Version 1: Heat energy always flows spontaneously from a hot to a cold system and never
naturally the other way round.

• Version 2: The entropy of the universe increases in the course of any spontaneous change.

In thermodynamics spontaneous means not needing to be driven by doing work of some kind. Broadly
speaking, ’spontaneous’ is a synonym of ’natural’.

A spontaneous process refers to a physical or chemical change that occurs with no outside intervention
other than maybe a little energy to get the process started e.g. a spark for combustion.

Imagine a stationary football on a grassy slope. After a little push (the equivalent of the activation
energy), the ball will roll down the hill without any further intervention. This is a spontaneous process.

What about if we want the ball to go back up the slope? This time, we have to continuously put energy in
for it to roll uphill. This is not a spontaneous process.

Diffusion is an example of a spontaneous process.

Let’s return to the second version of the second law:

The entropy of the universe increases in the course of any spontaneous change.

So, what is entropy? Mathematically, the entropy of a chemical system measures the number of ways,
W, of arranging the particles and sharing out the energy between the particles. Another way of saying
this is that the entropy, S, of a substance is a measure of how distributed and disordered its particles and
their energy are.
GETTING TO THE NUB OF CHEMISTRY
This last statement needs unpicking. We’ll begin with the idea of disorder, which is inextricably linked to
probability.

Imagine we throw 20 pennies up in the air. Here are three possible ways they could land:

Highly ordered; very improbable Disorded; very probable Even greater disorder
→ low entropy → high entropy → even higher entropy

Now let’s apply the above to the particle model we learned in y7:

Highly ordered → low entropy Highly disordered → high entropy

As we saw earlier, entropy also reflects the extent to which energy can be distributed across a system.
Solids have relatively low values for standard molar entropies. It might help for us to compare the
entropies of two different solid substances.

Substance So (J K-1 mol-1)


Diamond 2.4
Aluminium 27.2

In diamond, the carbon atoms are held in place by strong, highly directional covalent bonds. Iron has a
higher value for its standard molar entropy because its metallic bonding is not directional. The heavier,
larger Aluminium ions can vibrate more freely and share out their energy in more ways than carbon
atoms in diamond.

Liquids typically have higher standard molar entropies than solids because the atoms or molecules are
free to move. There are many more ways of distributing the particles and energy.

Molecules with more atoms have higher standard molar entropies because they can vibrate, rotate and
arrange themselves in yet more ways.
GETTING TO THE NUB OF CHEMISTRY
Gases have higher standard molar entropies than comparable liquids because the atoms or molecules
are not only free to move but are also widely spaced. There are even more ways of distributing the
particles and energy.

Substance So (J K-1 mol-1)


Methane (g) 186.2
Ethane (g) 229.5
Propane (g) 269.9
Butane (g) 310.1
Pentane (l) 261.1

As with liquids, molecules with more atoms have even higher standard molar entropies because they can
vibrate, rotate and arrange themselves in even more ways.

There is a sharp increase in entropy at the melting point as the structure change from a solid to the more
random arrangement of particles when liquid. There is an even greater increase in entropy at the boiling
point.

Our knowledge up to this point allows us to make a qualitative judgement regarding the change in
entropy, S, in many reactions.
CaCO3 (s) → CaO (s) + CO2 (g)

1 mole solid → 1 mole solid and 1 mole gas; Gas has high disorder; The entropy of the products is
greater than the reactants; There is a positive increase in entropy in this reaction

CH4(g) + 2O2(g) → CO2(g) + 2H2O(g)

The entropy change is close to zero as go from 3 moles of gas to 3 moles of gas.
GETTING TO THE NUB OF CHEMISTRY
So, the thermal decomposition of calcium carbonate results in an increase in entropy in this reaction.
However, thermal decomposition is not a spontaneous reaction. As with rolling a ball up a hill, energy
needs to be constantly supplied for the reaction to take place. How come? To answer this, we need to
unpack and rephrase the second law of thermodynamics:

‘For a spontaneous process, the total entropy change in a system and in the system’s surroundings will
increase’

System: this is the reactants and products


Surroundings: anything beyond the reactants and products e.g. the environment, such as water
molecules if the reaction takes place in solution, or the container the reaction takes place in.

So, the change entropy (S) can be expressed as follows:

STOTAL = SSYSTEM + SSURROUNDING

Reactions are only feasible (spontaneous) if the total entropy change is +ve. All spontaneous changes
correspond to an increase in the number of ways of arranging energy and particles in the system and
surroundings taken together.

Entropy Change of a System (SSYSTEM) is the difference between the sums (Σ) of the standard molar
entropies of the products and reactants. Standard molar entropies have units of J K–1 mol–1.

SSYSTEM = ΣSPRODUCTS – ΣSREACTANTS

For example, calculate ΔSSYSTEM for the reaction of nitrogen and hydrogen to form ammonia given the
following standard molar entropy values:

N2(g) + 3H2(g) → 2NH3(g)


GETTING TO THE NUB OF CHEMISTRY
The value of ΔSSURROUNDINGS depends on the size of the enthalpy change, ΔH, and the temperature, T.

ΔH
ΔSSURROUNDINGS = -
T

• T is the temperature in kelvin


• ΔH is enthalpy change in joules per mole (take care with units – values for the enthalpy may be
given in kJ!)

The equation above doesn’t feature in the AQA specification so you don’t need to learn it. However, I
think covering it provides us with a better understanding of what is still to come.

When ΔSTOTAL is +ve, reactions are spontaneous.

ΔSTOTAL = ΔSSYSTEM + ΔSSURROUNDINGS

ΔH
As we’ve just seen, ΔSSURROUNDINGS = -
T

So…

ΔH
ΔSTOTAL = ΔSSYSTEM -
T

Multiply both sides by T.

TΔSTOTAL = TΔSSYSTEM – ΔH

-TΔSTOTAL is called the Gibbs free energy change, ΔG. Multiply the formula above by -1:

-TΔSTOTAL = ΔH – TΔSSYSTEM

ΔG = ΔH – TΔS

Since -TΔSTOTAL = ΔG, for ΔSTOTAL to be positive, ΔG must be negative. So…

For all spontaneous changes, ΔG must be negative. This is another way of stating the Second Law of
Thermodynamics.

Units
ΔG kJ mol-1
ΔH kJ mol-1
ΔSSYSTEM J mol-1 K-1

• Watch out for kJ vs J when doing calculations.


GETTING TO THE NUB OF CHEMISTRY
GETTING TO THE NUB OF CHEMISTRY
ΔG = ΔH – TΔS

We can rearrange the above as follows:

ΔG = –ΔST + ΔH
y = mx + c
Plot ΔG vs. T (when T = 0, ΔG = ΔH)
(Remember that ΔG must be negative if a reaction is to be spontaneous (feasible))

Feasability
ΔH ΔS Graph
(i.e. ΔG is negative)

+ve +ve At high T

-ve +ve Always

+ve -ve Never

-ve -ve At low T


GETTING TO THE NUB OF CHEMISTRY
Plotting a graph, like those sketched above, can determine the temperature at which a reaction becomes
feasible. This can also be done by making ΔG = 0 and then rearranging the formula ΔG = ΔH – TΔS so that
T is its subject.

For example, given ΔH = +206 kJmol-1 and ΔS = 216 Jmol-1K-1 for a given reaction, calculate the
temperature which this reaction is feasible.

The reaction becomes feasible when ΔG = 0.

Kinetics vs. Thermodynamics

Although a reaction may be thermodynamically feasible, it still may not be able to occur due to kinetic
factors.

• A reaction may require a large activation energy that is not provided by the reaction conditions.
Therefore, it will not occur spontaneously.

E.g. methane does not react with oxygen in the air at room temperature even though it’s
thermodynamically feasible.
GETTING TO THE NUB OF CHEMISTRY
GETTING TO THE NUB OF CHEMISTRY
Extension: Gibbs Energy and Equilibrium
This is NOT in the A Level course but is too good to ignore.

In a spontaneous reaction at constant temperature and pressure, ΔG < 0. So, the system moves to a
lower Gibbs energy.

Let’s consider how Gibbs energy changes during the course of a reaction. The graph below shows how G
changes for a reaction that goes virtually to completion.

At any stage during the reaction, the Gibbs energy of the system decreases as the reaction proceeds. The
lowest Gibbs energy is achieved when the reactants are completely converted to products. An example
of such a reaction would be during the formation of water from hydrogen and oxygen.

Contrast this with a reaction that comes to equilibrium with observable amounts of reactants and
products present, such as the N2O4(g)/ NO2(g) reaction. In this case, the Gibbs energy plot looks like the
below.

This demonstrates the one of the fundamental statements of chemistry:

A system comes to equilibrium when it reaches its minimum Gibbs energy i.e. when ΔG is most negative.

The graph shows that, if you start with 100% products, the system can lower its Gibbs energy by the
reverse reaction taking place until we reach the minimum Gibbs energy. The same equilibrium
composition is obtained whatever the direction from which equilibrium is approached.

Changing the temperature changes the Gibbs energy which changes the position of equilibrium.
GETTING TO THE NUB OF CHEMISTRY
Solubility in Inorganic Chemistry
Like many other ionic compounds, sodium chloride dissolves easily in water. This raises a question: if
melting and dissolving both involve overcoming forces of attraction between ions, why does melting
occur at 801oC while dissolving can occur at room temperature?

When an ionic compound dissolves, the resulting free ions are surrounded by polar water molecules. In
solution, the positive sodium ion attracts the partial negative charge on the oxygen atoms of water.
Each negative chloride ion attracts the partial positive charges on the hydrogen atoms of the water
molecules.

If an ionic compound is soluble, the energy released because of the attractions between water
molecules and ions must be comparable with the energy required to separate the ions. This energy
requirement is not always met and several ionic compounds, such as magnesium and calcium
carbonate, dissolve only to a very small extent.

As water molecules have large dipoles, they can interact strongly with solute ions and release a lot of
energy. Non-polar solvents, such as hydrocarbons, do not dissolve ionic substances because much less
energy is released to compensate the positive lattice dissociation enthalpy.
GETTING TO THE NUB OF CHEMISTRY
Rate Equations 3.1.9
Rate is a measure of how quickly something changes. For example, speed is the rate of change of
distance.

𝑑𝑖𝑠𝑡𝑎𝑛𝑐𝑒
𝑠𝑝𝑒𝑒𝑑 =
𝑡𝑖𝑚𝑒

The rate of reaction is how fast reactants are converted into products.

1
𝐴𝑙𝑙 𝑟𝑎𝑡𝑒𝑠 ∝
𝑡𝑖𝑚𝑒

In 3.1.5 Kinetics we considered the factors that could affect the rate of a reaction along with why this
was through referring to the collision theory. These were:

• Temperature – increasing the temperature increased the energy of the particles along with the
frequency of collisions à an increase in the frequency of successful collisions.
• Concentration – increasing the concentration increases the frequency of successful collisions.
• Pressure – increasing the concentration increased the frequency of successful collisions.
• Catalysts – using a catalyst results in a lower activation energy and there’s a greater proportion
of successful collisions.
• Surface area – increasing the surface area increases the frequency of successful collisions.

Up until this point, we have only considered the rate of reactions in a qualitative sense. However, it’s
possible to collect data during the course of a reaction that allows us to plot a graph, which can then
provide us with quantitative information regarding the rate of a reaction.

The method we can use to collect data during a reaction (i.e. continuous monitoring) depends on the
reaction itself.

• Measure the volume, or mass, of a gas produced over time.

CaCO3(s) + 2HCl(aq) à CaCl2(aq) + H2O(l) + CO2(g)


GETTING TO THE NUB OF CHEMISTRY
• If a precipitate is formed (the solution goes cloudy – turbidity is the word that describes this
cloudiness), or there’s a change in colour, a colorimeter can be used to measure the absorbance of
light over time. In the following reaction, a precipitate of Sulfur is formed.

Na2S2O3(aq) + 2HCl(aq) → 2NaCl(aq) + H2O(l) + SO2(g) + S(s)

As the precipitate in the above reaction is formed, the absorbance of light increases. In the case of a
change of colour, absorbance is ∝ (proportional) to concentration. The actual concentration can be
found using a calibration curve prepared using known concentrations.

• Measure the concentration or a reactant or product (e.g. by titration) at specific times during the
reaction. This only works if the samples measured are quenched i.e. the reaction is, in effect,
stopped (in reality, it’s significantly slowed down). If there’s an acid catalyst, this can be done by
neutralising the acid. Otherwise, place the sample in an ice bath (reducing the temperature slows
the reaction right down) or dilute the reaction mixture (dilution reduces the concentration, slowing
the reaction right down).

Using the data collected via one of the methods above, we can plot a graph. The gradient at any point on
this graph will give us the rate of the reaction at that point. If the graph is a curve (they usually are), then
we will have to draw a tangent to the curve at that point.
GETTING TO THE NUB OF CHEMISTRY
The tangent when t=0 is the initial rate of the reaction.

Rate equations provide a mathematical description of the rate of reaction in terms of reactant
concentrations.

The general rate equation is given as follows:

Rate = k [A]m [B]n

• k is the rate constant for the reaction


• m and n are the orders of reaction w.r.t. (with respect to) the reactants A and B
• The overall order of the reaction is m + n
• The units for rate are mol dm-3 s-1

The power to which the concentration of A is raised in the experimental rate equation is called the order
of reaction with respect to A.

• Zero order [A]0 (= 1) The concentration of the reactant does not affect the rate of reaction
• First order [A]1 (= [A]) If the concentration doubles the rate of reaction will double
• Second order [A]2 (= [A] x [A]) If the concentration of the reactant doubles the rate of reaction
will be four times faster (22)

The overall order of the reaction is the sum of the powers the concentration terms are to.

Here’s an example. Let’s consider the following reaction:

BrO3- + 5Br- + 6H+ à 3Br2 + 3H2O

It’s determined experimentally that the rate equation for this reaction is:

Rate = k [BrO3 -] [Br -] [H+]

• The reaction is first order with respect to all three reactants.


• The overall order of the reaction = 3
GETTING TO THE NUB OF CHEMISTRY
The units for the rate constant depend on the rate equation. For example, let’s rearrange the rate
equation above so that its subject is k.

𝒓𝒂𝒕𝒆
𝒌=
[𝑩𝒓𝑶" " #
𝟑 ][𝑩𝒓 ][𝑯 ]

The units of rate are mol dm-3 s-1 and the units of concentration are mol dm-3. These allow us to work out
the units of k.

𝒎𝒐𝒍𝒅𝒎"𝟑 𝒔"𝟏
𝒖𝒏𝒊𝒕𝒔 𝒐𝒇 𝒌 =
(𝒎𝒐𝒍𝒅𝒎"𝟑 )(𝒎𝒐𝒍𝒅𝒎"𝟑 )(𝒎𝒐𝒍𝒅𝒎"𝟑 )

𝒎𝒐𝒍𝒅𝒎"𝟑 𝒔"𝟏
𝒖𝒏𝒊𝒕𝒔 𝒐𝒇 𝒌 =
(𝒎𝒐𝒍𝒅𝒎"𝟑 )(𝒎𝒐𝒍𝒅𝒎"𝟑 )(𝒎𝒐𝒍𝒅𝒎"𝟑 )

𝒔"𝟏
𝒖𝒏𝒊𝒕𝒔 𝒐𝒇 𝒌 =
𝒎𝒐𝒍𝟐 𝒅𝒎"𝟔

𝒖𝒏𝒊𝒕𝒔 𝒐𝒇 𝒌 = 𝒎𝒐𝒍"𝟐 𝒅𝒎𝟔 𝒔"𝟏

Rate equations can only be determined by experiment. The initial rates method is one way of doing this
and involves collecting and analysing experimental data. The initial rate of a reaction is the rate of the
reaction at the moment the reactants are first mixed.

The reaction can then be repeated with different concentrations and the different time taken recorded.
The concentration of one of the reactants is changed and the concentrations of all the other reactants is
kept constant. By comparing results, we can form a rate equation.
GETTING TO THE NUB OF CHEMISTRY
Let’s look at how we might determine the rate equation for a reaction. The following is an example of an
iodine clock reaction. Iodate ions react with sulfite ions in the presence of acid.

2IO3- + 5SO32- + 2H+ → I2 + 5SO42- + H2O

The reaction can be followed by using starch solution to detect the formation of iodine. It’s a clock
reaction because the colourless solution turns black very suddenly after a certain length of time. It’s this
time that’s measured and provides information on the initial rate of the reaction.

Here are our results from a series of experiments.

IO3-(aq) SO32-(aq) H+(aq) H2O(l) Time Rate /


Run
/cm3 /cm3 /cm3 /cm3 /s s-1

1 10 10 10 20 24 0.04

2 10 20 10 10 24 0.04

3 10 10 20 10 12 0.08

4 20 10 10 10 6 0.16

The total volume must be kept constant so that the concentration of each reactant is proportional to
volume.

Let’s start by considering how the rate of the reaction changes when we change the concentration of H+.
Comparing runs 1 and 3…

IO3- SO32-(aq) H+(aq) H2O(l) Time Rate /


Run (aq)
/cm3 /cm3 /cm3 /s s-1
/cm3
1 10 10 10 20 24 0.04
2 10 20 10 10 24 0.04
3 10 10 20 10 12 0.08
4 20 10 10 10 6 0.16
[H+] x 2 à Rate x 2
Rate µ [H+]
Reaction is first order with respect to [H+]
GETTING TO THE NUB OF CHEMISTRY
IO3- SO32-(aq) H+(aq) H2O(l) Time Rate /
Run (aq)
/cm3 /cm3 /cm3 /s s-1
/cm3
1 10 10 10 20 24 0.04
2 10 20 10 10 24 0.04
3 10 10 20
2 10 12 0.08
4 20 10 10 10 6 0.16
[IO3-] x 2 à Rate x 4 à Rate x 22
Rate µ [IO3-]2
Reaction is second order with respect to [IO3-]

IO3-(aq) SO32- H+(aq) H2O(l) Time Rate /


Run (aq)
/cm3 /cm3 /cm3 /s s-1
/cm3
1 10 10 10 20 24 0.04
2 10 20 10 10 24 0.04
3 10 10 20 10 12 0.08
4 20 10 10 10 6 0.16
[SO32-] x 2 à Rate is same
Rate µ [SO32-]0
Reaction is zero order with respect to [SO32-]

Pulling all this together, we get:

Rate µ [H+]1 [IO3-]2 [SO32-]0

As [SO32-] has no effect on the rate so it is omitted from the rate equation.

Rate = k[H+][IO3-]2

Notice that the rate equation bears no relation to the stoichiometric equation for the reaction:

2IO3- + 5SO32- + 2H+ → I2 + 5SO42- + H2O

Rate equations can only be determined experimentally.


GETTING TO THE NUB OF CHEMISTRY
Here’s another example. D and E react together to form F. Three experiments are performed with the
concentrations D and E changed the ensuing initial rate measured.

To begin with, our rate equation is:

𝑟𝑎𝑡𝑒 = 𝑘[𝐷]' [𝐸](

Begin by comparing two of the experiments in which the concentration of only one of the reactants is
different. In this case, experiments 2 and 3. Determine by what factor the concentration of E is changed
and by what factor the initial rate changes.

So, when the concentration of E was doubled, the initial rate quadrupled. This means the reaction is
second order with respect to E.

𝑟𝑎𝑡𝑒 = 𝑘[𝐷]' [𝐸])

Now we know how the concentration affects the rate of reaction, we can compare a different two
experiments to determine order with respect to D. Let’s choose 1 and 2.

We can see that the concentrations of both D and E are increased by a factor of 3/2 (try to work in
fractions – it’s easier) and the initial rate increases by a factor of 27/8 as a result. Let’s modify our
original rate equation so that it focuses on the change in rate.

𝑐ℎ𝑎𝑛𝑔𝑒 𝑖𝑛 𝑟𝑎𝑡𝑒 = (𝑐ℎ𝑎𝑛𝑔𝑒 𝑖𝑛 𝑐𝑜𝑛𝑐. 𝑜𝑓 𝐷)' (𝑐ℎ𝑎𝑛𝑔𝑒 𝑖𝑛 𝑐𝑜𝑛𝑐. 𝑜𝑓 𝐸))


GETTING TO THE NUB OF CHEMISTRY
Substitute in the changes we determined.

27 3 ' 3 )
=R T R T
8 2 2

27 3 ' 9
=R T R T
8 2 4

3 3 '
=R T
2 2

So, m = 1 and the reaction is first order with respect to D. So, the rate equation for this reaction is:

𝑟𝑎𝑡𝑒 = 𝑘[𝐷][𝐸])

Let’s look back at our first rate equation example, which was for the reaction:

2IO3- + 5SO32- + 2H+ → I2 + 5SO42- + H2O

Rate = k[H+][IO3-]2

Notice that concentration of one of the reactants (SO32-) does not affect the rate of the reaction. This is a
clear modification of the factors we learned at GCSE and in the AS level content that affect the rate of
reaction. Concentration does not always affect rate! This obviously raises the question, why not?

Imagine I have prepared ten worksheets and want to collect them into a complete set and ask for help.
The worksheets are arranged in ten piles and…

• Mani collects a page from each of the piles (step 1).


• Miayla takes the set of ten pages and tidies them ready for stapling (step 2).
• Matthew staples the set of notes together (step 3).

A fourth person then volunteers. Who should I put them with so that the complete set gets made faster?
Or, in other words, which step is the one that is determining how quickly the sets are compiled?

Step 1 in the process above is the one that determines how quickly the complete sets are made as it’s
clearly the slowest step of the three.
GETTING TO THE NUB OF CHEMISTRY
As you will have begun to appreciate from the organic mechanisms you have done so far, chemical
reactions can take place over a series of steps. This can be true or inorganic reactions too.

The overall rate of a reaction (the one which you would measure if you did some experiments) is
controlled by the rate of the slowest step.

The slow step of a reaction is known as the rate determining step (RDS).

Crucially

• In any mechanism, species that are only involved in steps after the RDS do not appear in the rate
equation.

• However, species that appear in the steps up to, and including, the RDS are in the rate equation.

We find the rate equation by experiment and then use the rate equation to deduce possible
mechanisms. A substance that acts as a catalyst will appear in the rate equation but not in the overall
chemical equation for the reaction (it hasn’t been used up so would appear on both sides of the
equation and therefore cancel out).

Let’s consider the reaction:

CH3COCH3(aq) + I2(aq) à CH3COCH2I(aq) + HI(aq)

The reaction is acid catalysed and, using a colorimeter, the following rate equation is determined:

Rate = [CH3COCH3(aq)][H+(aq)]

The reaction proceeds through the following steps:

As the iodine molecules are not involved in this proposed mechanism until after the rate-determining
step, the overall rate of the reaction would not depend on the concentration of iodine.
GETTING TO THE NUB OF CHEMISTRY
The Isolation Method would have also allowed us to determine how the rate of reaction is affected by
the concentration of one of the reactants. The dependence of the reaction rate on the chosen reactant
concentration is isolated by having all other reactants present in a large excess, so that their
concentration remains essentially constant throughout the course of the reaction.

Returning to our original reaction:

2IO3- + 5SO32- + 2H+ → I2 + 5SO42- + H2O

The rate equation will be: Rate = k[H+]a [IO3-]b [SO32-]c

If we want to know how the concentration of H+ affects the rate of the reaction, we can isolate this by
ensuring IO3- and SO32- are present at a concentration 1000 times greater than H+. When all of the H+ has
been used up, the concentrations of IO3- and SO32- will only have changed by 1/1000, or 0.1%, and so
99.9% of the original IO3- and SO32- will still be present. It is therefore a good approximation to treat its
concentration as constant throughout the reaction.

This greatly simplifies the rate law since the (constant) concentrations of all reactants present in
large excess may be combined with the rate constant to yield a single effective rate constant. For
example, the rate law for the reaction considered above will have become:

Rate = k1[H+]a

Where k1 = k[IO3-]b [SO32-]c from the original rate equation.


GETTING TO THE NUB OF CHEMISTRY
The rate of reaction can be expressed graphically.

• concentration vs. time


• rate vs. concentration

The shapes of these graphs can tell us the order of reaction with respect to the reactant.

By drawing tangents at different concentrations, we calculate the rate at different points in the reaction
and plot a corresponding graph.

When a reaction is second order with respect to a substance, then:

Rate µ [A]2

Therefore, there is a third graph we can draw:


GETTING TO THE NUB OF CHEMISTRY
As we increase the temperature a larger proportion of particles have energy equal to, or greater than,
the activation energy. So, as we learnt back at GCSE, if you increase the temperature, you increase the
rate. It begs the question of where, in our work on rate equations, does temperature feature?

Rate = k [A]m [B]n

The link between that information and our quantitative work on rates lies in our understanding the
Arrhenius equation.
"
! !
𝑘 = 𝐴𝑒 #$

k – rate constant (units depend on rate of equation but will always include s-1)
e – irrational number (like π)
T – temperature (in kelvins)
A – pre-exponential factor, a constant for each chemical reaction (units will be the same as k’s)
Ea – activation energy (J mol-1)
R – gas constant (J K-1 mol-1)

"!
𝑘 = 𝐴𝑒 !#$

• If we raise the value of T, the power to which Ae is raised would get smaller
• As the power is negative, the smaller the power the larger Ae becomes and hence k
• Decreasing the activation energy (by adding a catalyst) also has the same effect and increases k
• Raising the temperature increases the value of the rate constant and hence the rate

Here’s an example calculation:

The rate constant, k, for a reaction varies with temperature as shown by the equation

*!
𝑘 = 𝐴𝑒 "+,


For this reaction, at 25 °C, k = 3.46 × 10–8 s–1

The activation energy Ea = 96.2 kJ mol–1

The gas constant R = 8.31 J K–1 mol–1

Calculate a value for the Arrhenius constant, A, for this reaction. Give the units for A.

Always rearrange the equation before substituting in any values (watch out of the units of Ea and T).

𝑘
𝐴 = *!
𝑒 "+,

3.46 × 10"- 𝑠 "3


𝐴 = ./)00 = 2.57 𝑥 10. 6'78 "#
= 𝑠 "3
" "
𝑒 -.23 5 ).-
𝑒 69 "# '78 "# 5 9
GETTING TO THE NUB OF CHEMISTRY
It’s possible to determine the activation energy of a reaction from this equation and plotting a suitable
graph.

"!
𝑘 = 𝐴𝑒 !#$

Natural log (ln) both side of the equation (you don’t need to do this but it’s mentioned to explain what’s
happening).
−𝑬𝒂
𝒍𝒏 𝒌 = + 𝒍𝒏 𝑨
𝑹𝑻
−𝑬𝒂 𝟏
𝒍𝒏 𝒌 = + 𝒍𝒏 𝑨
𝑹 𝑻
y = mx + c
𝟏
Plot ln k vs. 𝑻
𝟏
y = ln k, x = 𝑻
"𝑬𝒂
m (gradient of the line) = 𝑹

EA= -mR
Do not write
6 outside the
2-
box
Sy0c
25202-
0 1 . 8 A general equation for a reaction is shown.=
1202 2u + 2 2420()
Fz(aq)
+

odineina
-

I
+ +

A(aq) + B(aq) + C(aq) → D(aq) + E(aq)


thisie In aqueous solution, A, B, C and D are all colourless but E is dark blue.
-
initial rates S20g
version A reagent (X) is available that reacts rapidly with E. This means that, if a small
of amount of X is included in the initial reaction mixture, it will react with any E produced
required until all of the X has been used up.
practical
7
.
Explain, giving brief experimental details, how you could use a series of experiments
to determine the order of this reaction with respect to A. In each experiment you
should obtain a measure of the initial rate of reaction.
- REMINDS
USE BULLET POINTS /NUMBERS YOU [6 marks]
THIS IS A METHOD ; REDUCES WAFFLE

.
1
Using a measuring
cylinder ,
measure 10cm

of A
,
B
,
C and X .
Place the volumes of

B C and X in a conical flask .

.
2 Place the flask on a white tile will a

black cross in a water such that the

cross is visible
through the bottom of the

flask .

the timer
.
3 Add A to the flask and start

4 .
Time how lo
ng
itlakes
for the crossno
disappear.
1-4 volumes
.
5
Repeat steps using the same

and concentrations of B
,
C and X and

different concentrations
of A
(same volume)
6 .
Time is a measure of the rate of reaction(s)
Plot time concentration of A
graph of
7 . a us .
.

.
8 If the resulting graph is :

El then the rate order A


is zero writ .

[A]

E men the rate is


1st order .
writ A
[A]

E then the rate is 200 order w r .


. t . .
A

[A]

*06*
IB/M/Jun18/7405/3
GETTING TO THE NUB OF CHEMISTRY
Equilibrium constant Kp for homogeneous systems 3.1.10
Homogenous: all components are in the same phase (less precisely, the same state)

To recap some aspects of Le Chatelier’s principle:

• Le Chatelier's principle is that the position of equilibrium will move (to the left or right) to
counteract any changes that are made to the system

• It helps chemists to predict the outcome of any changes made to the pressure, concentration or
temperature

When an equilibrium involves gases, it can be expressed in terms of the pressure exerted by the different
gases of the reactants and products rather than concentration. It depends on the data you have available
as to whether you use Kc or Kp.

Kp is calculated using the partial pressures of the reactants and products rather than their
concentrations. The partial pressure of a gas in a mixture of gases is the pressure it would exert if it alone
occupied the total volume occupied by the gas mixture.

partial pressure of A partial pressure of B Total pressure = pA + pB


pA pB

The equilibrium constant can be expressed in terms of the individual partial pressures and is given the
symbol Kp.

A(g) + 2B(g) ⇌ C(g) + D(g)

(𝑝𝐶)(𝑝𝐷)
𝐾! =
(𝑝𝐴)(𝑝𝐵)"

So, for the reaction 3H2(g) + N2(g) ⇌ 2NH3(g)

(𝑝𝑁𝐻# )"
𝐾! =
(𝑝𝐻" )# (𝑝𝑁" )
GETTING TO THE NUB OF CHEMISTRY
Sometimes the partial pressures will be given to you. If they’re not, you will need to calculate them using
mole fractions, which are moles (n) of one chemical divided by the total moles in the mixture. For
example, in the reaction:

A+B⇌C+D

𝑛𝐴
The mole fraction of A =
𝑛𝐴 + 𝑛𝐵 + 𝑛𝐶 + 𝑛𝐷

Here’s a concrete example – imagine you have a mixture of gases at equilibrium:

Nitrogen 20mol, Hydrogen 45mol, Ammonia 25mol

Total number of moles is 20+45+25 = 90.

"$
This means the mole fraction of nitrogen is = = 0.22
%$

All the mole fractions added together must equal 1.

We can now put all these ideas together: Calculate Kp of a mixture of Nitrogen (20 moles), Hydrogen (45
moles) and Ammonia (25 moles) at a pressure of 40 atm.

N2 H2 NH3

Moles 20 45 25

Mole fraction 0.22 0.50 0.28

Partial pressure (mole


fraction x total 8.89 20.00 11.11
pressure (40))

(𝑝𝑁𝐻# )"
𝐾! =
(𝑝𝐻" )# (𝑝𝑁" )

(11.11)"
𝐾! = = 0.00174
(20.00)# (8.89)

The units will be in atm or kPa and are determined in the same way as they are for Kc

(𝑎𝑡𝑚)"
𝑈𝑛𝑖𝑡𝑠 𝑜𝑓 𝐾! =
(𝑎𝑡𝑚)# (𝑎𝑡𝑚)

𝐾! = 𝟎. 𝟎𝟎𝟏𝟕𝟒 𝐚𝐭𝐦&𝟐
GETTING TO THE NUB OF CHEMISTRY
As with Kc:

• Changing the pressure change the position of equilibrium but does not change the value of Kp.
• Adding a catalyst, speeds up the time it takes to reach equilibrium but does not change the
position of equilibrium and does not change the value of Kp.
• Only changing the temperature changes the value of Kp.
GETTING TO THE NUB OF CHEMISTRY
GETTING TO THE NUB OF CHEMISTRY
3.1.11.1 Electrode Potential and Cells
This sounds rather niche. It isn’t. You can’t just leave it at that. Show me it isn’t niche. Okay, have you
ever owned a lemon clock? Niche! Sorry, are you just misspelling ‘nicht’ there? Are you saying you have
never owned a lemon clock? Nicht! Errm, I think I’ll just carry on… Here’s a photo of a lemon clock.

I want you to notice the fact that two different metals are stuck into the lemon – that’s very important
and completely relevant to this topic because if the metals strips were made of the same metal the clock
wouldn’t work.

Okay, I’m intrigued. Great. I’ll start at the beginning then. If you place a piece of metal foil in a solution
of its ions, an equilibrium is set up:

M(s) ⇌ Mz+(aq) + ze-

The position of equilibrium could lie to the left, to the right or be somewhere in the middle. Now, it’s
crucial to understand where the ze- are. They’re not in the solution like the positive ions. They’re
delocalised on the metal. Can you see what this means?

I’m grasping for it… The metal could become negatively charged due to electrons building up on it as
metal cations are released into the solution with the latter becoming positively charged because of the
excess of cations. Alternatively, metal ions in the solution could take up delocalized electrons in the strip
of metal and be discharged as metal atoms (increasing the mass of the foil). In this case, the metal strip
would become positively charged. Whichever happens, the result would be a difference in charge
between the metal and the solution.

If there’s a difference in charge, why do we need two different metals for our lemon clock? If we think
back to the reactivity series, we can view this as essentially a relative measure of how easily various
metals formed positive ions. Metals at the top, such as potassium and sodium, form ions readily and
those at the bottom, like gold and platinum don’t. So, if we put foil pieces of two different metals e.g.
magnesium and copper in solutions containing their respective ions, then we would expect the following
equilibriums to be established:

Mg(s) ⇌ Mg2+(aq) + 2e-


Cu(s) ⇌ Cu2+(aq) + 2e-

And our knowledge of the reactivity series suggests to us that the position of their respective
equilibriums would be different. As magnesium is higher in the reactivity series than copper, which as we
know means it forms ions more readily, we would expect its position of equilibrium to lie further to the
right and copper’s further to the left.
Mg(s) ⇌ Mg2+(aq) + 2e-
Cu(s) ⇌ Cu2+(aq) + 2e-
Which means…
GETTING TO THE NUB OF CHEMISTRY
The differences in charge set up in each equilibrium would be different and if these are different then
we have… a… potential difference. Exactly. This is exactly how Alessandro Volta created the first cell,
consisted of alternating zinc and silver discs, with brine-soaked cardboard between each disc and wires
connected at both ends. In this A level topic, electrode potentials are an attempt to quantify the
differences in the position of equilibria.

Great, let’s get started then. Whoa, cool your jets! The thing is, it’s not possible to measure the electric
potential between the metal and its ions in solution. WHAT???? You said this topic was all about
quantifying. Surely that means we have to measure something. We do, we do. The point is, we have to
measure these potentials against a reference just like we do with anything else. For example, a metre
was originally defined as one ten-millionth of the distance from the equator to the North Pole and is now
the distance travelled by light in a vacuum in 1299,792,458 of a second. We measure lengths against this
reference.

Oh, I see. What reference do we use for electrode potentials? Something called the Standard Hydrogen
Electrode (SHE). This has been assigned the potential of 0 and everything is measured against this. In
addition, just we saw in previous topics, measurements are made under standard conditions. We need
to know what these are before we look more closely at the SHE.

Fair enough. So, what are these standard conditions? No, wait, let me guess whatever you’re about to
write.

• Gases are at a pressure of 100 kPa


• The temperature is 298K (25oC)
• All ions present are at a concentration of 1 mol dm-3

Crikey, spot on! Finally, it’s very important that no charge flows (i.e. the current is 0 A) when an
electrode potential is measured. This is so the maximum voltage (potential difference) possible is
measured – this is called the electromotive force (e.m.f.) and a high resistance voltmeter is put in series
to prevent any charge flowing. How come? Remember V = IR from your GCSE Physics? Yuck! Well, given
‘V’ is essentially determined by the metal and solution, a very large ‘R’ means a very small ‘I’. This is
important because if charge is allowed to flow then the positions of equilibrium in our half-equations
change (we’ll look at this more later on).

So, can I hear about the Standard Hydrogen Electrode now? Sure, here’s a diagram.
GETTING TO THE NUB OF CHEMISTRY
You can see it’s under the standard conditions. The relevant equilibrium is:

H2(g) ⇌ 2H+(aq) + 2e-

The SHE is a half-cell we use to measure another half-cell e.g. copper in 1 mol dm-3 copper sulfate.
Combining two half-cells together produces an electrochemical cell that can allow charge to flow
because of the difference in their respective electrode potentials (however, they can’t actually do this
because of the high resistance voltmeter). When two half cells are joined together, a salt bridge is used
to complete the circuit. It does this by allowing ions to move from one half-cell to the other while
ensuring the solutions to remain separate. This is often a tube containing a saturated solution of KNO3 in
agar gel although it can also be filter paper soaked in this solution. A combination of a Group 1 metal and
nitrate ions ensures we have no possibility of precipitation (they’re always soluble), which would affect
out results.

Tell me more about half-cells. Okay, so it’s clear, a half-cell comprises of an element in two different
oxidation states. There are different types of half cells:

1. Metal / Metal ion half cells (e.g. Zn2+(aq) + 2e- ⇌ Zn(s))


2. Non-metal / non-metal ion half cells (e.g. Br2(g) + 2e- ⇌ 2Br-(aq))
3. Metal ion / Metal ion half cells (e.g. Fe3+(aq) + e- ⇌ Fe2+(aq))

Metal / Metal ion half cells

This is when the metal element, which has an oxidation state of 0, is placed in an aqueous solution of its
ions. In this type of half-cell an equilibrium exists between the metal element and metal ion.

Zn2+(aq) + 2e- ⇌ Zn(s)

Forward reaction = reduction


Backward reaction = oxidation

Hold on a minute, you’ve switched the equilibria around from how we wrote them earlier. Well
spotted. From now on we’re going to present them in this way. This is the convention when it comes to
electrode potentials and this is how they will always be presented in questions. Everything we said
earlier still applies. It’s still about how far to the right or left the position of equilibrium lies.
GETTING TO THE NUB OF CHEMISTRY
Hmmm. Okay. Tell me about the second type of half-cell then. So, non-metals can also be in equilibrium
with its non-metal ions. For example, a half cell of bromine and its ions:

Br2(g) + 2e- ⇌ 2Br-(aq)

Of course, the most common type of non-metal/non-metal ion half-cell is the hydrogen half-cell.

2H+(aq) + 2e- ⇌ H2(g)

You’ll notice that there is no solid piece of metal to act as an electrode. To overcome this issue, an inert
(unreactive) platinum electrode is used instead. The platinum electrode allows electrons to flow (it’s a
good conductor of electricity after all) between the non-metal (e.g. hydrogen) and its ions (e.g. hydrogen
ions). The fact it’s inert prevents side reactions from occurring.

And the final type of half-cell? A metal ion/metal ion half-cell contains ions of the same element in
different oxidation states. For example, a half cell can contain Fe3+(aq) and Fe2+(aq) ions:

Fe3+(aq) + e- ⇌ Fe2+(aq)

Similar, to the hydrogen half-cell, since there is no solid piece of metal, a platinum electrode is needed to
act as a point of contact between the two metal ions for electrons to flow. To things clear regarding
these cells and standard conditions, a standard Fe3+(aq)/Fe2+(aq) half-cell is made up a solution containing
Fe2+(aq) and Fe3+(aq) ions with the same 1.0M concentrations of each ion (equimolar). You need to be
careful here as this is concentration ion, which isn’t necessarily the same as the concentration of the
solution e.g. 0.5M Fe2(SO4)3 = 1.0M of Fe3+.

Okay, so we’ve got our various half-cells and you’ve mentioned that two half-cells make an
electrochemical cell in which electrons can flow (although they don’t because of the high resistance
voltemeter). So, tell me more about that. I’d like nothing better. In a simple electrochemical cell, a
redox reaction takes place (in some courses, this topic is referred to as redox equilibria). For example:

2Ag+(aq) + Cu(s) à 2Ag(s) + Cu2+(aq)

If a redox reaction takes place, then two half equations must be involved

Cu(s) à Cu2+(aq) + 2e-


2Ag+(aq) + 2e- à 2Ag(s)

Rather than drawing complicated diagrams of electrochemical cells, a shorthand form is written:

Pt(s)| H2(g) | H+(aq) || Zn2+(aq) | Zn(s)

• | represents a phase boundary (i.e. between species in different states). Note: (aq) and (l) are in
the same phase)
• || represents a salt bridge
• Note that the actual solid electrodes are written at the two ends (there must always be a solid
electrode – if an equilibrium doesn’t possess a solid, then a platinum electrode is used)
GETTING TO THE NUB OF CHEMISTRY
The most oxidised species are near the middle (ROOR – Reduced, Oxidised, Oxidised, Reduced). Here are
some more examples (ROOR is placed about the relevant species):

R O O R
3+ 2+
Al(s) | Al (aq) || Pb (aq) | Pb(s)

R O O R
Fe(s) | Fe2+(aq) || MnO4–(aq), H+(aq), Mn2+(aq)| Pt(s)

R O O R
Pt(s)| H2(g) | OH (aq) || O2(g) |OH-(aq)| Pt(s)
-

The convention is that the half-cell with the more positive potential is set up as the right-hand
electrode* so cell diagrams are drawn this way (*the exception is that the SHE is always the left-hand
electrode).

Okay, so we’ve measured the electrode potentials of a half-cell. Now what? Well, if we do this for a
number of half-cells, we end up with the electrochemical series, which looks like this.
GETTING TO THE NUB OF CHEMISTRY
Do you notice anything? Hmm, it looks a lot like the reactivity series I did at GCSE only with non-metals
as well as metals. That’s right. Really, it’s just another version of the reactivity series and just as the
reactivity series we learned at GCSE allowed us to predict if a redox reaction would happen or not (we
incorrectly referred to them as displacement reactions), so does the electrochemical series.

Ah! I think things are becoming a little clearer. Phew! So, how exactly can we use the series? Well look
at the electrode potentials listed. If we put together two half-cells, the more positive half-cell will attract
the electrons. This means electrons are gained at the positive half-cell à reduction; the equilibria
equation will run in the forward direction. Conversely, electrons are lost at the more negative half-cell
à oxidation.
Reduction à
Cu (aq) + 2e- ⇌ Cu(s)
2+
+0.34V
2+ -
Zn (aq) + 2e ⇌ Zn(s) -0.76V
ß Oxidation

Notice, if you’re given the simple cell diagram, then you’ve been given you overall reaction.
GETTING TO THE NUB OF CHEMISTRY
In summary…

Is there anything in particular I should know about how AQA does things? I’m glad you asked. AQA like
to use the following notation to represent the half-cells.

Notice, it’s the left-hand, oxidized species, over the right-hand, reduced species. We can use this to
determine whether a particular reaction takes place and what the equation would be.

What if a question asks you to explain why the reaction takes place… You say: this happens as E°
(Cu2+/Cu) is more positive than (or ‘>’) E°(Zn2+/Zn) and so Cu2+ gains electrons (it’s always LH species/RH
species). In effect, Cu2+ oxidises Zn and Zn reduces Cu2+.

Equally, they can ask why a reaction DOESN’T take place e.g why aren’t Zn2+ ions capable of oxidizing Cu?
Our answer is the same as above: E° (Cu2+/Cu) is more positive than (or ‘>’) E°(Zn2+/Zn)
GETTING TO THE NUB OF CHEMISTRY
For example…

Here’s another example. Based on the electrode potentials below, will H2O2(aq) react with itself resulting
in disproportionation?

Remember, electrochemical half-cells are redox half-equations. Often, it’s obvious what’s changing
oxidation number. When it isn’t, try to make it so.
GETTING TO THE NUB OF CHEMISTRY

At the more positive half-cell, we have reduction. Oxidation takes place at the more negative one.

What about actual calculations? Depending on the info you’re give, the Eѳcell or e.m.f. can be calculated
as follows...

Eѳcell = Eѳright – Eѳleft


Eѳcell = Eѳmore positive – Eѳmore negative

Cu2+(aq) + 2e- ⇌ Cu(s) +0.34V


Zn2+(aq) + 2e- ⇌ Zn(s) -0.76V

Eѳcell = +0.34 – (-0.76) = +1.10V

Another example?

Pt(s)| Fe2+(aq), Fe3+(aq) || Tl3+(aq), Tl+(aq)| Pt(s) Cell e.m.f. = +0.48V

Eѳright = e.m.f. + Eѳleft


Eѳright = +0.48 + 0.77
Eѳright = 1.25V
GETTING TO THE NUB OF CHEMISTRY
Going back to the electrochemical series, I’m sure I’ve seen exam questions that ask what the
best/weakest oxidizing/reducing agent is.

Remember, an oxidizing agent is itself reduced (it gains electrons) and a reducing agent is itself oxidized
(it loses electrons). So, for have the potential to be an oxidizing agent, it must be a left-hand species. The
PIONIR acronym tells you that the positive electrode potential contains the most powerful oxidizing
agent the most negative the most powerful reducing agent. So, based on the table above…

• Most powerful oxidizing agent: Cl2


• Most powerful reducing agent: Fe
• Weakest oxidizing agent: Fe2+
• Weakest reducing agent: Cl-

We can link this with 3.2.3 Group 7 (17), the Halogens. In group 7:

• We determined the oxidizing power of the halogens to be F2 > Cl2 > Br2 > I2
• We determined the reducing power of the halide ions to be I- > Br- > Cl- > F- using their reaction
with conc. H2SO4.

These results tally with the electrochemical series, part of which is reproduced below.
GETTING TO THE NUB OF CHEMISTRY
Near the beginning, you mentioned that we’d consider what happened if we did allow charge to flow.
Well remembered but we’ll consider another possibility prior to look as what happens when charge is
allowed to flow. Imagine the scenario below:

According to the electrochemical series, the right-hand cell would have an electrode potential of 0.34V.
However, the one on the left would be different as it’s not under standard conditions (its concentration
is not 1.0 mol dm-3). We can use the equilibrium equation to determine whether it would more positive
or more negative:
Cu2+(aq) + 2e- ⇌ Cu(s)

0.15 mol dm-3 < 1.0 mol dm-3. The equilibrium constant must remain… constant… so the equilibrium will
further to the left to ensure this is the case. This means there more electrons released so the potential
difference become more negative. We learned earlier that the right-hand cell is the more positive, which
confirms this. So the diagram is correctly drawn.

Perhaps weirdly, the conventional representation for this cell would be:

Cu(s)| Cu2+(aq) || Cu2+(aq) | Cu(s)

And allowing the charge to flow? Of course. Imagine we now replace the voltmeter with a bulb. Thinking
about V = IR, R is now much smaller so I will be much bigger – in effect, we now have a current, or flow
of charge. Electrons are now leaving the left-hand cell and flowing into the right-hand cell. We no longer
have a closed system.

Cu2+(aq) + 2e- ⇌ Cu(s)

• In the left-hand cell, there is a shift to the left because of the removal of electrons and so the
concentration of Cu2+(aq) ions increases.

• In the right-hand cell, there is a shift to the right because of the addition of electrons and so the
concentration of Cu2+(aq) ions decreases.

From the moment charge is allowed to flow, the e.m.f. of the cell decreases until it eventually reaches
0V. This happens when the concentration of Cu2+(aq) in the two half-cells is equal.
GETTING TO THE NUB OF CHEMISTRY
Phew! This has been mentally exhausting. There’s just one more thing I want us to look at. And that is?
The oxidation of Vanadium (II) ions. You’ll formally cover this when you do Transition Metals but there’s
a strong overlap with Electrode Potentials. Below are the colours of the various oxidation states of
Vanadium.

V2+ V3+ VO2+ VO2+


GETTING TO THE NUB OF CHEMISTRY
3.1.11.2 Commercial applications of electrochemical cells
Electrochemical cells can be used as a commercial source of electrical energy. Cells can be non-
rechargeable (irreversible), rechargeable or fuel cells.

Alkaline cells are very common and can produce a steady current over a longer time than many other
cells. Here is an example.

This is its cell diagram.

Zn(s) | Zn2+(aq) || MnO2(s) + H2O(l), Mn2O3(s) + 2OH–(aq)| C(graphite)

At the zinc anode, we get this half equation:

Zn(s) + 2OH–(aq) à ZnO(s) + H2O(l) + 2e–

At the positive carbon (graphite) electrode, the reaction is:

2MnO2(s) + H2O(l) + 2e– à Mn2O3(s) + 2OH–(aq)

Adding these two half-equations together gives the overall cell reaction:

Zn(s) + 2MnO2(s) à ZnO(s) + Mn2O3(s)


GETTING TO THE NUB OF CHEMISTRY
Rechargeable Cells

Lithium cells provide a steady voltage for the life of the cell (you need to know these equations).

At the positive electrode:

Li+ + CoO2 + e– → Li+[CoO2]–

At the negative electrode:

Li → Li+ + e–

So, the overall reaction is:

Li + CoO2 → LiCoO2

When a cell is recharged, an electric current is passed through it in the opposite direction to that which
flows when the cell is generating electricity. The original substances reform. This recharging is an
example of electrolysis and is the reverse of discharge.

LiCoO2 → Li + CoO2

In general terms, it’s…

Here’s how we consider things should we be given the relevant information as part of the
electrochemical series.

Rechargeable cells have the environmental advantage over non-rechargeable cells by virtue of the
metals, a finite resource, being reused. However, the electricity required to recharge the cell may have
been produced by the burning of fossil fuels.
GETTING TO THE NUB OF CHEMISTRY
Here’s an example question:

Fuel cells are electrochemical cells. They differ from the commercial cells above by having a continuous
supply of reactants. The allows the voltage to stay constant. They have an efficiency of approximately
70%, which is significantly better than the best combustion engines’ approximate 45% efficiency. Below
is a diagram demonstrating the principles of an alkaline hydrogen–oxygen fuel cell.
GETTING TO THE NUB OF CHEMISTRY
Acidic hydrogen–oxygen fuel cells also exist. Here’s the relevant electrochemical series.

Electrode half-equation Eθ / V
H2 + 2OH– → 2H2O + 2e– -0.83V
H2 → 2H+ + 2e– 0.00V
O2 + 4e– + 2H2O → 4OH– +0.40V
O2 + 4H+ + 4e– → 2H2O +1.23V

The overall equation for both alkaline and acidic hydrogen–oxygen fuel cells is the same, as is the
voltage.

2H2 + O2 → 2H2O E = +1.23V

You can memorise the half-equations above or you can construct them from scratch.

In acidic conditions, you know H2 à and O2 à and you then finish them using the rules you learned in
y12 i.e. balance the element oxidized/reduced, add H2O to balance oxygens, add H+ to balance
hydrogens and electrons to balance the charge.

So, the cell diagram is written as follows:

In alkaline conditions, you work it out exactly as above before adding enough hydroxide ions to both
sides of the equation so that you can neutralise all the hydrogen ions. Because it is now a balanced
equation, you must add the same number to both sides to maintain the balance. The hydrogen ions and
hydroxide ions on the same side would turn into water molecules.
GETTING TO THE NUB OF CHEMISTRY
There are inevitably benefits and risks to society associated with using these cells.

Benefit(s) Risk(s)
Cells • Portable store of energy • waste issues
Non-rechargeable cells • cheap • waste issues
Rechargeable cells • less waste • some waste issues (at end of
• cheaper in the long run useful life)
• lower environmental impact • energy needed for
recharging
Hydrogen fuel cells • only waste product is water • need a constant supply of
• do not need re-charging fuels
• very efficient • hydrogen is flammable &
explosive
• hydrogen usually made using
fossil fuels
• high cost of fuel cells
GETTING TO THE NUB OF CHEMISTRY
GETTING TO THE NUB OF CHEMISTRY
Acids and bases 3.1.12
3.1.12.1 Brønsted–Lowry acid–base equilibria in aqueous solution

The Brønsted–Lowry acid–base theory states that:

• An acid is a proton donor


• A base is a proton acceptor

However, a proton, H+, cannot exist independently in aqueous solution.

HCl (aq) + H2O (l) à H3O+(aq) + Cl-(aq)

H3O+ is the hydroxonium ion.

However, when writing equations, the less precise H+ is often used for simplicity.

In water, HCl behaves as an acid – it donates protons. However, if it is dissolved in methylbenzene it will
not show any acidic properties. This is because methylbenzene will not accept protons.

An interesting aspect of The Brønsted–Lowry acid–base theory is that once an acid has ‘donated’ a
proton it would become able to ‘accept’ a proton back and hence act as a base. We call these pairs of
chemicals conjugate acid-base pairs.
Conjugate pair

HCl (aq) + H2O (l) ⇌ H3O+(aq) + Cl-(aq)


Conjugate Conjugate
Acid Base
Acid Base

Conjugate pair
Another example:

CH3COOH (aq) + NH3 (aq) ⇌ NH4+(aq) + CH3COO-(aq)


Conjugate Conjugate
Acid Base
Acid Base
GETTING TO THE NUB OF CHEMISTRY
Water can act as an acid and a base – it can be described as amphoteric.

Pure water will always contain some H+ and OH- ions it is described as self-protonating.

3.1.12.2 Definition and determination of pH

We could use [H+] as a measure of acidity, but the numbers would be difficult to work with. To simplify
things, we use a log scale:

pH = -log10[H+]

[H+] pH
(mol dm-3) (ALWAYS give pH to two decimal places)
0.00500 2.30
0.0250 1.60
0.125 0.90

As [H+] increases pH decreases.

We need to be mindful of the stoichiometry when acids dissociate. Hydrochloric and nitric acid for
example dissociate to give one hydrogen ion (one proton). These are monoprotic acids.

HCl à H+ + Cl-

However, sulfuric acid and phosphoric acid both split up to give more than one proton.

H2SO4 à 2H+ + SO42- (diprotic)


H3PO4 à 3H+ + PO43- (triprotic)

This is a simple mistake to make in the heat of an exam!

Calculating [H+] from pH.

[H+] = 10-pH

pH [H+] (mol dm-3)


-0.50 3.16
2.24 5.75 x 10-3
6.95 1.12 x 10-7
GETTING TO THE NUB OF CHEMISTRY
3.1.12.3 The ionic product of water, KW

Just because a solution contains H+ doesn’t necessarily mean it’s acidic. All aqueous solutions will contain
H+ and OH- ions.

Remember, water is self-protonating:

2H2O (l) ⇌ H3O+(aq) + OH-(aq)


or, as it’s commonly written
H2O (l) ⇌ H+(aq) + OH-(aq)

You can now see that adding an alkali increases the pH by the shifting the equilibrium to the left, thereby
lowering [H+].

The equilibrium expression for the above is:

[" ! ][$" " ]


Kc =
["# $]

Let’s consider the concentration of water in pure water. Let’s make life easy and consider 1dm3 of water.

The density of water is 1g/cm3.

1dm3 = 1000cm3

So, the mass of water = 1000cm3 x 1g/cm3.

The Mr of H2O is 18.

Therefore, 1dm3 contains 1000/18 moles = 55.6. Therefore, the concentration of water is 55.6 mol dm-3

This means that the denominator of our Kc expression above is so much, much bigger when compared to
the numerator that, to all intents and purposes, [H2O] remains constant. This allows us to simplify the
original expression, resulting in the ionic product of water Kw.

Kw = [H+][OH-]

At 298K the value of Kw is 1 x 10-14 mol2dm-6

In pure water [H+] = [OH-]. So, Kw = [H+]2. Hence [H+] = √Kw

At 298K [H+] = √(1 x 10-14mol2 dm-6) = 1 x 10-7 mol dm-3. So, at 298K the pH of pure water is 7.00
GETTING TO THE NUB OF CHEMISTRY

Increasing the temp à shifts the eqm to the right à leading to an increase in Kw. It therefore also means
an increase in [H+]. So, at higher temperatures, the pH of water will decrease. However, the water is still
neutral as [H+] = [OH-].

Calculate the pH of water at 40°C when Kw = 2.09 x 10-14 mol2 dm-6

Kw gives us a method of calculating the pH of strong bases.

Kw = [H+][OH-]

𝐾%
[𝐻$ ] =
[𝑂𝐻– ]

Calculate the pH of 0.1M KOH at 298K. Since KOH is a strong base [OH-] = 0.1M

Kw = [H+][OH-] = 1 x 10-14 mol2dm-6

𝐾%
[𝐻$ ] =
[𝑂𝐻– ]

1 𝑥 10'()
[𝐻$ ] = = 1 𝑥 10'(*
0.1

pH = -log10(1 𝑥 10'(* )

pH = 13.00

Watch out for bases that release 2 x OH- ion e.g. Ba(OH)2
GETTING TO THE NUB OF CHEMISTRY
We now have the tools to calculate the pH of a solution that is a mixture of a strong acid and a strong
base.

Calculate the pH of the solution formed when 25 cm3 of 0.250 mol dm-3 H2SO4 is added to 100 cm3 of
0.200 mol dm-3 NaOH.
GETTING TO THE NUB OF CHEMISTRY
3.1.12.4 Weak acids and bases Ka for weak acids

Strong acids, such as HCl, dissociate completely.

HA(aq) à H+(aq) + A-(aq)

Weak acids, such as CH3COOH, only partially dissociate.

HA(aq) ⇌ H+(aq) + A-(aq)

As such, the acid dissociation constant, Ka, is introduced. (Ka – has units mol dm-3)

[" ! ][%" ]
Ka =
["%]

If only the weak acid is present, the equilibria shows that [H+] = [A-]

[" ! ]#
Ka =
["%]

[𝐻$ ]+ = Ka[HA]

[H+] = -𝐾, [𝐻𝐴]


Ka for Ethanoic Acid

CH3COOH (aq) ⇌ H+(aq) + CH3COO-(aq)

CH3COOH is a weak acid so fery few of its molecules dissociate

[CH3COOH ] hardly changes

[CH3COOH ]equilib ≈ [CH3COOH ]initial

[𝐻$ ][𝐶𝐻* 𝑂𝑂' ]


𝐾, =
[𝐶𝐻* 𝑂𝑂𝐻]-.-/-,0

Since Ka values for weak acids are very small it is convenient to use the pKa scale where:

pKa = -log10 Ka (and Ka = 10-pKa)

If Ka = 1.70 x 10-5 mol dm-3

pKa = -log10(1.70 x 10-5 mol dm-3)

pKa = 4.77
GETTING TO THE NUB OF CHEMISTRY
Stronger acids – higher Ka values
Stronger acids – lower pKa values

Acid Ka pKa
HCl 1.0 x 107 -7.00
HNO3 4.0 x 101 -1.60
HF 5.6 x 10-4 3.25
HCOOH 1.6 x 10-4 3.80
CH3COOH 1.7 x 10-5 4.82

When a weak acid reacts with a strong base, for every mole of OH- added, one mole of HA is used up and
one mole of A- is formed.
GETTING TO THE NUB OF CHEMISTRY
GETTING TO THE NUB OF CHEMISTRY
GETTING TO THE NUB OF CHEMISTRY
Half neutralisation of a weak acid: Calculate the pH of the solution formed when 100 cm3 of 0.200 mol
dm-3 ethanoic acid (pKa = 4.76) is added to 40 cm3 of 0.250 mol dm-3 KOH.
GETTING TO THE NUB OF CHEMISTRY
3.1.12.6 pH curves, titrations and indicators

The term equivalence point refers to the point at which the solutions have been mixed in exactly the
right proportions according to the equation. The term end point is where the indicator changes colour.
They are not the same thing.

How does pH change during a titration between a strong acid and a strong base? We can show this
mathematically using the following:

(23-4-.,0 62078 29 : ! ) ' (62078 29 <: " ,==7=)


[H+] = (>20?67 29 ,@-=) $ (>20?67 29 A,87 ,==7=)

pH = -log [H+]

Consider the reaction between 20cm3 of 0.100 moldm-3 HCl and 0.100 moldm-3 NaOH.

Vol of NaOH added (cm3) pH

0 1
10 1.48
19.0 2.59
19.9 3.60
19.99 4.60
20.00 7.00
20.01 9.40
21.00 11.40
25.00 12.05
40.00 12.70

14

12 • As the alkali is added, the pH


changes slowly at first. It
10
does, however, change very
8
rapidly from about 3.6 to
11.4 at the equivalence
pH

6 point.
• So, an indicator that changes
4
colour between pH 3.5 and
2 9.5 will identify the
equivalence point.
0
0 5 10 15 20 25 30 35 40 45
Volume of NaOH added (cm3)
GETTING TO THE NUB OF CHEMISTRY
The following is a useful rule of thumb regarding the four types of titration curve: strong acid/strong
base, weak acid/strong acid, strong acid/weak base and weak acid/weak base

The Rule of Two:

• A weak acid starts or finishes at 2 pH units higher than a strong acid.

• A weak acid has an equivalence point 2 pH units higher than a strong acid.

• A weak base starts or finishes at 2 pH units lower than a strong base.

• A weak base has an equivalence point 2 pH units lower than that of a strong base.

• For both weak acids and bases, the vertical range is +/- 2 pH units about the equivalence point pH
GETTING TO THE NUB OF CHEMISTRY
GETTING TO THE NUB OF CHEMISTRY
Diprotic titration

Indicators are weak acids or bases that have different colour to their conjugate bases / acids.

pH range of colour
Indicator Colour at low pH Colour at high pH
change
Methyl orange Red 3.2-4.4 Yellow

Phenolphthalein Colourless 8.2 – 10.0 Purple

For an indicator to be useful it must change colour in the vertical part of the pH curve.
GETTING TO THE NUB OF CHEMISTRY
How to sketch or complete a titration curve.

1. If you’re given a strong acid’s concentration, or the concentration and Ka of a weak acid,
calculate the starting pH.

2. If it’s a weak acid, remember the mini rise at the start before the gradient becomes shallower.
Also, in the case of a weak acid, pH = pKa at half-neutralisation.

3. Calculate the volume of base that need to be added so that the acid and base are mixed in the
right proportions for the equivalence point.

4. Find the pH of the mixture when all the base has been added to determine where the line ends
on your sketch.

For example:

Flask: 15.0 cm3 of 0.20 moldm-3 CH3COOH (Ka = 1.70 x 10-5)


Added: 50.0 cm3 of 0.10 moldm-3 NaOH
GETTING TO THE NUB OF CHEMISTRY
A 0.100 mol dm solution of sodium hydroxide was gradually added to 25.0 cm3 of a solution of a
−3

weak acid, HX, in the presence of a suitable indicator. The acid dissociation constant of HX, Ka, =
2.62 × 10−5 mol dm−3

Calculate the starting pH.

At half-neutralisation, pH = pKa
GETTING TO THE NUB OF CHEMISTRY

Another example.
GETTING TO THE NUB OF CHEMISTRY
3.1.12.6 Buffer action

Buffers are solutions that resist changes in pH.

An acidic buffer is made from a weak acid and its strong base salt e.g. ethanoic acid and sodium
ethanoate. A basic buffer is made from a weak base and its strong acid salt e.g. ammonia and
ammonium chloride.

How Buffers Work


HA(aq) ⇌ H+(aq) + A-(aq)

[" ! ][%" ]
Ka =
["%]

&B ["%]
H+ =
[%" ]

1. If you add an acid, you increase the concentration of H+ ions. This shifts the equilibrium to the
left.
2. If you add an alkali, the OH- ions react with the H+ ions, lowering their concentration, and so the
equilibrium shifts to the right.
3. If the starting concentrations of the weak acid and its salt are large enough, and the moles of
added acid/alkali relatively small, then [H+] remains almost constant and therefore so does the
pH.

Some assumptions are made

• The equilibrium concentration of acid = the initial concentration of the acid.


• The equilibrium concentration of salt = the initial concentration of the salt.
GETTING TO THE NUB OF CHEMISTRY
To see what we mean, let’s consider what happens to the pH of two different solutions that have the
same starting pH when we add 5.00 cm3 of 0.1 mol dm-3 of HCl to each of them.
GETTING TO THE NUB OF CHEMISTRY

We added 5.00 cm3 of 0.1 mol dm-3 of HCl to solutions 1 and 2. The pH of 1 changed by 0.50 whereas 2
changed by 0.02. 2 is an example of a buffer.

Buffers can be made by:

• Adding the necessary salt to a weak acid.

E.g. 50.0 cm3 of 1.00 mol dm-3 methanoic acid (Ka = 1.78 x 10-4 mol dm-3) mixed with 20.0 cm3 of 1.00
mol dm-3 sodium methanoate.
GETTING TO THE NUB OF CHEMISTRY
E.g. 1.00 g of potassium ethanoate is dissolved in 50.0 cm3 of 0.200 mol dm-3 ethanoic acid
(Ka = 1.74 x 10-5 mol dm-3).

• By partial neutralisation

E.g. 25.0 cm3 of 0.500 mol dm-3 methanoic acid (Ka = 1.78 x 10-4 mol dm-3) is mixed with 10.0 cm3 of 1.00
mol dm-3 sodium hydroxide.
GETTING TO THE NUB OF CHEMISTRY
Extension: Equivalence Points

To remind ourselves, the equivalence point, or stoichiometric point, of a chemical reaction is the point at
which chemically equivalent quantities of reactants have been mixed. Neither reagent is in excess.

As we would expect, we get a pH of 7 when a strong acid and strong base are mixed in exactly the right
proportions. However, when it’s a weak acid and strong base, the pH is around 9. And with a strong acid
and weak base it’s around 5. Why?

A weak acid and strong base result in the following reaction:

HA(aq) + OH– à A– (aq) + H2O(l)

One of the products, A– (aq), is a conjugate base (it’s a proton acceptor). This conjugate base reacts with
water to form a slightly basic solution.

A– (aq) + H2O (l) ⇌ HA + OH–

As a result, [OH–] > [H+] and the resulting mixture at the equivalence point has a pH > 7.

A weak base, such as NH3, and strong acid result in the following reaction:

NH3(aq) + H+(aq) à NH4+(aq)

The product, NH4+(aq), is a conjugate acid (it’s a proton donor). This conjugate acid reacts with water to
form a slightly acidic solution.

NH4+(aq) + H2O (l) ⇌ NH3(aq) + H3O+(aq)

As a result, [H+] > [OH–]


GETTING TO THE NUB OF CHEMISTRY
3.2.2 Group 2, the alkaline earth metals
The elements in Group 2 are called the alkaline earth metals. The trends in the solubilities of the
hydroxides and the sulfates of these elements are linked to their use. Barium sulfate, magnesium
hydroxide and magnesium sulfate have applications in medicines whilst calcium hydroxide is used in
agriculture to change soil pH, which is essential for good crop production and maintaining the food
supply.

As a topic, Group 2 is mostly about trends from Mg to Ba. Here they are:

Atomic radius Ionisation energy Reactivity


Mg
Ca
Sr
Ba
• Increasing no. of electron • Increasing distance between nucleus and outer electrons
shells • Increasing shielding
• Decreasing electrostatic attraction between nucleus and
outer electrons

Melting Point

• As you go down Group 2, the melting


point decreases.
!"#$%&
• Attraction ∝ $#'()*
• Down the group, charge remains the
same but ionic radii increases.
• Weaker attraction between positive
metal ions and delocalised electrons.
• Mg is an anomaly due to a change in the
way the ions pack together.

Solubility of hydroxides Solubility of sulfates


M(OH)2 MSO4
Very insoluble (white ppt) Very soluble
Mg
Ca
Sr
Ba
Very soluble Very insoluble (white ppt)
GETTING TO THE NUB OF CHEMISTRY
Group 2 metals are strong reducing agents, losing electrons to form ions with a charge of 2+ when they
react.

M à M2+ + 2e-

They all tarnish in air forming a layer of oxide on their surface and, apart from beryllium, burn brightly in
oxygen when heated. All Group 2 compounds are white solids at room temperature.

M(s) + O2(g) à MO(s)

Reactions of Group 2 metals with water


Very slow reaction with water at room temp. Mg(s) + 2H2O(l) à Mg(OH)2(aq) + H2(g)
Mg Vigorous reaction with steam, bright white flame Mg(s) + H2O(g) à MgO(s) + H2(g)

Ca Fizzes and white suspension eventually forms as Ca(s) + 2H2O(l) à Ca(OH)2(aq) + H2(g)
Ca(OH)2 is only slightly soluble.
Sr Fizzes Sr(s) + 2H2O(l) à Sr(OH)2(aq) + H2(g)
Ba Fizzes Ba(s) + 2H2O(l) à Ba(OH)2(aq) + H2(g)

How acidic a substance is depends on the concentration of H+ ions in solution and OH-(aq) ions affect this.
Because of the solubility trend of Group 2 hydroxides, we can consider a trend in alkalinity (assuming we
consider an equivalent amount of moles).

Mg(OH)2
Ca(OH)2 ≈ pH 10-11
Sr(OH)2
Ba(OH)2 ≈ pH 14
Increasingly alkaline

This diagram shows the method for reacting magnesium with steam.
GETTING TO THE NUB OF CHEMISTRY
Uses of Group 2 metals their compounds

Substance Use

Mg • Mg is used as a reducing agent in the extraction of titanium from its ore.

TiO2 + 2C + Cl2 à TiCl4 + 2CO

TiCl4 + 2Mg à Ti + 2MgCl2

Mg(OH)2 • Mg(OH)2 is used as an antacid in certain indigestion remedies.

Mg(OH)2(s) + 2HCl(aq) à MgCl2(aq) + 2H2O(l)

Ca(OH)2 • Commonly known as ‘slaked lime’, Ca(OH)2 is used in agriculture to raise the pH
of soil.

CaO / CaCO3 • CaO and CaCO3 are used to remove SO2 from flue gases in, say, coal-fired power
stations.

• SO2 is acidic and CaO or CaCO3 act as bases to neutralise it.

CaO(s) + SO2(g) à CaSO3(s)

CaCO3(s) + SO2(g) à CaSO3(s) + CO2(g)

BaSO4 • BaSO4 is insoluble in water. In medicine, an insoluble suspension of BaSO4,


known as a ‘Barium meal’, is sometimes used when investigating the digestive
tract using X-rays.

• The patient drinks a suspension of BaSO4 in water. X rays cannot pass through
the ‘meal’, which shows up on the image.

• Even though Ba2+ ions are toxic to the human body, this method isn’t harmful
because of the insolubility of BaSO4 in water.

BaCl2 • BaCl2(aq) is a source of Ba2+ ions and these can be used to test for SO42- ions – a
white precipitate is formed. This is the ionic equation:

Ba2+(aq) + SO42-(aq) à BaSO4(s)

• Acidified BaCl2(aq) is added to the solution being tested. Dilute hydrochloric acid is
used to acidify the solution to react with any impurities, such as aqueous
carbonate or sulfite ions, which would also give a white precipitate with Ba2+
ions.
GETTING TO THE NUB OF CHEMISTRY
3.2.3 Group 7(17), the halogens
The halogens in Group 7 are very reactive non-metals. Trends in their physical properties are examined
and explained. Fluorine is too dangerous to be used in a school laboratory but the reactions of chlorine
are studied. Challenges in studying the properties of elements in this group include explaining the
trends in ability of the halogens to behave as oxidising agents and the halide ions to behave as reducing
agents.

Appearance Melting / Boiling Point Explanation


F2 Pale yellow gas
Cl2 Pale green gas • Increasing number of electrons
• Increasing strength of van der Waals forces
Br2 Orange/brown liquid
of attraction
I2 Grey crystalline solid
Increasing

• Bromine is a volatile liquid and forms an orange gas readily.


• Iodine forms a purple vapour when heated.
• Iodine dissolved in water (aq) is a brown solution.

Atomic radius Ionisation energy Reactivity


F
Cl
Br
I
• Increasing no. of electron • Increasing distance between nucleus and outer electrons
shells • Increasing shielding
• Decreasing electrostatic attraction between nucleus and
outer electrons

Electronegativity Remember, electronegativity is power of an atom to attract the two


increasing electrons in the covalent bond
F Down the group:
• more shells
Cl
• more shielding
Br • weaker attraction between nucleus and pair of electrons in covalent
I bond
GETTING TO THE NUB OF CHEMISTRY
Chlorine plus Water

Chlorine dissolves in water to give chlorine water Cl2(aq), and some chlorine then reacts with water.

Cl2(aq) + H2O(l) ⇌ HCl(aq) + HClO(aq)

This is a disproportionation reaction as the oxidation state of Cl changes from 0 to -1 and +1. HClO(aq) is
called hypochlorous acid, which is a weak acid:

HClO(aq) + H2O(l) ⇌ H3O+(aq) + ClO-(aq)

The hypochlorite ion (ClO-) is a powerful disinfectant and bleach – it kills bacteria.

The hypochlorite ion can decompose when exposed to sunlight.

2 ClO-(aq) à 2Cl-(aq) + O2(g)

Society assesses the advantages and disadvantages when deciding if chemicals should be added to water
supplies. For example, the benefits to health of water treatment by chlorine outweigh its toxic effects.
Ditto the addition of sodium fluoride to drinking water

The equilibrium mixture of chlorine in water is acidic.

Cl2(aq) + H2O(l) ⇌ HCl(aq) + HClO(aq)

If an aqueous base, such as sodium hydroxide, is added, the equilibrium is shifted much further to the
right. For example, sodium hypochlorite is produced when chlorine reacts with cold aqueous sodium
hydroxide.

Cl2(aq) + 2NaOH(aq) à NaCl(aq) + NaClO(aq) + H2O(l)

As we’ve seen, the hypochlorite ion (ClO-) is a powerful disinfectant and bleach – it kills bacteria.
GETTING TO THE NUB OF CHEMISTRY
All the halogens are oxidising agents. In redox reactions, they accept electrons, forming halide ions. For
example:
2Na(s) + Cl2(g) à 2NaCl(s)

The half equations for this reaction are:

Na à Na+ + e- (oxidation)
2e- + Cl2 à 2Cl- (reduction)

The oxidising power of the halogens decreases down the group: F2 > Cl2 > Br2 > I2 (a smaller atomic
radius, less shielding = a stronger attraction with the electron gained).

This can be demonstrated using potential redox reactions involving each other in their elemental and
reduced forms.

Cl-(aq) Br-(aq) I-(aq)


Cl2(aq) Reaction observed Reaction observed
Cl2(aq) + 2Br-(aq) à2Cl-(aq) + Br2(aq) Cl2(aq) + 2I-(aq) à2Cl-(aq) + I2(aq)
Br2(aq) Reaction observed
No reaction
Br2(aq) + 2I-(aq) à2Br-(aq) + I2(aq)
I2(aq) No reaction No reaction

Similarly, there is a trend in the reducing power of the Halide ions: I- > Br- > Cl- > F- (a larger ionic radius,
more shielding = a weaker attraction with the electron lost). When a halide ion reduces another
substance, the halide is oxidised to a halogen. This following table compares how well the halides reduce
concentrated H2SO4.

Halide Product Observation Type of reaction Equations


Cl- HCl(g) misty fumes acid-base H+ + Cl- → HCl
HBr(g) misty fumes acid-base H+ + Br- → HBr
2Br- à Br2 + 2e-
Br2(g) orange fumes oxidation of Br- 2e + 2H+ + H2SO4 à SO2 + 2H2O
-

Br- Overall: 2Br- + 2H+ + H2SO4 → Br2 + SO2 + 2H2O


colourless gas 2Br- à Br2 + 2e-
SO2(g) (turns blue litmus reduction of H2SO4 2e + 2H+ + H2SO4 à SO2 + 2H2O
-

paper red) Overall: 2Br- + 2H+ + H2SO4 → Br2 + SO2 + 2H2O


Hl(g) misty fumes acid-base H+ + l- → Hl
2I- à I2 + 2e-
I2(s) black solid oxidation of I- 2e- + 2H+ + H2SO4 à SO2 + 2H2O
Overall: 2I- + 2H+ + H2SO4 → I2 + SO2 + 2H2O
colourless gas 2I- à I2 + 2e-
SO2(g) (turns blue litmus reduction of H2SO4 2e- + 2H+ + H2SO4 à SO2 + 2H2O
I- paper red) Overall: 2I- + 2H+ + H2SO4 → I2 + SO2 + 2H2O
2I- à I2 + 2e-
S(s) yellow solid reduction of H2SO4 6e + 6H + H2SO4 à S + 4H2O
- +

Overall: 6I- + 6H+ + H2SO4 → 3I2 + S + 4H2O


2I- à I2 + 2e-
colourless gas
H2S(g) reduction of H2SO4 8e- + 8H+ + H2SO4 à H2S + 4H2O
(rotten egg smell)
Overall: 8I- + 8H+ + H2SO4→ 4I2 + H2S + 4H2O
GETTING TO THE NUB OF CHEMISTRY
GETTING TO THE NUB OF CHEMISTRY
The silver halides are used to identify the chloride, bromide and iodide ions. Acidified silver nitrate can
be used to identify and distinguish between them. Dilute nitric acid is initially added to remove any other
dissolved anions that could form a precipitate with the silver ions in silver nitrate solution, such as
carbonate, CO32-(aq), or sulfite, SO32-(aq), ions.

The ionic equations for these precipitation reactions are:

Ag+(aq) + Cl-(aq) à AgCl(s)


Ag+(aq) + Br-(aq) à AgBr(s)
Ag+(aq) + l-(aq) à Agl(s)

• Ammonium ions (NH4+) can be identified in a solution by adding dilute sodium hydroxide solution
and gently heating. If ammonium ions are present, they will be converted to ammonia gas, which
turns damp red litmus paper or damp universal indicator paper blue.
• Carbonate ions (CO32-) can be identified by adding an acid – effervescence is observed.
• Sulfate ions (SO42-) can be identified by adding barium chloride – a white precipitate is formed.
• Hydroxide ions (OH-) can be identified by adding universal indicator or using red litmus paper – both
turn blue.
Element Na Mg Al Si P4 S8
Atomic radius / Atomic radius decreases (increasing nuclear charge, same shielding  stronger attraction with outer electrons)
Electronegativity /
Ionisation energy Electronegativity increases (increasing nuclear charge, same shielding, decreasing atomic radius)  stronger attraction)
Ionisation energy increases (increasing nuclear charge, same shielding, decreasing atomic radius)  stronger attraction)
Structure Metallic Metallic Metallic Giant Covalent Molecular Molecular
Mg(s) + H2O(g)  MgO(s) + H2(g)
(steam - fast)
Element + Water 2Na(s) + 2H2O(l)  2NaOH(aq) + H2(g) – – – –
Mg(s) + 2H2O(l)  Mg(OH)2(s) + H2(g)
(water - slow)
S(s) + O2(g)  SO2(g)
Element + Oxygen 4Na(s) + O2(g)  2Na2O(s) 2Mg(s) + O2(g)  2MgO(s) 4Al(s) + 3O2(g)  2Al2O3(s) Si(s) + O2(g)  SiO2(s) P4(s) + 5O2(g)  P4O10(s)
2SO2 (g) + O2(g) ⇌ 2SO3 (g)
Flame colour Orange-yellow White White White Blue
Oxide Na2O MgO Al2O3 SiO2 P4O10 SO2, SO3
Structure Ionic Ionic Ionic Giant Covalent Molecular Molecular

Structure of
molecule

Na2O + H2O  2NaOH MgO + H2O  Mg(OH)2 P4O10 + 6H2O  4H3PO4 SO2 + H2O ⇌ H2SO3 (sulfurous acid)
Oxide + Water O2- + H2O  2OH- O2– + H2O  2OH– Insoluble Insoluble (phosphoric acid) SO3 + H2O ⇌ H2SO4 (sulfuric acid)
~ pH 14 ~ pH 9 (slightly soluble) ~ pH 1 ~ pH 2-3 ~ pH 1

Structure of acid

Reaction with acid Basic Basic Amphoteric Acidic Acidic Acidic


Al2O3 + 6H+  2Al3++ 3H2O SiO2 + 2OH–  SiO32– + H2O SO2 + 2OH–  SO32–+ H2O
/ alkali Na2O + 2H+  2Na+ + H2O MgO + 2H+  Mg2+ + H2O P4O10 + 12OH–  4PO43– + 6H2O
Al2O3 + 2OH– + 3H2O  2Al(OH)4– (must be hot, concentrated NaOH) SO3 + 2OH–  SO42– + H2O
Melting Points of Elements Melting Points of Oxides
GETTING TO THE NUB OF CHEMISTRY
3.2.5.1 General properties of transition metals
What’s the difference between a Transition Metal and a ‘d block’ Element?

• A transition metal is an element that has an incomplete d sub-level in either its atoms or ions.
• A ‘d block’ element contains its outer energy electrons in the d block.

This definition excludes and zinc from the list

• The electronic configuration of Zn: 1s2 2s2 2p6 3s2 3p6 4s2 3d10
• The electronic configuration of Zn2+: [1s2 2s2 2p6 3s2 3p6 3d10]2+

The characteristic properties of transition metals include:

1. They form coloured ions.

V2+ V3+ VO2+ VO2+

2. They exhibit variable oxidation states.

Remember, the 4s electrons that are lost first when transition metal ions are formed.

• Ti 1s2 2s2 2p63s2 3p6 3d24s2


• Ti2+ 1s2 2s2 2p63s2 3p6 3d2
• Cr 1s2 2s2 2p63s2 3p6 3d54s1
• Cr3+ 1s2 2s2 2p63s2 3p6 3d3
GETTING TO THE NUB OF CHEMISTRY
3. They form complexes with ligands. A ligand is a molecule or ion that forms a co-ordinate bond
with a transition metal by donating a pair of electrons. A complexes’ co-ordination number is the
number of co-ordinate bonds to the central metal atom or ion.

[Cu(H2O)6]2+ [CuCl4]2-
OH2 2+ Cl 2-
H2O OH2
Cu Cu
H2O OH2 Cl
Cl
OH2 Cl
Co-ordination no. = 6 Co-ordination no. = 4

4. They show catalytic activity. E.g.

• Ni margarine production
• V2O5 making SO3 for H2SO4
• Fe Haber process to make NH3
• Pt, Pd catalytic converters

Why do variable oxidation states arise?

The more highly charged the ion, the more electrons you must remove and the greater the ionisation
energy you have to provide.

Transition metals from Sc to Cu have electrons of similar energy in both the 3d and 4s levels. There isn't a
huge jump in the amount of energy you need to remove the third electron compared with the first and
second.

The extra energy required to remove a third electron from an Fe2+ ion (3rd I.E.) to form an Fe3+ ion is only
+2960 kJ mol-1. Offsetting this, the more highly charged the ion, the more energy is released either as
lattice enthalpy or the hydration enthalpy of the metal ion

This means that one element can form ions of roughly the same stability by losing different numbers of
electrons.

The common oxidation states for each element include +2 or +3 of both. The transition metals usually
exhibit their highest oxidation states in compounds with oxygen and fluorine.

Ti, V, Cr and Mn never form simple ions in their highest oxidation state (+4 or higher) as this would result
in ions of extremely high charge density. Hence these oxidation states are covalently bonded, often in an
anion e.g. MnO4–.

In general, the maximum oxidation state for a transition metal is the number of 4s electrons plus the
number of unpaired 3d electrons.
GETTING TO THE NUB OF CHEMISTRY
3.2.5.2 Substitution Reactions
What is a complex ion?

A complex ion has a metal ion at its centre with a number of other molecules or ions surrounding it.
These are attached to the central ion by co-ordinate (dative) covalent bonds. All ligands have a lone
pairs of electrons in the outer energy level that are used to form co-ordinate (dative) bonds with the
metal ion. The exact nature of the bonding is beyond A level.

The positive metal ion is stabilised by the ligands and so (with a few exceptions) an ion will make as
many bonds with as many ligands as it is able, up to 6.

Unidentate ligands – form one co-ordinate bond (dentate means ‘bite’). e.g. H2O:, :OH-, :NH3, :CN-, :Cl-

[Cu(H2O)6]2+ [CuCl4]2-
OH2 2+ Cl 2-
H2O OH2
Cu Cu
H2O OH2 Cl
Cl
OH2 Cl
Co-ordination no. = 6 Co-ordination no. = 4

Bidentate ligands have two lone pairs, both of which can bond to the central metal ion. E.g. 1,2-
diaminoethane (NH2CH2CH2NH2) and ethandioate (C2O42-).

[Cr(NH2CH2CH2NH2)3]3+ [Cr(C2O4)3]3-
O
NH2 3+ O C
C O 3-
H2N NH2 O O C O
Cr Cr
H2N NH2 O O C O
C NH2
NH2 O C
O

Co-ordination no. = 6 Co-ordination no. = 6

Multidentate ligands attach to the central metal ion with 3 or more dative bonds. Haem is a co-factor
found in the blood. It consists of an Fe2+ ion surrounded by a porphyrin ligand. Haem is an important
part of haemoglobin
GETTING TO THE NUB OF CHEMISTRY
The biochemical importance of iron in blood relates to its role in haemoglobin and ligand substitution.
Each haemoglobin protein is made up of 4 haem units like the one in the picture. Each iron atom is able
to bind to one oxygen molecule in addition. The oxygen molecule is bound to the iron atom by a
coordinate bond and as such is considered another ligand.

Carbon monoxide is toxic to the body as it can replace the oxygen in the haem group. It is a stronger
more stable bond and so this exchange is essentially irreversible much, much slower release than O2
from the iron atom as it bonds around 200 times more strongly so dangerous levels can build up.

EDTA (Ethylenediamminetetraacetic acid) is a hexadentate ligand. Laid out EDTA has the following
structure:

When EDTA attaches itself as a ligand the 4 oxygen atoms lose a hydrogen ion and are left with a 1-
charge. The other 2 bonds are from the lone pairs of the two nitrogen atoms.

It is able to ‘trap’ ions e.g. Fe2+ and Ca2+ and in doing so reduces their reactivity. It has many industrial
and medical applications including: Reduction of metal ion impurities; In chelation therapy to remove
mercury and lead when somebody is poisoned.
GETTING TO THE NUB OF CHEMISTRY
What is ligand substitution?

A ligand substitution reaction is one where one ligand is replaced by another ligand. If the ligands are a
similar size (actually, the atom/ion in the ligand that bonds to the metal), then there will be no change in
co-ordination number if a substitution reaction takes place. For example, H2O and NH3 ligands are a
similar size.

[Co(H2O)6]2+ + 6NH3 à [Co(NH3)6]2+ + 6H2O

In some ligand substitution reactions, only some of the ligands are replaced, for example

[Cu(H2O)6]3+ + 4NH3 à [Cu(H2O)2(NH3)4]2+ + 4H2O

Substitution by bigger/smaller ligands. If the ligands are a different size then the co-ordination number
may change. For example, Cl– ligands are significantly bigger than the O in H2O or the N in NH3 and only
four Cl– ligands can bond to most transition metal ions. In this reaction, the co-ordination number
changes from 6 to 4.

[Co(H2O)6]2+ + 4Cl– à [CoCl4]2– + 6H2O

Steric effects are the effects seen in molecules that come from the fact that atoms occupy space. When
atoms are put close to each other, this costs energy due to repulsion. The electrons near the atoms
minimise this repulsion by being as far apart as possible from each other. This can change the way
molecules react. It can also change the shape (or conformation) of the molecule.

Substitution by ligands that form more co-ordinate bonds (the chelate effect)

Chelates are very stable as they often replace multiple unidentate ligands: a reaction which has an
increase in entropy.

[Cu(H2O)6]2+ + EDTA4- à [Cu(EDTA)]2- + 6H2O

[Cu(H2O)6]2+ + 3C2O42- à [Cu(C2O4)3]4- + 6H2O

[Co(C2O4)3]4– + EDTA4– à [Co(EDTA)]2– + 3C2O42–

In many ligand substitution reactions, DH is negligible as the same number of the same type of similar
bonds are being broken and formed. In this example above, six Cr-N bonds are being broken and six Cr-N
bonds being formed and so DH is negligible.

As the feasibility of a reaction depends on DG = DH – TDS, if DH is negligible and DS very positive (a big
increase in entropy), then DG is very negative and the reaction is feasible.

In general, in a ligand substitution reaction where ligands are replaced with by those that form more co-
ordinate bonds, the reaction is feasible and this is driven by the entropy increase. Reactions which would
have ligands replaced by those that make fewer co-ordinate bonds are not feasible due to an entropy
decrease.
GETTING TO THE NUB OF CHEMISTRY
3.2.5.3 Shapes of complex ions
Linear Square Planar Tetrahedral Octahedral

Co-ordination no. 2 4 4 6

Shape

Bond angle(s) 180o 90o 109.5o 90o

Most complexes
2+ 2+
Pt and Ni With larger with small ligands,
Occurrence Ag+ complexes
complexes ligands, such as Cl- such as H2O and
NH3

[Ag(NH3)2]+
Example (in Tollen’s [PtCl4]2- [CuCl4]2- [Cu(H2O)6]2+
reagent)

Octahedral complexes can display cis–trans isomerism (a special case of E–Z isomerism) with unidentate
ligands. With trans complexes, a pair of ligands are across from each other (∠ ligand-metal ion-ligand =
180o). In cis complexes, the ligands on the same side and ∠ ligand-metal ion-ligand = 90o.

Square planar complexes are also formed and can display cis–trans isomerism.
GETTING TO THE NUB OF CHEMISTRY
Octahedral complexes can display optical isomerism with bidentate ligands.

3.2.5.4 Formation of coloured ions


Most of the compounds of transition metals are coloured. This is because they only absorb light in
certain areas of the visible spectrum. The compound takes on the colour of the light it transmits /
reflects. For example, a solution of copper (II) sulfate is blue because red and yellow wavelengths are
absorbed.

The absorption of light can be related to incompletely filled d orbitals in the transition metal ion. When
radiation is absorbed by substances, electrons are promoted form lower to higher energy levels. In most
cases, this jump is so large that the radiation absorbed is likely in the ultraviolet region of the
electromagnetic spectrum. In the case of transition metals, however, this promotion of electrons
happens between orbitals within the d sub-shell.

Placing ligands around a central ion causes the energies of the d orbitals to change. Some of the d
orbitals gain energy and some lose energy.

OCTAHEDRAL TETRAHEDRAL

3d 3d
GETTING TO THE NUB OF CHEMISTRY
Once ligands bond, the five d orbitals are no longer have the same energy. The degree of splitting
depends on the central ion and the ligand. The energy difference between the levels affects how much
energy is absorbed when an electron is promoted. The amount of energy governs the colour of light
absorbed.

Quantum theory tells us that energy is emitted in discrete amounts known as energy quanta. The energy
gap (∆E) between the higher and lower energy orbitals corresponds to the energy of a photon of light
and is part of Bohr's frequency condition.

DE = hν

Where h is Planck’s constant and ν is the frequency of the radiation. It tells us that the frequency of a
photon absorbed or emitted during the transition of an electron is related to the energy difference (ΔE)
between the two energy levels.

Bohr's frequency condition, ∆E = hν, can be taken one step further. We can use c the speed of light and
the wavelength (lambda) 𝝀 to replace frequency (ν) in the equation.

𝒉𝒄
DE = hν =
𝝀

DE = gap in energy (J)


h = Planck’s constant (6.63 x 10–34 Js)
ν = frequency of light (s-1)
c = velocity of light (3.00 x 108 ms–1)
𝝀 = wavelength of light (m)

Various factors affect the colour seen.

Factor Complex 1 Complex 2


Metal [Cu(H2O)6]2+ Blue [Fe(H2O)6]2+ Green
Oxidation state [Fe(H2O)6]3+ Pale Violet [Fe(H2O)6]2+ Green
Ligand(s) [Cu(H2O)6]2+ Blue [Cu(H2O)4(NH3)2]2+ Deep Blue
Co-ordination number* [Cu(H2O)6]2+ Blue [CuCl4]2- Yellow

*The number of ligands changes the co-ordination number but you will also be changing the ligand so
there are two factors changing the energy gap in this instance.
GETTING TO THE NUB OF CHEMISTRY
Both zinc and scandium form colourless (in solution) and white (as solids) compounds.

↑↓ ↑↓
When Zinc dx - y2
2 d z2
forms an ion
its d-orbital 3d ↑↓ ↑↓ ↑↓ ↑↓ ↑↓
is full: 3d10 Average energy in the
presence of ligands ↑↓ ↑↓ ↑↓
dxy dxz dyz
When
Scandium
forms an ion dx2 - y2 dz2
its d-orbital
is empty: 3d

3d0 Average energy in the


presence of ligands
dxy dxz dyz

Colorimeters can be used to track the formation of a coloured product of removal of a coloured
reactant. They can also be calibrated against solutions of known concentration in order to establish the
concentration of an unknown sample

If we plotted a graph of absorbance against different known concentrations we could then test a sample
of unknown concentration.
Absorbance
×
×
×
X
×

M Molarity
GETTING TO THE NUB OF CHEMISTRY
3.2.5.5 Variable oxidation states
Transition elements show variable oxidation states, and these have uses.

Use Reagent
• Silver mirror formed in presence of aldehydes.
Tollen’s reagent • Ag(+1) in [Ag(NH3)2]+ reduced to Ag(0) in silver
(contains [Ag(NH3)2]+ mirror.
• Aldehyde oxidised to carboxylic acid
Testing for aldehydes
• Brick-red precipitate (Cu2O) formed in presence
Fehling’s solution of aldehydes.
(contains Cu2+) • Cu(+2) reduced to Cu(+1) in Cu2O.
• Aldehyde oxidised to carboxylic acid
• Colour change from orange (Cr2O72-) to green
Testing for alcohols
(Cr3+).
(primary & secondary) Acidified K2Cr2O7
• Cr(+6) in Cr2O72- reduced to Cr(+3).
aldehydes
• Alcohols / aldehyde oxidised
Redox titrations • E.g. acidified KMnO4 can be used to analyse
Fe2+, C2O42- and H2O2
• Many reactions are catalysed by transition
Catalysis metals and this often involves the transition
metal changing oxidation state

Vanadium has many different oxidation states with associated colour changes.

Oxidation State Species Present Colour


+5 VO2+ Yellow
+4 VO2+ Blue
+3 V3+ Green
+2 V2+ Violet

Vanadium (V) ions of the form VO2+ can be reduced by zinc in acidic solution all the way down to V2+

VO2+ + 2H+ + e– ⇌ VO2+ + H2O


VO2+ + 2H+ + e– ⇌ V 3+ + H2O
V3+ + e– ⇌ V2+

As the reaction takes place, the colour changes in succession from yellow to blue to green to violet.
(remember: You Better Get Vanadium).

The reduction of [Ag(NH3)2]+ (Tollens’ reagent) to metallic silver is used to distinguish between aldehydes
and ketones.
[Ag(NH3)2]+ + e– à Ag + 2NH3
CH3CHO +H2O à CH3COOH + 2H+ + 2e–

2[Ag(NH3)2]+ + CH3CHO +H2O à 2Ag + 4NH3 + CH3COOH + 2H+


GETTING TO THE NUB OF CHEMISTRY
If a compound has a high oxidation state it will have the potential to act as an oxidising agent

VO2+ + 2H+ + e– ⇌ VO2+ + H2O ; Eɵ +1.00V


Cu2+ + 2e ⇌ Cu ; Eɵ +0.34V

Eɵcell = 1.00 - 0.34 = +0.66V ∴ it’s feasible

2VO2+ + 4H+ + Cu ⇌ 2VO2+ + 2H2O + Cu2+


The copper is oxidised.

If a compound has a low oxidation state it will have the potential to act as a reducing agent.

V 3+ + e– ⇌ V2+ Eɵ -0.26V
Cu2+ + 2e ⇌ Cu Eɵ +0.34V

Eɵcell = 0.34 – 0.26 = +0.60V (feasible)

Cu2+ + 2V2+ ⇌ Cu + 2V3+


The copper ions are reduced

The redox potential for a transition metal ion changing from a higher to a lower oxidation state is
influenced by pH and by the ligand.

When aqueous transition metal ions undergo a change in oxidation state, the reactions frequently
involve hydrogen ions. By analysing the reaction equations, we can see the redox processes taking place.
The reduction of manganese in the manganate(VII) ion takes place readily in acidic solutions

MnO4- (aq) + 8H+ (aq) + 5e- ⇌ Mn2+ (aq) + 4H2O (l) Eꝋ = +1.51 V

The manganese changes from +7 to +2. The positive value of Eꝋ indicates the equilibrium position is well
over to the right. An excess of acid drops the pH and provides the hydrogen ions to ensure this reaction
usually goes to completion. However, in a neutral solution the reduction is limited to the +4 oxidation
state for the manganese

MnO4- (aq) + 2H2O (l) + 3e- ⇌ MnO2 (s) + 4OH- (aq) Eꝋ = +0.59 V

Notice that the value of Eꝋ has changed as a result of changing the pH of the solution

Changing the ligand also influences the redox potential. If you compare the aqueous nickel(II) ion with
the hexaamminenickel(II) ion, the Eꝋ becomes more negative when ammonia ligands replace water
ligands.

[Ni(H2O)6]2+ + 2e− ⇌ Ni + 6H2O Eꝋ = -0.26


[Ni(NH3)6]2+ + 2e− ⇌ Ni + 6NH3 Eꝋ = -0.49

Ammonia is a stronger ligand than water meaning that it binds better to the nickel(II) ion. The position of
the second equilibrium is slightly more to the left than the first one (hence the more negative Eꝋ value).
GETTING TO THE NUB OF CHEMISTRY
Redox reactions can be used in titrations (just as you have met acid-base reactions previously).
Potassium manganate (VII), which contains the MnO4– ion is a very common oxidising agent used in
redox titrations.
MnO4– + 8H+ + 5e– à Mn2+ + 4H2O
purple colourless

As the half-equation shows, it needs to be used in acidic conditions, and dilute sulfuric acid is the acid
used. Hydrochloric acid cannot be used as the MnO4– would also oxidise Cl– to Cl2 so affect the volume of
KMnO4 required in the titration.

In a normal redox titration like this, the purple potassium manganate (VII) is in the burette. The sample
being analysed is in the flask with an excess of dilute sulfuric acid. As the purple potassium manganate
(VII) is added it reacts and forms colourless Mn2+. At the end point, the purple MnO4– ions no longer
react and their colour remains. This means that the end point is the first hint of pink. It also means that
the titration does not need an indicator as it is self-indicating.

Fe2+ is analysed in redox titrations with acidified potassium manganate (VII).

• Half equation: Fe2+ à Fe3+ + e–

The ethanedioate ion (C2O42–) is analysed in redox titrations with acidified potassium manganate (VII).

• Half equation: C2O42– à 2CO2 + 2e–


GETTING TO THE NUB OF CHEMISTRY
GETTING TO THE NUB OF CHEMISTRY
Questions can feature a combination of acid-base and redox titrations. These typically involve a starting
material that is mixture of two substances, which we’ll call A and B. One of the titrations will give you the
moles of A and B combined. The other will give you the moles of just one of the substances e.g. A. So, if
we need to determine the moles of B, then substract the the smaller titration answer from the larger
one.
GETTING TO THE NUB OF CHEMISTRY
3.2.5.6 Catalysts
A catalyst is a substance that increases the rate of a chemical reaction but is not used up in the reaction.
Catalysts work by providing an alternative route with a lower activation energy (catalysts do not alter
DG, DH orDS for a reaction, only the activation energy). Catalysts speed the rate at which an equilibrium
is reached by speeding up the forward and reverse reaction - they do not alter the position of the
equilibrium.

A catalyst that is in a different phase to the reactants is a heterogeneous catalyst. The catalyst is usually
a solid and the reaction takes place on the surface. The bulk of industrial processes involve
heterogeneous catalysis.

e.g. Haber process 3H2 + N2 ⇌ 2NH3 Fe catalyst


e.g. Contact process 2SO2 + O2 ⇌ 2SO3 V2O5 catalyst
e.g. Hydrogenation C2H4 + H2 à C2H6 Ni catalyst

At least one of the reactants is adsorbed onto the surface (i.e. forms bonds to the atoms in the solid
surface). Note that adsorption is to go onto something, but absorption is to go into something. The
places on the surface where molecules are adsorbed are called active sites.

In an effective catalyst, the molecules can move about the surface, bonding to different active sites. The
adsorption of reactants onto the surface can result in increased reaction in several ways:

a. adsorption onto the surface effectively concentrates the reactants, i.e. brings them closer
together than in the gas phase, so increasing the likelihood of collision;
b. it may weaken some of the bonds in the molecule, making reaction easier;
c. it may position the molecule in a favourable orientation for reaction.

For the catalyst to work, molecules must be adsorbed onto the surface. However, if:

• adsorption is too weak: not many molecules will be adsorbed so the catalyst will have very little
effect
• adsorption is too strong: molecules will not be able to move around the active sites, and so be less
likely to meet another reactant and so be less likely to react (also any product will tend to remain
adsorbed on the surface)
GETTING TO THE NUB OF CHEMISTRY
Nickel and platinum are often used as catalysts for good reason. Some transition metals e.g. Tungsten
are too strong and stop the reactants from being desorbed. Others like silver are too weak and reactants
do not adsorb in the first place

More concentrated H2SO4 is manufactured than any other chemical. The key step in this process (the
Contact process) is the conversion of SO2 to SO3 which is a slow, reversible reaction. It is catalysed by
vanadium (V) oxide, V2O5.

V2O5 + SO2 à V2O4 + SO3


V2O4 + ½O2 à V2O5

Overall: SO2 + ½O2 à SO3

This is a rare example of a heterogenous catalyst being changed during the reaction and then
regenerated during a later step.

It is obvious that the larger the surface area of these catalysts, the less quantity of catalyst that is needed
to produce the same effect.

• Many surface catalysts are very expensive, and so maximising surface area has important cost
savings.
• The surface area is maximised by using a very thin coating of the catalyst on some type of
support medium (a support is required as the layer is too thin to support itself - often a ceramic
"honeycomb" structure is used as support - note that the catalyst must be able to bind to the
support).

Over time, Heterogeneous catalysts can become poisoned by impurities that block the active sites and
consequently have reduced efficiency; this has a cost implication.

A homogeneous catalyst is changed during the reaction and then regenerated during a later step.

Most reactions involving homogeneous catalysts take place in solution (where all species are either
liquids or are dissolved, so they are all in the same phase).

Reactions involving homogeneous catalysis proceed via an intermediate species formed from a reactant
and the catalyst, which then reacts further and regenerates the catalyst.

For example, the I–/S2O82– reaction

2I-(aq) + S2O82-(aq) → I2 (aq) + 2SO42- (aq) Fe2+(aq) or Fe3+(aq) catalyst

Both I- and S2O82- ions are negative and as such the rate of reaction is slow due to repulsion that results in
a high EA. The iron ions help to accelerate this reaction

2Fe2+ + S2O82- → 2Fe3+ + 2SO42-


2Fe3+ + 2I- → 2Fe2+ + I2
GETTING TO THE NUB OF CHEMISTRY
Another example of homogeneous catalysis is autocatalysis. An autocatalytic reaction speeds up
gradually as the reaction progresses because one of the products acts as a catalyst for the reaction. A
common example is the reaction between manganate ions and ethandioate ions (note how this is also a
reaction between two negative ions, which means there’s a high EA:

2MnO4- (aq) + 5C2O42- (aq) + 16H+ (aq) à 10CO2 (g) + 2Mn2+ (aq) + 8H2O (l)

The Mn2+ ions catalyse the reaction.

4Mn2+(aq) + MnO4-(aq) + 8H+(aq) → 5Mn3+(aq) + 4H2O(l)

2Mn3+(aq) + C2O42-(aq) → 2CO2(g) + 2Mn2+(aq)

Conc. vs time (most reactions) Conc. vs time (autocatalysed reactions)

You can see the slow (uncatalysed) reaction at


Concentrations are high at the beginning and so
the beginning. As catalyst begins to be formed in
the reaction is fast - shown by a rapid fall in the
the mixture, the reaction speeds up - getting
reactant concentration. As things get used up, the
faster and faster as more and more catalyst is
reaction slows down and eventually stops as one
formed. Eventually, of course, the rate falls again
or more of the reactants are completely used up.
as things get used up.

Q. In the autocatalytic reaction between manganate ions and ethandioate ions, how could you cause the
reaction to start more quickly before the autocatalysis got going? What observations would you expect to
make during this reaction?

A. Add a little Mn2+ ions or heat to increase the rate, Purple à pale pink colour change and effervescence
from the CO2
GETTING TO THE NUB OF CHEMISTRY
3.2.6 Reactions of ions in aqueous solution
In aqueous solution, the following metal-aqua ions are formed:

[M(H2O)6]2+ limited to M = Fe and Cu in the AQA specification


[M(H2O)6]3+ limited to M = Al and Fe in the AQA specification

Metal hexa-aqua ions can take part in one of three types of reactions:

1. Acid-Base / Hydrolysis – loss of H+ from an H2O ligand or ligands.


2. Substitution – the replacement of H2O ligands
3. Redox – when the metal ion changes oxidation state

Acid-Base / Hydrolysis Reactions

The addition of NaOH or NH3 results in the removal of H+ from water ligands until the overall charge of
the complex structure is 0.

For example:

[Cu(H2O)6]2+(aq) + 2OH–(aq) à M(H2O)4(OH)2(s) + 2H2O(l)

[Cu(H2O)6]2+(aq) + 2NH3(aq) à M(H2O)4(OH)2(s) + 2 NH4+(aq)

The resulting complex ions in the reactions have no overall charge. This makes them insoluble, and
precipitates result.

Metal hexa-aqua ions are acidic.

[M(H2O)6]2+ [M(H2O)6]3+
pH 6 3
GETTING TO THE NUB OF CHEMISTRY
This is because the charge density of the metal ion withdraws electron density from the O-H bonds in the
water molecule ligands thereby weakening them and releasing H+.

Metal hexa-aqua ions with a 3+ charge are more acidic (have a lower pH) than those with a 2+ charge. To
understand why, we need to look at the relevant equilibria.

The position of equilibrium lies further to the right in equation 1, which means a lower pH. This is
because the higher charge density of the 3+ hexa-aqua ion withdraws greater electron density from the
O-H bonds in the water molecules and weakens them to a greater extent.

As a result, 3+ complex ions react with CO32- to produce CO2 and effervescence is observed.

Some metal hydroxides show amphoteric character by dissolving in both acids and bases (e.g. hydroxides
of Al3+).

Metal (II) hexa-aqua ions react with CO32- to produce carbonate precipitates.

As we’ve seen, metal (III) hexa-aqua ions react with CO32- to produce CO2 AND a precipitate (in this case,
the hydroxide with no overall charge). This is because Metal (III) hexa-aqua ions have a lower pH.

Finally, certain metal hexa-aqua ions can undergo ligand substitution reactions.
GETTING TO THE NUB OF CHEMISTRY

Flowcharts and Reactions of AQA’s Ions


GETTING TO THE NUB OF CHEMISTRY

Test for… Reagent Observation Ionic Equation


Al 3+ NaOH / [Al(H2O)6]3+(aq) + 3OH¯(aq) à Al(H2O)3(OH)3 (s) + 3H2O(l)
[Al(H2O)6]3+ NH3 White precipitate
colourless
dropwise [Al(H2O)6]3+(aq) + 3NH3(aq) à Al(H2O)3(OH)3 (s) + 3NH4+(aq)
Al3+ White precipitate Al(H2O)3(OH)3 (s) + OH- à [Al(H2O)2(OH)4]- + H2O(l)
NaOH in
[Al(H2O)6]3+ dissolves à a
colourless excess Excess only: [Al(H2O)6]3+(aq) + 4OH¯(aq) à Al(H2O)2(OH)4]- + 4H2O(l)
colourless solution
Al3+ White precipitate
[Al(H2O)6]3+ Na2CO3 2[Al(H2O)6]3+(aq) + 3CO32-(aq) à 2Al(H2O)3(OH)3 (s) + 3H2O(l) + 3CO2(g)
colourless effervescence
Cu2+ NaOH / [Cu(H2O)6]2+ (aq) + 2OH¯ (aq) à Cu(H2O)4(OH)2 (s) + 2H2O (l)
[Cu(H2O)6]2+ NH3 Blue precipitate
blue
dropwise [Cu(H2O)6]2+(aq) + 2NH3 (aq) à Cu(H2O)4(OH)2 (s) + 2NH4+(aq)
Cu2+
NH3 in
[Cu(H2O)6]2+ Deep blue solution Cu(H2O)4(OH)2 (s) + 4NH3(aq) à [Cu(H2O)2(NH3)4]2+(aq) + 2H2O(l) + 2OH¯(aq)
blue excess

Cu2+
Blue-green
[Cu(H2O)6]2+ Na2CO3 [Cu(H2O)6 ]2+ (aq) + CO32-(aq) → CuCO3 (s) + 6H2O(l)
blue precipitate

Cu2+
[Cu(H2O)6]2+ HCl Green solution [Cu(H2O)6 ]2+(aq) + 4Cl-(aq) → [CuCl4]2-(aq) + 6H2O(l)
blue

Fe2+ NaOH / Green precipitate [Fe(H2O)6]2+ (aq) + 2OH¯ (aq) à Fe(H2O)4(OH)2 (s) + 2H2O (l)
[Fe(H2O)6]2+ NH3 (turns brown on
green
dropwise exposure to air) [Fe(H2O)6]2+ (aq) + 2NH3 (aq) à Fe(H2O)4(OH)2 (s) + 2NH4+(aq)
Fe2+
[Fe(H2O)6]2+ Na2CO3 Green precipitate [Fe(H2O)6 ]2+ (aq) + CO32-(aq) → FeCO3 (s) + 6H2O(l)
green

Fe2+
[Fe(H2O)6]2+ HCl Yellow solution [Fe(H2O)6 ]2+(aq) + 4Cl-(aq) → [FeCl4]2-(aq) + 6H2O(l)
green

Fe3+ NaOH / [Fe(H2O)6]3+ (aq) + 3OH¯ (aq) à Fe(H2O)3(OH)3 (s) + 3H2O (l)
[Fe(H2O)6]3+ NH3 Brown precipitate
purple
dropwise [Fe(H2O)6]3+ (aq) + 3NH3(aq) à Fe(H2O)3(OH)3 (s) + 3NH4+(aq)
Fe3+ Brown precipitate
[Fe(H2O)6]3+ Na2CO3 2[Fe(H2O)6]3+(aq) + 3CO32-(aq) à 2Fe(H2O)3(OH)3 (s) + 3H2O(l) + 3CO2(g)
purple and effervescence
Fe3+
[Fe(H2O)6]3+ HCl Yellow solution [Fe(H2O)6 ]3+(aq) + 4Cl-(aq) → [FeCl4]-(aq) + 6H2O(l)
purple
GETTING TO THE NUB OF CHEMISTRY
3.3.1 Introduction to Organic Chemistry
3.3.1.1 Nomenclature

Organic chemistry is the study of carbon compounds. The ability of carbon to form strong bonds to itself
and to hydrogen leads to the formation of stable compounds. Hydrocarbons contain carbon and
hydrogen only. Chains of up to thousands of carbons in length are possible. This means there are
potentially millions of organic compounds.

In organic chemistry, functional groups are specific groups of atoms or bonds within molecules that are
responsible for the characteristic chemical properties and reactions of these myriad molecules. The same
functional group will undergo the same or similar chemical reaction(s) regardless of the size of the
molecule it is a part of.

A homologous series is a family of compounds that all contain the same functional group. In a series,
each member differs by the addition of a CH2 group and there is a gradual change in physical properties.

In the formulas, the symbols R and R' usually denote an attached hydrogen, or a hydrocarbon side
chain of any length ( a.k.a. an alkyl group), but may sometimes refer to any group of atoms.

Homologous Functional
Prefix / Suffix Example Example’s name
series group

Alkane -ane N/A propane

bromo-
—X
choro-
Halogenoalkane (X represents 2-bromopropane
fluoro-
any halogen)
iodo-

Alkene -ene C=C propene

hydroxy-
Alcohol —O—H propan-2-ol
-ol

Aldehyde -al propanal


GETTING TO THE NUB OF CHEMISTRY

Ketone -one propanone

Carboxylic acid -oic acid propanoic acid

Ester -yl -oate methyl ethanoate

Acyl chloride -oyl chloride propanoyl chloride

amino-
Amine ethylamine
-ylamine

Amide -amide ethanamide

phenyl-
Aromatic methylbenzene
-benzene

amino- -oic 2-aminoethanoic


Amino acid
acid acid

Nitrile -nitrile propanenitrile


GETTING TO THE NUB OF CHEMISTRY
Types of Organic Formulae

Example Type Description

This only tells you how many


atoms of each element is present
Molecular formulae in each molecule. NEVER use
molecular formulae in your exam
unless it’s very clear you can.

This gives you the simplest whole


Empirical formulae number ratio of the different
types of atom in the substance.

Each carbon atom is listed


separately with the atoms
(Condensed) structural attached to it following. An
formulae exception is cyclic parts of
molecules e.g. benzene where the
carbons are grouped.

Not all the bonds are shown.


Carbon - hydrogen, oxygen-
Structural formulae
hydrogen and nitrogen-hydrogen
bonds are often simplified.

Displayed formulae Shows ALL of the bonds

Most hydrogen atoms are


omitted, and line ends and
vertices represent carbons.
Skeletal formulae
Functional groups and atoms
other than carbon and hydrogen
are still shown.
GETTING TO THE NUB OF CHEMISTRY
Naming Organic Compounds

1. Identify the longest unbranched chain containing the functional group to establish the stem
name.

Longest C-C
unbranched 1 2 3 4 5 6 7 8 9 10
chain

Prefix Meth- Eth- Prop- But- Pent- Hex- Hept- Oct- Non- Dec-

2. Name any substituent groups that are bonded to a carbon atom. If there are two or more do so
alphabetically.

Substituent Group Formula

Methyl- CH3

Ethyl- CH3CH2

Propyl- CH3CH2CH2

Butyl- CH3CH2CH2CH2

3. Identify the position of the substituents by numbering the longest C chain (that contains the
functional group) from the end that gives the functional group(s) the lowest numbers.

Extra notes

• If there are two identical groups on the same carbon the prefix ‘di-‘ is added and the number
repeated.

• If there are two or more identical groups on different carbons the prefixes ‘di-‘,’tri-‘ and ‘tetra’
are used.

• Commas separate numbers

• Dashes separate words

• Note that some functional groups are carbon-1 in the chain by definition and do not need a
number themselves (e.g. aldehydes, carboxylic acids and nitriles. These define which end the
numbers start for other groups.

• Note that we should only include numbers in names when they are necessary – but we usually
only omit numbers in the simplest molecules where it is obvious that no number is needed.
GETTING TO THE NUB OF CHEMISTRY

3.3.1.2 Reaction mechanisms

There are three important types of reaction in organic chemistry:

• Substitution – an atom, or group of atoms, is substituted by another.

• Addition – Two molecules react together to form a single molecule. A double bond is typically
involved.
GETTING TO THE NUB OF CHEMISTRY
• Elimination – a small molecule is removed (eliminated) from a larger one. This usually results in a
double bond being formed.

All reactions involve the breaking and making of bonds. The breaking of covalent bonds is often called
bond fission. Bonds can break in two ways:

• Homolytic fission – one of the two shared electrons in the covalent bond goes to each atom.
This results in chemical species possessing an unpaired electron called free radicals. There is no
overall charge because there is an equal number of protons and electrons. Free radicals are
highly reactive.

• Heterolytic fission – both of the shared electrons go to just one of the atoms when the bond
breaks. As a result, one atom becomes negatively charged because it has one more electron than
protons and the other becomes positively charged.

Many organic reactions involve nucleophiles and electrophiles.

• Nucleophiles are lone pair donors. Examples include :NH3 and :OH–

• Electrophiles are lone pair acceptors. Examples include HBr and NO2+

When reaction mechanisms are being described, curly arrows are used to show the movement of
electrons. A curly half-arrow is used to show the movement of a single electron. In both cases, the
beginning of the arrow shows where the electron(s) start from and the (half-)head of the arrow shows
where it/they ends up.

Reactions have a beginning, middle and end. A reaction equation shows only the first and third of these.
Reaction mechanisms allow us to understand the journey taken.
GETTING TO THE NUB OF CHEMISTRY
GETTING TO THE NUB OF CHEMISTRY
3.3.1.3 Isomers

Isomers are compounds with the same molecular formula but with their atoms arranged in different
ways. There are two distinct types of isomer:

• Structural isomers have the same molecular formula but a different structural formula.

• Stereoisomers have the same structural formula but their atoms are arranged differently in
space.

When answers questions asking how many isomers there are, skeletal formulae are extremely useful in
working these out.
GETTING TO THE NUB OF CHEMISTRY
How many different chain isomers of C5H12 are there?

Draw and name the position isomers of pentanol:

At A level, functional groups isomers often come in pairs. Here are the common ones:
GETTING TO THE NUB OF CHEMISTRY

When answering questions, it’s important to consider ALL possible isomers e.g. position AND chain.

Geometric isomerism, a form of stereoisomerism, arises in alkenes because of restricted rotation about
the C=C bond. Note, not all alkenes display this form of isomerism.
GETTING TO THE NUB OF CHEMISTRY
The simplest form of geometric isomerism is cis-trans isomerism.

However, the cis-trans label can only takes us so far. For example, which of the following geometric
isomers is ‘cis’ and which is ‘trans’?

To be able to name molecules such as the above, we use the Cahn-Ingold-Prelog rules prescribed by
IUPAC (International Union of Pure and Applied Chemistry). These assign priorities to each of the groups
attached to a C in the C=C bond.

1. Look at the first atoms attached directly to the carbon atoms at each end of the carbon carbon
double bond. The atom with higher atomic number is given the higher priority.

2. Sometime, the atoms being compared have the same atomic number. In this case, you look at
what’s attached to these and compare the priorities of these next atoms.

3. If the two groups of higher priority are on opposite sides (like trans – across from each other),
then we use the label E e.g. E-but-2-ene. If they’re on zee zame zide, then we use Z e.g. Z-but-2-
ene.
GETTING TO THE NUB OF CHEMISTRY
Let’s consider one of the isomers above.

Looking at the left-hand-side (LHS) carbon in the C=C, the atoms directly attached are C (in the CH3) and
Cl. The former has an atomic number of 6 and the latter 17. So, the Cl takes priority.

Looking at the right-hand-side (RHS) carbon in the C=C, the atoms directly attached are Br and F. The
former has an atomic number of 35 and the latter 9. So, the Br takes priority.

The groups with higher priority are across from each other and so the molecule’s name is:

E-1-bromo-2-chloro-1-fluoropropene

Let’s consider another example.

Starting with the left-hand-side (LHS) carbon in the C=C, both of the atoms directly attached are carbons
so we consider the atoms one bond further away. C (H and H) from the ethyl group has a higher atomic
number than H (H and H) from the methyl group. So, CH3CH2 group takes priority.

Looking at the right-hand-side (RHS) carbon in the C=C, the atoms directly attached are H and C. The
former has an atomic number of 1 and the latter 6. So, the CH2OH takes priority.

The highest priority groups, CH3CH2 and CH2OH are on zee zame zide of the C=C so the isomer is Z.

Z-3-methylpent-2-en-1-ol

Note: for there to be geometric isomerism, there MUST be two different groups attached to each
carbon in the C=C bond. For example, when C=C is at the end of the chain, 2 Hs will be attached to one
of the carbons and so geometric isomerism doesn’t occur.
GETTING TO THE NUB OF CHEMISTRY
Some Example Exam Questions
GETTING TO THE NUB OF CHEMISTRY
GETTING TO THE NUB OF CHEMISTRY
3.3.2 Alkanes
3.3.2.1 Fractional distillation of crude oil

Alkanes are saturated hydrocarbons. This means there are only single bonds between the carbon atoms.
Crude oil is a mixture consisting mainly of alkane hydrocarbons with different boiling points that can be
separated by fractional distillation.

The larger the Mr of the molecule, the more electrons there are and the greater the van der Waals forces
between the molecules. Remember, van der Waals forces because the random movement of electrons in
molecular substances can lead to temporary dipoles i.e. regions of these substances can become
temporarily d+ and d- (these symbols represent partial charges). While these exist (they are continually
appearing and disappearing), they can induce (cause) d+ and d- regions in adjacent molecules leading to
weak electrostatic attraction between them.

If the Mr of the different molecules is the same, then consider whether there’s any branching i.e. are
they chain isomers?

Straight-chain isomers have a greater contact surface area than branched isomer, which results in higher
boiling points. The more branched the isomer is, the smaller its contact surface area, the weaker the van
der Waals forces between the molecules and the lower its melting and boiling points.
GETTING TO THE NUB OF CHEMISTRY
So, the larger the molecule, the higher the boiling point. A substance’s boiling point is the same as its
condensation point. In fractional distillation:

• Crude oil is vaporised


• This vapour passes into a column with a temperature gradient – it’s hotter at the bottom and
cooler at the top.
• As the vapour rise, it cools.
• As the molecules in the mixture have different boiling points due to their different sizes, they
condense at different height in the column.
• The larger the molecule, the higher its boiling point, the lower down the column it condenses.
• Fractions of the original crude oil are collected – a fraction is a mixture of hydrocarbons with
similar boiling points.

3.3.2.2 Modifications of alkanes by cracking

There is a greater or lesser demand for the different fractions collected in Fractional Distillation. Larger
alkanes, which are typically less in demand, can be ‘cracked’ to produce smaller, more useful (valuable)
molecules. This involves the breaking of C-C bonds and is an example of a decomposition reaction. For
example:
GETTING TO THE NUB OF CHEMISTRY
There are two type of cracking, thermal and catalytic, which produce different products.

Thermal cracking Catalytic cracking


Temperature 900oC 450oC
Pressure 70 atm 1-2 atm
Catalyst N/A Zeolites
Products Alkenes Fuels (aromatics, cyclic alkanes, branched alkanes)

3.3.2.3 Combustion of alkanes

It’s the combustion of alkanes that provides their most important uses. These are very exothermic
reactions. This combustion can be complete or incomplete.

The products of complete combustion are CO2 and H2O. E.g.

CH4(g) + 2O2(g) à CO2(g) + 2H2O(l) DH = -890 kJ mol-1

The products of incomplete combustion are CO and/or C and H2O – these reactions are less exothermic
and so much less desirable when the alkane is being used as a fuel.

!
CH4(g) + O2(g) à CO(g) + 2H2O(l)
"

CH4(g) + O2(g) à C(s) + 2H2O(l)

Furthermore, CO is toxic (don’t use the word poisonous in an exam) and C causes respiratory problems.

Many combustion reactions involving alkanes take place in engines. Other reactions can also take place
here.

Nitrogen is usually an inert (unreactive) gas due to the large bond enthalpy of its triple bond. However, in
can react with oxygen when in the high temperatures present in an internal combustion engine e.g. a car
engine, to produce nitrous oxides. E.g.

N2(g) + O2(g) à 2NO(g)

Fuels often have sulfur containing compounds in them. This can result in SO2 being produced when
they’re burned.

Finally, some fuels may fail to be combusted, which wastes fuel.


GETTING TO THE NUB OF CHEMISTRY
Pollutant How is it formed? What problems does How can the problem be
this cause? reduced?
Complete combustion of Greenhouse gas à
CO2 Burn fewer fossil fuels
carbon containing fuels climate change
Incomplete combustion of Toxic / less energy is Ensure a good supply of
CO
carbon containing fuels released oxygen when burning fuels
Respiratory
Incomplete combustion of problems, global Ensure a good supply of
C (soot)
carbon containing fuels dimming / less oxygen when burning fuels
energy is released
Combustion of hydrogen
H2O N/A
containing fuels
Remove S from fuel before
Combustion of sulfur
SO2 Acid rain burning, flue gas
containing fuels
desulfurisation
The reaction of N2 in the air
NOx (NO, NO2) with O2 in the air at very high Acid Rain Use catalytic converters
temperatures (in engines etc.)
Unburned Ensure a good supply of
Fuels that doesn’t get burned Wasted fuel
hydrocarbons oxygen when burning fuels

SO2 can be removed from waste gases by flue gas desulfurisation – calcium oxide or calcium carbonate
react with the SO2.
CaO + SO2 à CaSO3
CaCO3 + SO2 à CaSO3 + CO2

CO and NOx gases can be removed from exhaust gases using catalytic converters.

2CO + 2NO à2CO2 + N2


GETTING TO THE NUB OF CHEMISTRY
3.3.2.4 Chlorination of alkanes

Under special conditions (uv light), alkanes can be made to react with a halogen to produce a
halogenoalkane. This is an example of the mechanism free radical substitution.

The mechanism itself involves three steps: initiation, propagation and termination.
GETTING TO THE NUB OF CHEMISTRY
Let’s take a closer look at the propagation step:

Now let’s take an even closer look at the propagation steps.


GETTING TO THE NUB OF CHEMISTRY

Some example exam questions


GETTING TO THE NUB OF CHEMISTRY
GETTING TO THE NUB OF CHEMISTRY
3.3.3 Halogenoalkanes
Halogenoalkanes are much more reactive than alkanes. Key to their reactivity is the carbon-halogen
bond (C–X).

Except for iodine, all halogenoalkanes are more electronegative than carbon. This results in a polar bond.

Even though C–I doesn’t have a permanent dipole, the bond is very easily polarised by anything
approaching it.

Bond strength is also an important factor in halogenoalkane reactivity and determines their relative
reactivity when compared to each other.

The C–F bond is so strong that fluoroalkanes are very unreactive.

3.3.3.1 Nucleophilic Substitution

Halogenoalkanes are capable of undergoing substitution reactions with the nucleophiles :OH–, :CN– and
:NH3 to produce alcohols, nitriles and amines.
GETTING TO THE NUB OF CHEMISTRY
Here’s the mechanism when the nucleophile (electron pair donor) is :OH–

The mechanism is the same with the nucleophile :CN–


GETTING TO THE NUB OF CHEMISTRY
When the nucleophile is :NH3, there are a few additional features.

3.3.3.2 Elimination

Elimination is the removal of a simple molecule from adjacent atoms to form a carbon-carbon double
bond (C=C). Under different conditions, mainly to do with the solvent, :OH– can act as a base (a proton
(H+) acceptor) instead of as a nucleophile.
GETTING TO THE NUB OF CHEMISTRY
Nucleophilic substitution vs Elimination
in halogenoalkanes

Elimination throws up the possibility of more than one product being formed when the halogenoalkane
is asymmetric about the carbon-halogen bond i.e. the molecule has different length carbon chains either
side of the C-X bond. For example, 2-bromobutane:

The product formed will depend on which carbon the hydrogen is removed from.
GETTING TO THE NUB OF CHEMISTRY
Watch out! Elimination can’t take place if the carbon(s) adjacent to the C-X bond doesn’t have an H on it.

Halogenoalkanes and Rates of Reaction

The rate of nucleophilic substitution depends on which halogen is present. As we saw at the start of this
section, carbon-halogen bonds have different strengths: C-F > C-Cl > C-Br > C-I. As a result, iodoalkanes
substitute faster than bromoalkanes, and bromoalkanes substitute faster than chloroalkanes.

Hydrolysing halogenoalkanes in the presence of acidified silver nitrate shows how the rate of reaction
varies with the halogen.

For example, when the halogenoalkanes CH3CH2Cl, CH3CH2Br and CH3CH2I react with silver nitrate in
ethanol solution, the yellow colour of silver iodide forms faster than the cream colour of silver bromide,
which itself forms faster than the white colour of silver chloride.

3.3.3.3 Ozone depletion

Ozone, formed naturally in the upper atmosphere, is beneficial because it absorbs ultraviolet radiation.
Chlorine free radical atoms are formed in the upper atmosphere when ultraviolet radiation causes C–Cl
bonds in chlorofluorocarbons (CFCs) to break.

Note: The C-F bond is too strong to undergo homolytic fission due to uv light.

These free radicals catalyse the decomposition of ozone and contribute to the hole in the ozone layer.

Because the Cl• is regenerated, one Cl• can destroy many, many O3 molecules. This is known as a chain
reaction. In recent years, CFCs have been replaced by fluorocarbons, which don’t produce free radicals.
GETTING TO THE NUB OF CHEMISTRY
Some Example Exam Questions
GETTING TO THE NUB OF CHEMISTRY
GETTING TO THE NUB OF CHEMISTRY
GETTING TO THE NUB OF CHEMISTRY
Solubility in Organic Chemistry
Most organic reactions are carried out in solution. The solvent needs to dissolve the reactants to
allow the species to be in the same phase so that they can collide and react (the solvent also helps
absorb and dissipate the energy released).

A helpful simple rule is ‘like dissolves like’ i.e. polar molecules tend to dissolve in polar solvents (e.g.
water, alcohols) and non-polar molecules in non-polar solvents (e.g. hexane). This is why ionic
compounds like sodium chloride or compounds like sugar, dissolve in water but do not dissolve to any
great extent in most organic solvents. It also applies to the separation of oil and water i.e. they are
immiscible.

The polarity of organic molecules is determined by the presence of polar bonds due to the presence of
electronegative atoms (e.g. N, O) in polar functional groups such as amines (-NH2) and alcohols (-OH).
Overall, polarity is a balance of the non-polar parts and polar parts of a molecule, and it can be
convenient to mentally break a molecule into its polar and non-polar components. A larger non-polar
‘organic’ hydrocarbon part will tend to make the molecule less soluble in polar solvents while more
polar parts will tend to make it more water-soluble (polar solvent) and vice versa for non-polar solvents.

Ethanol is a suitable solvent for the reaction between a halogenoalkane and NaOH as it’s capable of
dissolving both reactants. Hexanol would be less appropriate because of its much larger non-polar part.

Carboxylate salts of soaps form micelles, spherical structures in which the polar carboxylate is oriented on
the outside diameter of the sphere and the non-polar hydrocarbon chains of the fatty acids are oriented into
the centre of the sphere. Dirt and oil associate with the non-polar tails of the carboxylate salts (‘like dissolves
like’) through hydrophobic interactions and are trapped in the centre of the sphere. Since the outer surface
of the micelle is polar, it is water soluble and can be washed or rinsed out with water (polar solvent).
GETTING TO THE NUB OF CHEMISTRY
3.3.4 Alkenes
3.3.4.1 Structure, bonding and reactivity

In alkenes, the high electron density of the carbon–carbon double bond leads to attack on these
molecules by electrophiles (electron pair acceptors).

3.3.4.2 Addition reactions of alkenes

Before we look carefully at the mechanism, we first need to understand what a carbocation is.
Carbocations are molecules that carry a positive charge on a carbon atom.

Carbocations can be classed at primary (1o), secondary (2o) or tertiary (3o). In a primary carbocation, the
carbon that carries the positive charge is only attached to one other carbon chain.

In a secondary carbocation, the positively charged carbon is attached to two carbon chains.

And tertiary is…


GETTING TO THE NUB OF CHEMISTRY
There is a difference in their relative stability:

The inductive effect is the reason for this difference in stability.

Carbon is not very electronegative. This means that many functional groups attract electrons more than
carbon, which means they withdraw electrons from the carbon atom they’re bonded to.

This is called the negative inductive effect and is a key feature of nucleophilic mechanisms.

Alkyl groups do precisely the opposite. They are in effect ‘electron donating’ or ‘electron pushing’
groups. As a result, we get a positive inductive effect.

Anything that ‘donates’, or ‘pushes’, electron density towards a positively charged carbon will help
stabilise it.

We can now look at the electrophilic addition mechanism that alkenes undergo. Here it is when an
alkene reacts with a hydrogen halide to produce a halogenoalkane.
GETTING TO THE NUB OF CHEMISTRY
It’s with asymmetrical alkenes that the relative stability of carbocations comes into play.

The second of these happens more readily. This is because, with the second, a secondary carbocation is
formed, which is more energetically favourable than the primary one formed in the first. This results in
there being major and minor products.

2-chlorobutane is the major product because it is formed via a secondary (2o) carbocation, which is more
stable than the primary (1o) carbocation 1-chlorobutane is formed via. This is due to the positive
inductive effect of two alkyl groups as opposed to one.

The standard test for an alkene is bromine water, which then goes from orange to colourless. Although
Bromine is a non-polar molecule, a dipole is temporarily induced as Br2 gets close the electron-rich
double-bond in an alkene as this repels the electron pair in the Br-Br bond.
GETTING TO THE NUB OF CHEMISTRY
Alkenes can also react with sulfuric acid.

3.3.4.2 Addition polymers

Addition polymers are formed from alkenes and substituted alkenes. When polymerisation takes place,
the double bonds in the separate monomers break, allowing the monomers to join up.

The name of the polymer above is polyethene (the part after poly is the name of the original monomer).

A reaction equation for addition polymerisation would be:


GETTING TO THE NUB OF CHEMISTRY
Questions can ask you to draw the repeating unit from a section of the polymer chain or draw the
structure of the monomer from a section of the polymer. For example, if we were given the following
section of a polymer polypropene…

Then the repeating unit would be…

And the monomer would have been…

Similarly, you could be asked to draw the repeating unit from a monomer structure. For example, if we
were given the following monomer polyphenylethene (also known as polystyrene)…

The repeating unit would be…


GETTING TO THE NUB OF CHEMISTRY
Addition polymers are famously non-biodegradable. This is because their C-C bond are non-polar and
therefore aren’t susceptible to attack from nucleophiles.

All addition polymers are solid at room temperature. This is because there are strong van der Waals
forces between the large, long chains.

Some addition polymers are particularly hard and rigid. In the case of PVC (polychloroethene), this is
because of dipole-dipole interactions between the chains in addition to van der Waals forces.

For certain functions, this is perfect. However, they can be made more flexible through the addition of a
plasticiser. These reduce the effectiveness of these attractions by getting between the chains making the
plastic more flexible. The more plasticiser you add, the more flexible it becomes.

Poly(chloroethene) is used to make a wide range of things including guttering, plastic windows, electrical
cable insulation, sheet materials for flooring and other uses, footwear, clothing, and so on and so on.
GETTING TO THE NUB OF CHEMISTRY
3.3.5 Alcohols
3.3.5.1 Alcohol production

Alcohols are produced industrially by hydration of alkenes in the presence of an acid catalyst. The
mechanism for the formation of an alcohol by the reaction of an alkene with steam in the presence of an
acid catalyst is…

Ethanol is produced industrially by fermentation of glucose and is separated by fractional distillation. The
product can then be used as a biofuel. It is a biofuel because it is made from plants that are renewable.

C6H12O6 à 2CH3CH2OH + 2CO2

One of great advantages of using ethanol as a fuel, besides it being renewable, is that it’s considered
‘carbon neutral’. By this we mean that it releases the same amount of CO2 when it is burned as was
taken in from the atmosphere when the plants grew and photosynthesised.

Photosynthesis in crops: 6CO2 + 6H2O à C6H12O6 + 6O2 Uses up 6CO2


Fermentation to make ethanol: C6H12O6 à 2CH3CH2OH + 2CO2
Burning ethanol: 2CH3CH2OH + 6O2 à 4CO2 + 6H2O
Overall, for fermentation + burning: C6H12O6 + 6O2 à 6CO2 + 6H2O Releases 6CO2

However, in the process of going from crop to using the fuel, there are other processes to consider
including fuel for machinery and transport meaning that it is not 100% carbon neutral. Plus, there are
growing concerns that too much agricultural land is being used for growing crops to make biofuels rather
than food.
GETTING TO THE NUB OF CHEMISTRY
Fermentation Reaction of Ethene with Steam
Equation C6H12O6 à 2CH3CH2OH + 2CO2 CH2=CH2 + H2O à CH3CH2OH
Conditions
Temperature: 35oC
300 oC
Pressure: Normal atmospheric pressure
70 atm
Catalyst Enzyme in yeast
Conc. Phosphoric Acid
Other: Aqueous, anaerobic
Raw materials Carbohydrate crops (e.g. sugar, maize) Crude oil
Raw material type Renewable Non-renewable (finite)
Type of process Batch (stop-start) Continuous
Reaction rate Slow Fast
Impure (must be purified by fractional
Purity of the ethanol Pure
distillation

Which… Fermentation Hydration of steam


…is the fastest? ✓
…uses renewable raw materials ✓
…takes place at lower

temperatures and pressures?
…gives purer ethanol? ✓
…is a continuous process? ✓
…has lower labour costs? ✓

Some alcohols contain more than one -OH group. They are named using suffixes -diol, -triol etc.

Short chain alcohols, such as ethanol, tend to be good solvents for polar as well as non-polar solutes.
This is because they contain both a highly polar -OH group and a non-polar hydrocarbon chain. For
example, both sodium hydroxide (ionic) and hexane (molecular) dissolve well in ethanol.

In addition, the -OH group in alcohols gives them low volatility and high miscibility with water because of
hydrogen bonding. The influence of the -OH group becomes less as the carbon chain gets longer.
GETTING TO THE NUB OF CHEMISTRY
3.3.5.2 Oxidation of alcohols

As with carbocations, alcohols can be classified as primary (1o), secondary (2o) and tertiary (3o).

Depending on whether they’re primary (1o), secondary (2o), alcohols can be oxidised to produce
aldehydes, ketones or carboxylic acids.

Functional group General structure Example Name

Aldehyde propanal

Ketone pentan-2-one

Carboxylic acid butanoic acid

The oxidising agent in these reactions is typically acidified potassium dichromate (VI), which is reduced
to Cr3+. A colour change from orange to green is observed. This is the relevant half-equation for this
reduction:

Primary (1o) alcohols can be oxidised initially to an aldehyde and then to a carboxylic acid.
GETTING TO THE NUB OF CHEMISTRY
When writing the reaction equation for the above, [O] is used as shorthand to represent the oxidising
agent.

Going straight to the carboxylic acid, the equation would be:

It’s very hard to control the reaction so that just the aldehyde that is produced. The aldehyde needs to
be separated from the reaction mixture as it’s formed. This can be achieved because of the difference in
boiling point of alcohols, aldehydes and carboxylic acids because of their relevant intermolecular forces.

Secondary (2o) alcohols can only be oxidised to a ketone and no further.

Tertiary (3o) alcohols can’t be readily oxidised, which means acidified potassium dichromate can be used
as a test for 1o and 2o alcohols.

Orange à Green
Reduction 6e + 14H+ + Cr2O72-
- à 2Cr3+ + 7H2O
Oxidation 1o / 2o alcohol à Aldehyde or Carboxylic Acid / Ketone
3o alcohol X No reaction
GETTING TO THE NUB OF CHEMISTRY
Intermolecular forces and organic molecules…

Type Key thoughts Relevant to… Disco lights analogy


van der Waals Temporary / instantaneous Every molecule! The lights are constantly
dipoles are formed due to Alkanes, alkenes and flashing and hence on
constantly moving electrons aromatic compounds with temporarily. However, they
and these, in turn, induce only alkyl substituents flash at different times so the
temporary dipoles in adjacent feature these forces dance floor is always
molecules. ONLY. illuminated, albeit dimly.
Greater Mr à greater number A greater number of flashing
of electrons à stronger vdW lights (a bigger Mr) means
forces. the floor is more illuminated.
Straight-chain > branched due
to there being a greater contact
surface area.
Permanent dipole- These arise when atoms with • Acyl chlorides Some of the lights are
dipole different electronegativities are • Aldehydes and ketones permanently switched on
covalently bonded resulting in a • Halogenoalkanes while the others continue to
polar bond AND there is a net • 3o amines flash. As a result, the dance
dipole. The shape of the • Polyesters floor is more brightly lit.
molecule is typically very
important when determining if
there’s the latter.
Hydrogen-bonding These arise when hydrogen is • Water and ammonia As with PDD, some of the
bonded to fluorine, oxygen or • Alcohols lights are permanently
nitrogen (the three most • Ammonia switched on. However,
electronegative elements). This • 1o and 2o amines brighter bulbs are being used
is really a type of permanent • Amides and so the dance floor is
dipole-dipole interaction but its • Carboxylic acids even more brightly lit.
greater strength results in it • Hydrogen fluoride
having its own classification. • Polyamides and DNA

There are only four shapes of molecules in organic chemistry…

Shape Diagram When relevant IMF (all have vdW)

Whenever C has four PDD is C is bonded to F, Cl or


Tetrahedral
single bonds Br à halogenoalkanes

Whenever N has three H-bonding when 1o or 2o


Trigonal pyramidal single bonds i.e. amines amine or amide with H
and amides bonded to N. Otherwise PDD.
Around any carbonyl
group, or a C with two PDD when it’s a carbonyl e.g.
Trigonal planar single bonds and one aldehyde, ketone, ester and
double bond or a acyl chloride.
carbocation.
Around any O with two PDD – esters and ethers. H-
Bent single bonds i.e. alcohols bonding when it’s alcohols
and carboxylic acids. and carboxylic acids.
GETTING TO THE NUB OF CHEMISTRY
GETTING TO THE NUB OF CHEMISTRY

There are two main chemical tests to distinguish between aldehydes and ketones. Both are REDOX
reactions in which the aldehyde is oxidised (a positive result). Relevant reaction equations will use [O] to
represent the oxidising agent. Ketones cannot be oxidised.

The first involves Tollen’s Reagent:

1. Place about 1 cm3 of silver nitrate solution in a clean test tube.


2. Add one drop of sodium hydroxide solution to form a precipitate of silver oxide.
3. Add dilute ammonia drop by drop until the brown precipitate just redissolves (this is now
Tollen’s reagent).
4. Add a few drops of the organic compound.
5. Heat in a hot water bath for a few minutes. A silver mirror indicates the presence of an aldehyde.

Silver mirror
+ - à
Reduction [Ag(NH3)] (aq) + e Ag(s) + 2NH3 (aq)
Oxidation Aldehyde à Carboxylic acid
Ketone à No reaction

The second involves Fehling’s Solution:

Blue solution Red precipitate


Reduction 2Cu2+(aq) + 2OH- + 2e- à Cu2O + H2O
Oxidation Aldehyde à Carboxylic acid
Ketone à No reaction
GETTING TO THE NUB OF CHEMISTRY
3.3.5.3 Elimination

Alkenes can be formed from alcohols by acid-catalysed elimination reactions. This is also know as a
dehydration reaction. Alkenes produced by this method can be used to produce addition polymers
without using monomers derived from crude oil.

If the alcohol is 2o or 3o and has a chain length of four carbons or more, then it’s possible to get three or
four alkene products.

Elimination can’t take place if the carbon(s) adjacent to the C-OH bond doesn’t have an H attached to it.
GETTING TO THE NUB OF CHEMISTRY
3.3.6 Organic Analysis
3.3.6.1 Identification of functional groups by test-tube reactions

Functional Group Reagent Observation


1o / 2o Alcohol Acidified Potassium Dichromate Orange à Green
Aldehyde Tollen’s reagent Silver Mirror
Aldehyde Fehling’s solution Blue solution à Red precipitate
Carboxylic acid Sodium Carbonate Fizzing / effervescence
Alkene Bromine water Orange à colourless (decolourises)

3.3.6.2 Mass Spectrometry

Mass spectrometry was first looked at to determine the relative atomic masses and relative abundances
of isotopes. However, it can also be used to analyse molecules, particularly organic molecules.

The molecular ion peak is produced by the loss of just one electron from the complete molecule.

M à M+. + e-

So, the mass at which the molecular ion peak is found represents the relative formula mass (Mr) of the
molecule. This is the peak with the greatest m/z value.
GETTING TO THE NUB OF CHEMISTRY
In the case of compounds such as halogenoalkanes, there will be more than one significant molecular ion
peak due to the high abundance of the different isotopes of bromine (79Br : 81Br = 1:1 and 35Cl : 37Cl = 3:1)

High resolution mass spectrometers measure the m/z values to enough accuracy to find the molecular
formula.

If the Mr of a compound is measured to several decimal places, based on the most common isotopes), it
gives a unique molecular formula.

Low-res Mr Hi-res Mr Molecular formula Possible molecules


CH3COOH (ethanoic acid)
60 60.0211 C2H4O2
HCOOCH3 (methyl methanoate)
CH3CH2CH2OH (propan-1-ol)
60 60.0575 C3H8O CH3CH(OH)CH3 (propan-2-ol)
CH3OCH2CH3 (methoxyethane)
60 60.0324 CH4N2O H2NCOONH2 (urea)

However, there is often more than one compound with the same molecular formula. This means that
while high resolution mass spectrometry gives the molecular formula, it may not identify the compound.
For example, if the Mr is found to be 60.0575, and so the molecular formula is C3H8O, it does not tell us
whether the compound is propan-1-ol, propan-2-ol or methoxyethane.
GETTING TO THE NUB OF CHEMISTRY
GETTING TO THE NUB OF CHEMISTRY
3.3.6.3 Infrared Spectroscopy

All organic molecules absorb energy in the infrared region of the electromagnetic spectrum. The IR
wavelengths they absorb correspond to natural frequencies at which vibrating bonds in the molecules
stretch and bend. The bonds that absorb most strongly as they vibrate are polar covalent bonds such as
O-H, C-O and C=O (greenhouse gases have polar bonds!).

Bonds such as C=O and O-H vibrate in characteristic ways and absorb specific wavelengths. This means
they can be easily identified.

Below 1500 cm-1 is called the fingerprint region, which is unique to EVERY compound. This region can be
compared with a database

Putting it all together…


GETTING TO THE NUB OF CHEMISTRY
GETTING TO THE NUB OF CHEMISTRY
Optical isomerism 3.3.7
Stereoisomerism is when molecules have the same structural formulae but their atoms are arranged
differently in space. There are two types of stereoisomerism: geometric and optical.

Compounds that contain an asymmetric carbon atom form optical isomers, which have the same
structure but are mirror images of each other. Substances that show optical isomerism exist as two
isomers known as enantiomers.

They are so called because of their effect on plane polarised light.

Optical isomers are non-superimposable i.e. you can’t slide one isomer exactly over the other. Hence,
they’re stereoisomers – molecules with the same structural formula but with a different spatial
arrangement.

Optical Isomers have no plane of symmetry and are described as chiral. A carbon atom is described as a
chiral centre if it has four different groups attached – this makes it asymmetric.

Some examples:
GETTING TO THE NUB OF CHEMISTRY
Learn how to determine and draw optical isomers. Draw a tetrahedral shape and add the four different
groups. One will be an H; another will likely be CH3. The third will either be a functional group or a
branch with it attached e.g. CN, CHO, CH2OH etc. The fourth will be whatever is left. For example, draw
an optical isomer of an alcohol will the molecular formula C4H10O.

To determine whether a substance rotates plane polarised light, we us an instrument called a


polarimeter.

Initially light is sent through a sheet of Polaroid and the light which comes through the other side is
‘polarised light’ – all the waves are vibrating in the same plane.

Optical isomers rotate the plane of polarisation – one enantiomer rotates the plane clockwise and the
other enantiomer rotates the plane anticlockwise. As the plane polarised light travels into the sample
containing one enantiomer of the optical isomers, the plane of polarised light is rotated (5).
GETTING TO THE NUB OF CHEMISTRY
When optical isomers are made in the lab, they're often in 50:50 proportions of the two enantiomers –
this is known as a racemic mixture. Here’s how this can happen in an electrophilic addition reaction.

The two enantiomers rotate plane-polarised light to the same extent but in opposite directions. As they
exist in equal proportions, there is no net rotation. The rotations cancel out because two isomers are
present in equal amounts. So, racemic mixtures have no effect on plane polarised light.

Optical isomers are non-superimposable i.e. you can’t slide one isomer exactly over the other. Hence,
they’re stereoisomers – molecules with the same structural formula but with a different spatial
arrangement.
Optical Isomers have no plane of symmetry and are described as chiral. A carbon atom is described as a
chiral centre if it has four different groups attached – this makes it asymmetric.
Some real-life examples chiral molecules include:

• All amino acids except glycine. E.g. alanine.

• Thalidomide: When first released, thalidomide was promoted for anxiety, trouble sleeping, ‘tension,
and morning sickness.
GETTING TO THE NUB OF CHEMISTRY
Aldehydes and Ketones 3.3.8
This is the bonding, and shape around, the carbonyl group in organic chemistry.

Aldehydes can be reduced to primary alcohols, and ketones to secondary alcohols, using NaBH4(aq)
(sodium borohydride) or LiAlH4(dry ether) (lithium aluminium hydride). These provide a source of :H–
(hydride) ions, which acts as a nucleophile. The reducing agent is then followed by the addition of a
dilute acid, which provides H+ ions.

These reduction reactions are examples of nucleophilic addition.

Aldehydes and ketones can also undergo nucleophilic addition reactions with :CN- (from KCN), followed
by dilute acid, to produce hydroxynitriles. These provide the H+ and :CN- ions needed for the reaction.
HCN isn’t used because it’s a very toxic gas.

Notice that this adds a carbon atom the original structure.


GETTING TO THE NUB OF CHEMISTRY
If an asymmetric ketone is reduced, optical isomers are formed.

If aldehydes or asymmetric ketones undergo nucleophilic addition to form hydroxynitriles, optical


isomers are formed.

When optical isomers are formed in a nucleophilic addition reaction, the proportion of the two isomers
is 50:50 i.e. a racemic mixture is formed. This is because of the way the molecule is attacked by the
nucleophile. Molecules with C=O have a trigonal planar shape around the carbon atom in the C=O bond.
This results in an equal chance that the nucleophile will attack the δ+ carbon from above or below.

Because the mixture is racemic (in 50:50 proportions), it will not rotate plane polarised light despite
comprising optical isomers.
GETTING TO THE NUB OF CHEMISTRY
Extra: Alkenes à Optical Isomers àRacemic Mixtures
Optical isomers are also be formed whenever asymmetric halogenoalkanes are formed from alkenes.

However, as with aldehydes and asymmetric ketones, a racemic mixture if produced. This is because
carbocations are also trigonal planar round the carbon being attacked and there is an equal probability of
attack from above and below by the halide ion (i.e. nucleophile).
GETTING TO THE NUB OF CHEMISTRY
Carboxylic Acids and Derivatives 3.3.9
Carboxylic Acids and Esters 3.3.9.1

Carboxylic acids can be made through the oxidation of a primary (1o) alcohol or aldehyde using acidified
potassium dichromate.

When writing the reaction equation for the above, [O] is used as shorthand to represent the oxidising agent.

Going straight to the carboxylic acid, the equation would be:

In 3.3.5 Alcohols, we learned about the difference in the relative strengths of the intermolecular forces in
comparable alcohols, aldehydes and carboxylic acids and how these affected the practical
equipment/technique used depending on the desired product.

The hydrogen bonding in carboxylic acids is particularly strong, as these boiling points show.

Ethanol (Mr = 46) Ethanal (Mr = 44) Ethanoic acid (Mr = 60)
o
Boiling point ( C) 78 20* 118
*this volatility is why the glassware we collect the aldehyde in during distillation is cooled in ice.

The boiling points of carboxylic acids are higher than would be expected by considering their Mr only; two
hydrogen bonds form between molecules, holding them together in a dimer.
GETTING TO THE NUB OF CHEMISTRY
Carboxylic acids are the most important class of organic acids.

X-ray diffraction studies of the carboxylate ion show that the two carbon-oxygen bonds are of equal length.
This suggests the negative charge is distributed evenly over the carboxylate group and is delocalised.

This delocalisation means the carboxylate is less likely to join up with the H+ ion again (the carboxylate ion is
made more stable). This is why carboxylic acids are more acidic than alcohols.

Bases neutralise carboxylic acids to form salts. You need to remember the general equations for these
reactions (notice that the convention is to put the metal at the end of the salt’s formula in these instances).

The test for carboxylic acids is to add a metal carbonate of hydrogen carbonate – usually sodium carbonate
(Na2CO3) or sodium hydrogencarbonate (NaHCO3). Effervescence is seen.
GETTING TO THE NUB OF CHEMISTRY
Esters have the functional group:

R1 can be an H but R2 cannot (otherwise it’d be a carboxylic acid).

Esters can be named by reading them as alkyl derivatives of carboxylic acids. Thus. the name is obtained
from a stem indicating the alcohol from which the ester is derived with the suffix -yl, followed by a stem
indicating the acid, with the suffix -oate.

The most common way of making an ester (esterification) is through the reaction of an alcohol and
carboxylic acid.

The reaction of a carboxylic acid and alcohol to form an ester is fairly slow, even in the presence of an acid
catalyst (a few drop of conc. H2SO4). It is also an equilibrium reaction.
GETTING TO THE NUB OF CHEMISTRY

Another way of making an ester is through the reaction of an alcohol and acyl chloride.

The reverse of esterification is hydrolysis (again involving an acid catalyst). Note that equilibrium can be
reached from either side of the equation and will be the same regardless.
GETTING TO THE NUB OF CHEMISTRY
You can also perform the hydrolysis of esters using NaOH.

This reaction goes to completion – the carboxylic acid formed by hydrolysis reacts with excess alkali (NaOH)
to form the carboxylate salt. This removes the carboxylic acid from the equilibrium mixture as it’s formed,
which means the ester hydrolysis proceeds to completion. This cannot happen when an acid catalyst is used.

The reaction is still slow, and the mixture must be boiled under reflux.

Vegetable oils and animal fats are esters of propane-1,2,3-triol (glycerol). These can be hydrolysed in alkaline
conditions to give soap (salts of long-chain carboxylic acids) and glycerol.

Biodiesel is a mixture of methyl esters of long-chain carboxylic acids, which can be produced by reacting
vegetable oils with methanol in the presence of a catalyst.

Oils and Fats can be hydrolysed with water although this is a slow reaction.
GETTING TO THE NUB OF CHEMISTRY
Acylation 3.3.9.2

The substitution of an H atom for RC=O is called acylation. There are two main types of acylating agents:

Like halogenoalkanes, acyl chlorides (and acid anhydrides) are important in synthesis. Just as R-X is used to
attach an alkyl group to a nucleophile, RCOCl is used to attach an acyl group, RCO, to make amides or esters.
GETTING TO THE NUB OF CHEMISTRY

Acyl Chloride vs Acid Anhydride: there are industrial advantages to using acid anhydrides over acyl
chlorides. This is because acid anhydrides are:

• Cheaper
• Less corrosive
• Less dangerous to use
• Do not produce toxic HCl fumes
• Result in an easier to control reaction
• Undergo hydrolysis less readily

When water is added to ethanoyl chloride, which is a liquid, a violet reaction occurs. The liquid boils and
fumes of hydrogen chloride are evolved.

With acid anhydrides, we get a much less vigorous reaction.


GETTING TO THE NUB OF CHEMISTRY
The mechanism involved in these reactions is nucleophilic addition-elimination.
GETTING TO THE NUB OF CHEMISTRY
Extension: The effect of neighbouring groups on carboxylic acid strength.

The strengths of weak acids can be compared using Ka

The larger the value of Ka, the stronger the acid.

Nitric acid Ka = 40 mol dm-3


Ethanoic acid Ka = 1.7 x 10-5 mol dm-3

So, nitric acid is a stronger acid than ethanoic acid.

In the case of carboxylic acids, you are breaking an oxygen-hydrogen bond. Two of the factors that
influence the ionisation of carboxylic acids are:

• the strength of the bond being broken


• the stability of the ion being formed.

Different carboxylic acids have different strengths depending on what their make-up is aside from the COOH
group.

Carboxylic Acid Ka
CHCl2COOH 5.1 x 10-2
CH2ClCOOH 1.4 x 10-3
CH2ICOOH 6.8 x 10-4
HCOOH 1.8 x 10-4
CH3COOH 1.7 x 10-5

Generally, the more polarised the O-H bond is, the weaker and easier it is to break. The presence of
electron-withdrawing and donating groups affects this polarisation.

So, chloroethanoic acid is a stronger acid when compared with ethanoic acid, which the table data confirms.
GETTING TO THE NUB OF CHEMISTRY
In general, the higher the density of negative charge on the RCOO– ion, the more likely it will attract and
reform the unionised RCOOH. In other words, the more negative the charge density on the RCOO– ion, the
weaker the acid will be and vice versa.
Recrystallisation
This is a
very standard
method for purifying a solid
by removing both

and insoluble impurities


got trapped first
soluble that likely in
your
batch of crystals . insoluble
-
impurities
are left behind
-
#

① ② ③
⑪ ⑪

solid
filterbygraystation
dissolve in
impure
Crude product) miniumum amount

of hot
-
solvent
crystals
saturated Buchner funnel collected
- Iwash with
solvent)
⑭ ⑤ ⑥
cold

filter paper
RECRYSTALLISATON Filter
⑪ ⑪ Vacuum

loo ·
under pump
pressure

soluble
Allow filhate t impurities
-

the
---

infiltrate
-

Vacuum/suction
cool to room temperature
filtration
OR bath
place in an ice

If necessary

⑰ ⑪
⑧ ⑨

weg
hto calculate measure
melting
Dry crystals yeld

on s
point to check

watch
glass or
purity
felter paper

Point
Melting be products,
can reactants other
used
,

any catalyst
used and the solvent .

.
E it not have been dried sufficiently
.
g may

If the will be than expected.


the
product contains impurities , mething point lower

In addition the product will melt several °C rather than


range of
over a have a sharp
,

melting point .
GETTING TO THE NUB OF CHEMISTRY
Aromatic Chemistry 3.3.10
Bonding 3.3.10.1

Benzene has the formula C6H6. In 1865, August Kekulé proposed the following structure:

These subsequent pieces of evidence showed Kekulé to be wrong.

1. Unlike alkenes, benzene does not undergo electrophilic addition (e.g. it does not decolourise
bromine water).
2. X-ray analysis of solid benzene shows a regular hexagon with all the bonds the same length. This
length is intermediate between a typical single and double bond.
3. DH associated with hydrogenating the three unsaturated C=C bonds would be expected to be 3 x
that for hydrogenating the one double bond of cyclohexene (-120 kJ mol-1). However, the
experimental value of -208 kJ mol-1 is much smaller than the predicted value of -360 kJ mol-1.
GETTING TO THE NUB OF CHEMISTRY
The last point above makes it clear that the actual structure of benzene is considerably more stable than
Kekulé’s structure. The 152 kJ mol-1 is extra stability is called the delocalisation energy.

The explanation to all of the above is that the p orbitals at right angles to the plane of the carbon ring do
not just overlap in pairs to form double bonds, all six overlap together.

The electrons in the pi-bond do not ‘belong’ to any particular carbon atom but are delocalised and free
to moves throughout the entire pi system. It is this delocalisation that gives benzene its extra stability –
any system in which electron delocalisation occurs is stabilised.

The word aromatic is used to describe any system like that in benzene, which is stabilised by a ring of
delocalised pi-electrons. Non-aromatic compounds, such as alkanes and alkenes, are called aliphatic.

Naming isn’t as straight-forward as naming chain compounds. Often, more than one name is acceptable.
As a rule of thumb: if the group attached is one that never determines the suffix of a compound’s name
(e.g. halogen, alkyl, nitro etc.) or if it’s only the functional group attached with no chain, then we use
benzene.
GETTING TO THE NUB OF CHEMISTRY
If the group attached has a functional group that changes the suffix of an organic compound and/or has
a carbon chain, the prefix ‘phenyl’ is used. You get a phenyl group, C6H5, by removing a hydrogen from a
benzene ring, C6H6. Like a methyl or ethyl group, a phenyl group is always attached to something else.
Phenylamine is an exception to this.
GETTING TO THE NUB OF CHEMISTRY
Electrophilic Substitution 3.3.10.2

Electrophilic attack on benzene rings results in substitution, limited to monosubstitutions. Nitration is an


important step in synthesis, including the manufacture of explosives and formation of amines.

1. An electron pair from the ring forms a bond between one of the carbon atoms and the attacking
electrophile. This carbon atom now has four single bonds and so no longer contributes to the
delocalised system. The intermediate species is positively charged.
2. The pair of electrons in the C-H bond move back into the benzene ring reforming the delocalised
pi system.
GETTING TO THE NUB OF CHEMISTRY
Friedel–Crafts acylation reactions are also important steps in synthesis. Remember, acylation involved
replacing an H atom with an acyl group, RCO-
GETTING TO THE NUB OF CHEMISTRY
Amines 3.3.11
Amines can be primary, secondary and tertiary. If the nitrogen has four bonds, it’s a quaternary ammonium
ion.

When naming amines, they have the suffix -amine and are named after the groups surrounding the nitrogen.
If the groups are all the same, you add di- for secondary amines, tri- for tertiary amines and tetra for
quaternary ammonium ions. If the amine has more than one type of organic group attached, you list the
different groups in alphabetical order.

Preparation 3.3.11.1 and Nucleophilic Properties 3.3.11.3

Amines can be produced by heating excess ammonia in ethanol with a halogenoalkane (nucleophilic
substitution).
GETTING TO THE NUB OF CHEMISTRY
Unfortunately, the product is itself a nucleophile and can react with more chloroalkane to form successively
secondary and tertiary amines and ultimately a quaternary ammonium salt.

So, this method produces a mixture of products, which must be separated by distillation.

Halogenoalkane Ammonia à Products


Limiting reagent In excess à Mostly 1o amine
In excess Limiting reagent à Mostly quarternary

Having ammonia in excess if a required condition if you’re seeking to make a 1o amine otherwise further
substitution takes place.

The following flow diagram helps to picture the conditions required depending on the amine desired.

The reactant in excess is the one highlighted.

A more controlled method of producing primary aliphatic amines (the only product) is to reduce a nitrile
with LiAlH4 in dry ether followed by dilute sulfuric acid.

or, H2 and nickel catalyst with high temperature and pressure.


GETTING TO THE NUB OF CHEMISTRY
An aromatic amine can be produced by reducing a nitro compound, such as nitrobenzene. There are two
steps:

Quaternary ammonium salts can act as cationic surfactants.

Base Properties 3.3.11.2

1o, 2 o and 3o amines have a lone pair of electrons on their nitrogen atom. This enables them to accept a
proton (H+).
GETTING TO THE NUB OF CHEMISTRY
The salts of amines, i.e. protonated amines, are more soluble than the amine itself because ionic trumps H-
bonding. For example, Phenylamine is only very slightly soluble in water, but dissolves freely in dilute
hydrochloric acid.

So, like ammonia, 1o, 2 o and 3o amines are bases and form alkaline solutions.

When an acid is added to an amine, a salt is formed.

Is ethylamine a stronger or weaker base than ammonia? To answer this question, consider the following
equilibria:

The stronger the base, the further the position of equilibrium lies to the right. Alkyl groups are electron
pushing (c.f. carbocations). Because of this, the electron density on the nitrogen atom increases. We say the
lone pair of electrons on the N atom is more available, increasing the N atom’s ability to accept a proton
(H+). Alkyl groups have a positive inductive effect.
GETTING TO THE NUB OF CHEMISTRY
When a benzene ring is attached to the N…

Phenylamine is a weaker base than aliphatic amines and ammonia because the lone pair of electrons on the
nitrogen atom is partially delocalised around the benzene ring, so it’s less available to accept a proton (H+).

The simplest way to measure the comparative pHs of equimolar amines and ammonia is using a pH meter.
GETTING TO THE NUB OF CHEMISTRY
Polymers 3.3.12
Condensation Polymers 3.3.12.1

These are polymers made as a result of a series of condensation reactions (a reaction in a small molecule
is lost). The two types of condensation polymer we cover in this unit are polyesters and polyamides.

If you recall, esters could be formed by reacting a carboxylic acid and an alcohol.

And by reacting an acyl chloride with an alcohol.

Polyesters can form if the monomer(s) possess two of the relevant functional groups. For example:

This is how Terylene (PET) is made:


GETTING TO THE NUB OF CHEMISTRY
Note the direction of the ester linkages when we have two monomers.

Also, the repeating unit and polymer structures will show 2 linkages.

However, if it’s a hydroxycarboxylic acid (a single monomer)

Also, the repeating unit and polymer structures will show 1 linkage.

If you recall, amides could be forms as follows:


GETTING TO THE NUB OF CHEMISTRY
Polyamides can form if the monomer(s) possess two of the relevant functional groups. For example:

This is how Kevlar is made:

This is how Nylon 6,6 is made

Note the direction of the amide linkages.


GETTING TO THE NUB OF CHEMISTRY
Also, the repeating unit and polymer structures will show 2 linkages.

However, if it’s an amino acid.

Also, the repeating unit and polymer structures will show 1 linkage.

For the exam, you have to able to:

1. Draw the repeating unit from the monomer structure(s)

2. Draw the repeating unit from a section of the polymer chain

3. Draw the structure(s) of the monomer(s) from a section of the polymer.


GETTING TO THE NUB OF CHEMISTRY
Biodegradability and Disposal of Polymers 3.3.12.2

Just as esters and amides can be broken down by hydrolysis, so can polyesters and polyamides. This
allows them to be biodegradable. On the other hand, polyalkenes (addition polymers such as
polypropene) are not. This allows the possibility of recycling condensation polymers, such as PET.

Hydrolysis is possible because the δ+ C in the C=O group is susceptible to attack by a nucleophile. The C-C
bonds in addition polymers are non-polar and therefore not susceptible to attack.

All polymers, addition and condensation, have strong van der Waals forces between the chains.
Polyesters, in addition, will also possess permanent dipole-dipole interactions and polyamides hydrogen
bonds.

This often means that condensation polymers are stronger and more rigid than addition polymers.
Kevlar, for example, is used in stab-proof vests and puncture resistant tyres.
GETTING TO THE NUB OF CHEMISTRY
Amino acids, proteins and DNA 3.3.13
Amino Acids 3.3.13.1

Amino acids contain the amine group (-NH2) and the carboxylic acid group (-COOH). As such, they have both
acidic (proton donating) and basic (proton accepting) properties.

As solids and in solution, they exist as zwitterions (a molecule having a net charge of zero due to an equal
number of positive and negative charged functional groups). This is when the hydrogen atom from the
carboxylic acid group protonates the basic amine group e.g.

The strong electrostatic attraction between the oppositely charged parts of the ion account for the relatively
high melting-points of amino acids.

The charges on the zwitterions also means that amino acids move in an electric field. This property is the
basis of their separation by electrophoresis. They can also be separated using thin-layer chromatography
(see 3.3.16 Chromatography).

In solution, the zwitterions exist as a unique pH value – the isoelectric pH. Change the pH and you affect the
zwitterion.
GETTING TO THE NUB OF CHEMISTRY
There are 20 naturally occurring amino acids in humans, all with the general formula RCH(NH2) COOH. This
means they’re all 2-amino acids with the amino group on the carbon adjacent to the acid group.

With the exception of glycine, where R = H, all contain a carbon attached to four different groups ad
therefore display optical activity.

In biochemistry, the names of the 20 naturally occurring amino acids are often abbreviated to a 3-letter
‘code’, typically the first three letters of the name e.g. glycine (gly), alanine (ala) and serine (ser).

Amino acids undergo condensation reactions to form amides.

The C-N bond in -CONH- is called the peptide bond and the molecules above ate called dipeptides. If a third
amino acid were added e.g. gly-ala-ser, this would be a tripeptides. Chains containing many amino acids are
called polypeptides.

Hydrolysis of the peptide bond produces the constituent amino acids.

* strictly speaking it would be under acidic conditions and so the resulting amino acids would be protonated
GETTING TO THE NUB OF CHEMISTRY
Proteins 3.3.13.2

Proteins are condensation polymers of amino acids joined by peptide links. Their structure can be broken
into levels: primary secondary and tertiary.

The primary structure is the sequence of amino acids in the polypeptide chain that makes up a protein. E.g.

The secondary structure described how the protein is folded depending on the sequences of amino acids
that are next to each other. These can be α-helix and β–pleated sheets and hydrogen bonds hold the folded
structures in place.

The tertiary structure is the overall folding of the chain help by interactions between more distant amino
acids. These interactions include ionic attractions and disulfide bridges (covalent bonds that form between
sulfur atoms on the oxidation of two cysteine amino acids).
GETTING TO THE NUB OF CHEMISTRY
Enzymes are proteins that are biological catalysts. The molecules that they act on are called substrates and
each enzyme will only work with specific substrates. In the lock-and-key model, the part of the enzyme that
binds the substrate, known as the active site, accurately fits its particular substrate molecule.

The active sites of enzymes are stereospecific – they will only work on one enantiomer of the subtrate,

DNA (deoxyribonucleic acid) is a condensation polymer of nucleotides with a phosphate-sugar backbone.

Adenine pairs with thymine; Guanine pairs with cytosine. Remember to label lone pairs and partial charges.
GETTING TO THE NUB OF CHEMISTRY

DNA strands are complementary because the two strands match all A to T and all C to G.

Cisplatin is a drug for treating certain cancers. It has the following structure:

Cisplatin works by binding to two different Guanine (G) base groups on a strand of DNA. The cisplatin
undergoes a ligand substitution reaction with each Guanine – the N atom on the guanine replaces the Cl
atom on the cisplatin and a dative covalent bond is formed with Pt.

This prevents this section of DNA from being replicated. If the DNA is part of a cancerous cell, this will
prevent its multiplication. The disadvantage is that it also binds to the DNA in healthy cells and prevents
them from multiplying also. However, cancerous cells replicate faster so the drugs have a greater effect on
cancerous cells.
GETTING TO THE NUB OF CHEMISTRY
Organic Synthesis 3.3.14
For most synthetic routes, you could get from compound A to compound B in numerous ways. So, which
should chemists choose?

Synthetic routes which will be successful in an industrially setting should avoid solvents, use non-
hazardous starting materials and have steps with a high atom economy

Solvents may have environmental or toxicity concerns. In addition, solvents reduce the overall yield as it
is difficult to get 100% of the product solute out of the solution.

To have the greatest % yield and the most cost-effective and least wasteful synthesis, synthetic routes
with a high atom economy are always preferred if they prove to be cost-effective. Atom economy is the
calculation used to determine the efficiency of a reaction.

𝑡𝑜𝑡𝑎𝑙 𝑚𝑎𝑠𝑠 𝑜𝑓 𝑡ℎ𝑒 𝑑𝑒𝑠𝑖𝑟𝑒𝑑 𝑝𝑟𝑜𝑑𝑢𝑐𝑡


𝑃𝑒𝑟𝑐𝑒𝑛𝑡𝑎𝑔𝑒 𝑎𝑡𝑜𝑚 𝑒𝑐𝑜𝑛𝑜𝑚𝑦 = 𝑥 100%
𝑡𝑜𝑡𝑎𝑙 𝑚𝑎𝑠𝑠 𝑜𝑓 𝑡ℎ𝑒 𝑝𝑟𝑜𝑑𝑢𝑐𝑡𝑠

Let’s consider the atom economy of the fermentation of glucose to produce ethanol.

Large numbers of steps also reduce the % yield. This is because you lose some of your product in each
step along a synthetic route (maybe because it's hard to get all of your product out of a solvent or the
reaction doesn’t go to completion).

For example, in a three-step synthesis where each step had a 90% yield, the final overall yield would be
only 73%. If each of these steps instead had a yield of 50%, the overall yield would be 12.5%!

We can calculate the yield by working out how much we could get in theory, assuming all our starting
compound could be converted into product. This figure is then compared with the actual amount
obtained practically.

𝑎𝑐𝑡𝑢𝑎𝑙 𝑚𝑎𝑠𝑠 𝑜𝑏𝑡𝑎𝑖𝑛𝑒𝑑


𝑝𝑒𝑟𝑐𝑒𝑛𝑡𝑎𝑔𝑒 𝑦𝑖𝑒𝑙𝑑 = 𝑥 100%
𝑚𝑎𝑥𝑖𝑚𝑢𝑚 𝑡ℎ𝑒𝑜𝑟𝑒𝑡𝑖𝑐𝑎𝑙 𝑚𝑎𝑠𝑠
GETTING TO THE NUB OF CHEMISTRY
For example, a research chemist worked out a way to convert ethanol into ethanoyl chloride. She started
with 92.0g of ethanol and managed to prepare 68.0g of ethanoyl chloride. What was her yield?
GETTING TO THE NUB OF CHEMISTRY
When determining a synthetic route, watch out for the number of carbons increasing by one during the
synthesis as this would mean a nitrile was made in one of the steps.
GETTING TO THE NUB OF CHEMISTRY
Nuclear Magnetic Resonance Spectroscopy 3.3.15
Nuclear Magnetic Resonance (NMR) Spectroscopy is an analytical technique that can be used to determine
the structures of compounds by providing information about the position of 13C or 1H atoms in a molecule.

In 1H NMR (proton NMR), it’s vital that the sample being investigated is dissolved in a solvent that doesn’t
contain any hydrogen atoms, as these would produce confusing peaks in the spectrum. Suitable solvents are:

• Deuterated solvents e.g. CDCl3 (D = deuterium = 2H), which is a polar solvent.

• CCl4 – this is a non-polar solvent.

All measurements we make are against a standard. In 1H NMR, this is Tetramethylsilane (TMS) determines
where zero lies on the horizontal scale of the spectrum.

It is used because:

• It has 12 Hydrogens in the same environment, which results in a single, strong peak.

• It is non-toxic and inert

• It can be easily removed

• It gives a signal that is further right than most of the signals from organic compounds.

• At A-level, you’re unlikely to get a molecule that consists of more than 7 carbons and two functional
groups.

• You’ll likely be given the molecular formula and may be given some IR data. The latter will tell you if
it has a C=O group or not and whether there is an -OH. This could be from an alcohol or a carboxylic
acid – DON’T ASSUME it’s the latter if you also have a C=O).

• ALWAYS be specific about which you have: OH (alcohol) or OH (carboxylic acid).


GETTING TO THE NUB OF CHEMISTRY
When looking at an NMR spectrum, each signal corresponds to all hydrogens in a particular molecular
environment. The area under each signal is proportional to the number of hydrogens in that environment. In
ethanol, CH3CH2OH, for example, there are hydrogens in three environments: three in the CH3 group, two in
the CH2, and one in the OH. Here are some more examples:

Each signal may present itself as a singlet (s), doublet (d), triplet (t) or quartet (q). This is called the splitting
pattern.

The splitting pattern follows the rule n+1, where ‘n’ is the number of H’s attached to any carbons or nitrogen
adjacent to the environment the signal represents. NOTE: when the environment is -OH, it is always a singlet
as it’s never affected by adjacent C’s. It doesn’t affect the either.
GETTING TO THE NUB OF CHEMISTRY
The spectra are recorded on a scale known as the chemical shift (δ), which is how far the signal is away from
the signal for TMS is parts per million. The chemical shift depends on what other atoms/groups are near the
H. The closer the H is to electronegative atoms (e.g. O, Cl), the greater the shift. Also, the more
electronegative atoms there are close by, the greater the shift. The data sheet can be used to find
approximate values for chemical shifts.

If you’re given a spectrum, annotate the splitting patterns and number of Hs in each environment (signal).

1. Using the table over the page, try to identify what makes up each signal.

2. Keep an eye out for symmetry by considering how many signals there are vs. the number of carbons
in the molecular formula.

3. Having identified the parts that make up the overall molecule, draw the ends out and try joining
them with the atoms you still have left. Take care to think about how many more environments
there are.

4. Double-check your answer by checking how many environments it has, what the splitting patterns of
these would be, and their related chemical shifts.
GETTING TO THE NUB OF CHEMISTRY
No. of H’s Splitting Possible Fragment(s)

9 s

6 s

6 d

3 s

The above give you one end of the molecule. Use the chemical shift to determine
which is correct.

3,2 t,q

One end of the molecule. If you get 6,4 and t,q then it’s 2 x CH3CH2–

1 s
You’re like to have IR data, or other info, confirming the presence of this
functional group.

2,2 t,t

The adjacent atoms can’t have any H’s attached.


GETTING TO THE NUB OF CHEMISTRY

13
C NMR works in a very similar way to 1H NMR. The main differences are that there is no splitting in 13C
NMR, and there are no relative intensities. In this respect, the spectra are much simpler. Also, the chemical
shifts are very different so there’s a separate table on the data sheet. Here are some examples:
GETTING TO THE NUB OF CHEMISTRY
GETTING TO THE NUB OF CHEMISTRY
GETTING TO THE NUB OF CHEMISTRY
GETTING TO THE NUB OF CHEMISTRY
GETTING TO THE NUB OF CHEMISTRY
GETTING TO THE NUB OF CHEMISTRY
GETTING TO THE NUB OF CHEMISTRY
Chromatography 3.3.16
Chromatography is a method for separating mixtures that allows for the possibility of identifying the
components of that mixture.

All chromatography comprises:

• A mobile phase (this is always a liquid or gas)

• A stationary phase (this is a solid or a liquid on a solid support)

There are three types of chromatography that you need to be aware of:

1. Thin-layer chromatography (required practical 12)

2. Column chromatography (primarily a separation technique)


GETTING TO THE NUB OF CHEMISTRY
3. Gas chromatography

The Rf Value (Rf = retention factor) compared the distance moved by each component of the solute with the
distance moved by the solvent during the experiment (Thin-layer chromatography). You measure to the
centre of the spot. Rf can’t be greater than one.

The time taken for a particular compound to travel through the column to the detector is known as
its retention time. This time is measured from when the sample is injected to the point at which the display
shows a maximum peak height for that compound. Different compounds have different retention times.
Below, F takes the most time to pass through the column and, therefore, has the highest retention time.
GETTING TO THE NUB OF CHEMISTRY
Chromatography works as a separation technique because each component of a mixture will interact
differently with the two phases. Substances are separated according to relative affinity (attraction) to the
stationary and mobile phases. How fast/far each compound moves depends on the relative attraction
(affinity) of each compound to the mobile and the stationary phases.

• If they have a stronger affinity for the mobile phase – they’re more soluble and move quickly.
• If they have a stronger affinity for the stationary phase – they move slowly

Components that are more soluble (have a greater affinity with the mobile phase) will travel further (faster)
than components that are more strongly adsorbed (stuck/attached to the surface) to the stationary phase.

Usually, one phase will be polar and the other non-polar.

The stationary phase typically comprises SiO2 (silica) or Al2O3 (alumina), which are regarded as polar powders
due to the nature of their surface structure.

The mobile phase can be a non-polar (e.g. an alkane) or polar solvent (e.g. water, alcohols).

If, for example, hexane were used as the mobile phase and silica as the stationary phase, then non-polar
molecules would move quickly and polar molecules would move slowly.

Some comparisons.

Component Stationary Phase Mobile Phase Speed

(strong affinity) (weak affinity)


Polar Polar Non-polar Slow
(adsorbs well) (not v. soluble)

(weak affinity) (strong affinity)


Non-polar Polar Non-polar Fast
(adsorbs poorly) (v. soluble)

(weak affinity) (strong affinity)


Polar Non-polar Polar Fast
(adsorbs poorly) (v. soluble)

(strong affinity) (weak affinity)


Non-polar Non-polar Polar Slow
(adsorbs well) (not v. soluble)
GETTING TO THE NUB OF CHEMISTRY
Chromatography and Amino Acids: As different amino acids have different ‘R’ groups, they will have
different solubilities in the same solvent. This means you can easily separate and identify the different amino
acids in a mixture using thin-layer chromatography.

Amino acids aren’t coloured – you need to make them visible. This can be achieved by:

• Adding a fluorescent dye to the silica or alumina layer that glows when UV light shines on it.

• Spraying ninhydrin solution on the plate – this causes the amino acids to turn purple.

Sometimes, two or more components in a mixture can have similar Rf values in a particular solvent, hence
overlapping occurs. So, two-way chromatography is used.

M represents the position of the original spot.

1. How many spots would there have been in the chromatogram after you had used solvent 1 and
before you used solvent 2? 4: ‘B and C’ and ‘D and E’ wouldn’t have separated out.

2. Which spot is due to a substance almost insoluble in solvent 2? D – it has hardly moved at all when
solvent 2 was used.

3. Which spot is due to a substance almost insoluble in solvent 1, but very soluble in solvent 2? F – it
only travelled a little way up the paper when solvent 1 was used but was carried a long way by
solvent 2.

You might also like