Sie sind auf Seite 1von 56

THE COLLISION THEORY OF

REACTION RATES
This page describes the collision theory of reaction rates. It
concentrates on the key things which decide whether a particular
collision will result in a reaction - in particular, the energy of the
collision, and whether or not the molecules hit each other the right
way around (the orientation of the collision).
The individual factors which affect the rate of a reaction
(temperature, concentration, and so on) are discussed on
separate pages. You can get at these via the rates of reaction
menu - there is a link at the bottom of the page.
We are going to look in detail at reactions which involve a collision
between two species.
Species: This is a useful term which covers any sort of
particle you like - molecule, ion, or free radical.

Reactions where a single species falls apart in some way are


slightly simpler because you won't be involved in worrying about
the orientation of collisions. Reactions involving collisions between
more than two species are going to be extremely uncommon (see
below).

Reactions involving collisions between two species


It is pretty obvious that if you have a situation involving two species
they can only react together if they come into contact with each
other. They first have to collide, and then they may react.
Why "may react"? It isn't enough for the two species to collide they have to collide the right way around, and they have to collide
with enough energy for bonds to break.
(The chances of all this happening if your reaction needed a
collision involving more than 2 particles are remote. All three (or
more) particles would have to arrive at exactly the same point in
space at the same time, with everything lined up exactly right, and
having enough energy to react. That's not likely to happen very
often!)

The orientation of collision


Consider a simple reaction involving a collision between two
molecules - ethene, CH2=CH2, and hydrogen chloride, HCl, for
example. These react to give chloroethane.

As a result of the collision between the two molecules, the double


bond between the two carbons is converted into a single bond. A
hydrogen atom gets attached to one of the carbons and a chlorine
atom to the other.

converted by W eb2PDFConvert.com

Note: The mechanism for this reaction is dealt with on a


separate page. This might help you to understand why the
orientation of the two molecules is so important.
If you want to read a bit more about this, follow this link and
use the BACK button on your browser to return to this page.

The reaction can only happen if the hydrogen end of the H-Cl bond
approaches the carbon-carbon double bond. Any other collision
between the two molecules doesn't work. The two simply bounce
off each other.

Of the collisions shown in the diagram, only collision 1 may


possibly lead on to a reaction.
If you haven't read the page about the mechanism of the reaction,
you may wonder why collision 2 won't work as well. The double
bond has a high concentration of negative charge around it due to
the electrons in the bonds. The approaching chlorine atom is also
slightly negative because it is more electronegative than hydrogen.
The repulsion simply causes the molecules to bounce off each
other.
Note: If you aren't sure about electronegativity , you might
like to follow this link.
Use the BACK button on your browser to return to this page.

converted by W eb2PDFConvert.com

In any collision involving unsymmetrical species, you would expect


that the way they hit each other will be important in deciding
whether or not a reaction happens.

The energy of the collision


Activation Energy
Even if the species are orientated properly, you still won't get a
reaction unless the particles collide with a certain minimum energy
called the activation energy of the reaction.
Activation energy is the minimum energy required before a
reaction can occur. You can show this on an energy profile for the
reaction. For a simple over-all exothermic reaction, the energy
profile looks like this:

Note: The only difference if the reaction was endothermic


would be the relative positions of the reactants and products
lines. For an endothermic change, the products would have a
higher energy than the reactants, and so the green arrow
would be pointing upwards. It makes no difference to the
discussion about the activation energy.

converted by W eb2PDFConvert.com

If the particles collide with less energy than the activation energy,
nothing important happens. They bounce apart. You can think of
the activation energy as a barrier to the reaction. Only those
collisions which have energies equal to or greater than the
activation energy result in a reaction.
Any chemical reaction results in the breaking of some bonds
(needing energy) and the making of new ones (releasing energy).
Obviously some bonds have to be broken before new ones can be
made. Activation energy is involved in breaking some of the
original bonds.
Where collisions are relatively gentle, there isn't enough energy
available to start the bond-breaking process, and so the particles
don't react.

The Maxwell-Boltzmann Distribution


Because of the key role of activation energy in deciding whether a
collision will result in a reaction, it would obviously be useful to
know what sort of proportion of the particles present have high
enough energies to react when they collide.
In any system, the particles present will have a very wide range of
energies. For gases, this can be shown on a graph called the
Maxwell-Boltzmann Distribution which is a plot of the number of
particles having each particular energy.
Note: The graph only applies to gases, but the conclusions
that we can draw from it can also be applied to reactions
involving liquids.

The area under the curve is a measure of the total number of


particles present.
Note: The reason for this lies in some maths beyond the
scope of an A'level chemistry course. It is important that you
remember that the area under the curve gives a count of the
number of particles even if you don't understand why!

converted by W eb2PDFConvert.com

The Maxwell-Boltzmann Distribution and activation energy


Remember that for a reaction to happen, particles must collide
with energies equal to or greater than the activation energy for the
reaction. We can mark the activation energy on the MaxwellBoltzmann distribution:

Notice that the large majority of the particles don't have enough
energy to react when they collide. To enable them to react we
either have to change the shape of the curve, or move the
activation energy further to the left. This is described on other
pages.
Note: You can change the shape of the curve by changing
the temperature of the reaction. You can change the position
of the activation energy by adding a catalyst to the reaction.
You could either go straight to these pages if you are
interested, or access them later via the rates of reaction
menu (link at the bottom of the page).

Questions to test your understanding


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

Where would you like to go now?


To the rates of reaction menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002 (modified October 2013)

converted by W eb2PDFConvert.com

THE EFFECT OF SURFACE AREA ON


REACTION RATES
This page describes and explains the effect of changing the
surface area of a solid on the rate of a reaction it is involved in.
This applies to reactions involving a solid and a gas, or a solid and
a liquid. It includes cases where the solid is acting as a catalyst.

The facts
What happens?
The more finely divided the solid is, the faster the reaction
happens. A powdered solid will normally produce a faster reaction
than if the same mass is present as a single lump. The powdered
solid has a greater surface area than the single lump.
Note: Why normally? What exceptions can there be?
Imagine a case of a very fine powder reacting with a gas. If
the powder was in one big heap, the gas may not be able to
penetrate it. That means that its effective surface area is
much the same as (or even less than) it would be if it were
present in a single lump.
A small heap of fine magnesium powder tends to burn rather
more slowly than a strip of magnesium ribbon, for example.

Some examples
Calcium carbonate and hydrochloric acid
In the lab, powdered calcium carbonate reacts much faster with
dilute hydrochloric acid than if the same mass was present as
lumps of marble or limestone.

The catalytic decomposition of hydrogen peroxide


This is another familiar lab reaction. Solid manganese(IV) oxide is
often used as the catalyst. Oxygen is given off much faster if the
catalyst is present as a powder than as the same mass of
granules.

Catalytic converters
Catalytic converters use metals like platinum, palladium and
rhodium to convert poisonous compounds in vehicle exhausts into
less harmful things. For example, a reaction which removes both
carbon monoxide and an oxide of nitrogen is:

Because the exhaust gases are only in contact with the catalyst for
a very short time, the reactions have to be very fast. The extremely
expensive metals used as the catalyst are coated as a very thin
layer onto a ceramic honeycomb structure to maximise the surface
area.

The explanation

converted by W eb2PDFConvert.com

You are only going to get a reaction if the particles in the gas or
liquid collide with the particles in the solid. Increasing the surface
area of the solid increases the chances of collision taking place.
Imagine a reaction between magnesium metal and a dilute acid
like hydrochloric acid. The reaction involves collision between
magnesium atoms and hydrogen ions.

Increasing the number of collisions per second increases the rate


of reaction.

Questions to test your understanding


You will find questions about all the factors affecting rates of reaction on the
page about catalysts at the end of this sequence of pages.

Where would you like to go now?


To the rates of reaction menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002 (modified October 2013)

converted by W eb2PDFConvert.com

THE EFFECT OF CONCENTRATION ON


REACTION RATES
This page describes and explains the way that changing the
concentration of a solution affects the rate of a reaction. Be aware
that this is an introductory page only. If you are interested in orders
of reaction, you will find separate pages dealing with these. You
can access these via the rates of reaction menu (link at the bottom
of the page).

The facts
What happens?
For many reactions involving liquids or gases, increasing the
concentration of the reactants increases the rate of reaction. In a
few cases, increasing the concentration of one of the reactants
may have little noticeable effect of the rate. These cases are
discussed and explained further down this page.
Don't assume that if you double the concentration of one of the
reactants that you will double the rate of the reaction. It may
happen like that, but the relationship may well be more
complicated.
Note: The mathematical relationship between concentration
and rate of reaction is dealt with on the page about orders of
reaction. If you are interested, you can use this link or read
about it later via the rate of reaction menu (link at the bottom
of the page).

Some examples
The examples on this page all involve solutions. Changing the
concentration of a gas is achieved by changing its pressure. This
is covered on a separate page.
Note: If you want to explore the effect of changing pressure
on the rate of a reaction, you could use this link.
Alternatively, use the link to the rates of reaction menu at the
bottom of this page.

Zinc and hydrochloric acid


In the lab, zinc granules react fairly slowly with dilute hydrochloric
acid, but much faster if the acid is concentrated.

The catalytic decomposition of hydrogen peroxide


Solid manganese(IV) oxide is often used as a catalyst in this
reaction. Oxygen is given off much faster if the hydrogen peroxide
is concentrated than if it is dilute.

The reaction between sodium thiosulphate solution and


hydrochloric acid
This is a reaction which is often used to explore the relationship
between concentration and rate of reaction in introductory courses
converted by W eb2PDFConvert.com

(like GCSE). When a dilute acid is added to sodium thiosulphate


solution, a pale yellow precipitate of sulphur is formed.

As the sodium thiosulphate solution is diluted more and more, the


precipitate takes longer and longer to form.

The explanation
Cases where changing the concentration affects the rate of
the reaction
This is the common case, and is easily explained.
Collisions involving two particles
The same argument applies whether the reaction involves collision
between two different particles or two of the same particle.
In order for any reaction to happen, those particles must first
collide. This is true whether both particles are in solution, or
whether one is in solution and the other a solid. If the concentration
is higher, the chances of collision are greater.

Reactions involving only one particle


If a reaction only involves a single particle splitting up in some way,
then the number of collisions is irrelevant. What matters now is how
many of the particles have enough energy to react at any one time.
Note: If you aren't sure about this, then read the page about
collision theory and activation energy before you go on. Use
the BACK button on your browser to return to this page.

converted by W eb2PDFConvert.com

Suppose that at any one time 1 in a million particles have enough


energy to equal or exceed the activation energy. If you had 100
million particles, 100 of them would react. If you had 200 million
particles in the same volume, 200 of them would now react. The
rate of reaction has doubled by doubling the concentration.

Cases where changing the concentration doesn't affect the


rate of the reaction
At first glance this seems very surprising!
Where a catalyst is already working as fast as it can
Suppose you are using a small amount of a solid catalyst in a
reaction, and a high enough concentration of reactant in solution
so that the catalyst surface was totally cluttered up with reacting
particles.
Increasing the concentration of the solution even more can't have
any effect because the catalyst is already working at its maximum
capacity.
In certain multi-step reactions
This is the more important effect from an A' level point of view.
Suppose you have a reaction which happens in a series of small
steps. These steps are likely to have widely different rates - some
fast, some slow.
For example, suppose two reactants A and B react together in
these two stages:

The overall rate of the reaction is going to be governed by how fast


A splits up to make X and Y. This is described as the rate
determining step of the reaction.
If you increase the concentration of A, you will increase the
chances of this step happening for reasons we've looked at above.
If you increase the concentration of B, that will undoubtedly speed
up the second step, but that makes hardly any difference to the
overall rate. You can picture the second step as happening so fast
already that as soon as any X is formed, it is immediately pounced
on by B. That second reaction is already "waiting around" for the
first one to happen.
Note: The overall rate of reaction isn't entirely independent of
the concentration of B. If you lowered its concentration
enough, you will eventually reduce the rate of the second
reaction to the point where it is similar to the rate of the first.
Both concentrations will matter if the concentration of B is
low enough.
However, for ordinary concentrations, you can say that (to a
good approximation) the overall rate of reaction is unaffected
by the concentration of B.

The best specific examples of reactions of this type comes from


organic chemistry. These involve the reaction between a tertiary
halogenoalkane (alkyl halide) and a number of possible
substances - including hydroxide ions. These are examples of
nucleophilic substitution using a mechanism known as SN1.

converted by W eb2PDFConvert.com

Note: If you are interested in exploring nucleophilic


substitution reactions further, you could follow this link.
Otherwise, you can find more about how the relationship
between concentration and rate of reaction is affected by
reaction mechanisms by exploring the topics at the bottom of
the rates of reaction menu (link below).

Questions to test your understanding


You will find questions about all the factors affecting rates of reaction on the
page about catalysts at the end of this sequence of pages.

Where would you like to go now?


To the rates of reaction menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002 (modified October 2013)

converted by W eb2PDFConvert.com

THE EFFECT OF PRESSURE ON


REACTION RATES
This page describes and explains the way that changing the
pressure of a gas changes the rate of a reaction.

The facts
What happens?
Increasing the pressure on a reaction involving reacting gases
increases the rate of reaction. Changing the pressure on a
reaction which involves only solids or liquids has no effect on the
rate.

An example
In the manufacture of ammonia by the Haber Process, the rate of
reaction between the hydrogen and the nitrogen is increased by
the use of very high pressures.

In fact, the main reason for using high pressures is to improve the
percentage of ammonia in the equilibrium mixture, but there is a
useful effect on rate of reaction as well.
Note: If you want to explore equilibria you will find the topic
covered in a separate section of the site.

The explanation
The relationship between pressure and concentration
Increasing the pressure of a gas is exactly the same as increasing
its concentration. If you have a given mass of gas, the way you
increase its pressure is to squeeze it into a smaller volume. If you
have the same mass in a smaller volume, then its concentration is
higher.
You can also show this relationship mathematically if you have
come across the ideal gas equation:

Rearranging this gives:

Because "RT" is constant as long as the temperature is constant,


this shows that the pressure is directly proportional to the
concentration. If you double one, you will also double the other.
converted by W eb2PDFConvert.com

Note: If you should be able to do calculations involving the


ideal gas equation, but aren't very happy about them, you
might be interested in my chemistry calculations book.

The effect of increasing the pressure on the rate of reaction


Collisions involving two particles
The same argument applies whether the reaction involves collision
between two different particles or two of the same particle.
In order for any reaction to happen, those particles must first
collide. This is true whether both particles are in the gas state, or
whether one is a gas and the other a solid. If the pressure is
higher, the chances of collision are greater.

Reactions involving only one particle


If a reaction only involves a single particle splitting up in some way,
then the number of collisions is irrelevant. What matters now is how
many of the particles have enough energy to react at any one time.
Note: If you aren't sure about this, then read the page about
collision theory and activation energy before you go on. Use
the BACK button on your browser to return to this page.

converted by W eb2PDFConvert.com

Suppose that at any one time 1 in a million particles have enough


energy to equal or exceed the activation energy. If you had 100
million particles, 100 of them would react. If you had 200 million
particles in the same volume, 200 of them would now react. The
rate of reaction has doubled by doubling the pressure.

Questions to test your understanding


You will find questions about all the factors affecting rates of reaction on the
page about catalysts at the end of this sequence of pages.

Where would you like to go now?


To the rates of reaction menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002 (modified October 2013)

converted by W eb2PDFConvert.com

THE EFFECT OF TEMPERATURE ON


REACTION RATES
This page describes and explains the way that changing the
temperature affects the rate of a reaction. It assumes that you are
already familiar with basic ideas about the collision theory, and
with the Maxwell-Boltzmann distribution of molecular energies in a
gas.
Note: If you haven't already read the page about collision
theory, you should do so before you go on.
Use the BACK button on your browser to return to this page,
or come back via the rates of reaction menu.

The facts
What happens?
As you increase the temperature the rate of reaction increases. As
a rough approximation, for many reactions happening at around
room temperature, the rate of reaction doubles for every 10C rise
in temperature.
You have to be careful not to take this too literally. It doesn't apply
to all reactions. Even where it is approximately true, it may be that
the rate doubles every 9C or 11C or whatever. The number of
degrees needed to double the rate will also change gradually as
the temperature increases.
Note: You will find the effect of temperature on rate explored
in a slightly more mathematical way on a separate page.

converted by W eb2PDFConvert.com

Examples
Some reactions are virtually instantaneous - for example, a
precipitation reaction involving the coming together of ions in
solution to make an insoluble solid, or the reaction between
hydrogen ions from an acid and hydroxide ions from an alkali in
solution. So heating one of these won't make any noticeable
difference to the rate of the reaction.
Almost any other reaction you care to name will happen faster if
you heat it - either in the lab, or in industry.

The explanation
Increasing the collision frequency
Particles can only react when they collide. If you heat a substance,
the particles move faster and so collide more frequently. That will
speed up the rate of reaction.
That seems a fairly straightforward explanation until you look at the
numbers!
It turns out that the frequency of two-particle collisions in gases is
proportional to the square root of the kelvin temperature. If you
increase the temperature from 293 K to 303 K (20C to 30C), you
will increase the collision frequency by a factor of:

That's an increase of 1.7% for a 10 rise. The rate of reaction will


probably have doubled for that increase in temperature - in other
words, an increase of about 100%. The effect of increasing
collision frequency on the rate of the reaction is very minor. The
important effect is quite different . . .

The key importance of activation energy


Collisions only result in a reaction if the particles collide with
enough energy to get the reaction started. This minimum energy
required is called the activation energy for the reaction.
Note: What follows assumes you have a reasonable idea
about activation energy and its relationship with the MaxwellBoltzmann distribution. This is covered on the introductory
page about collision theory.
If you aren't confident about this, follow this link, and use the
BACK button on your browser to return to this page.

You can mark the position of activation energy on a MaxwellBoltzmann distribution to get a diagram like this:

converted by W eb2PDFConvert.com

Only those particles represented by the area to the right of the


activation energy will react when they collide. The great majority
don't have enough energy, and will simply bounce apart.
To speed up the reaction, you need to increase the number of the
very energetic particles - those with energies equal to or greater
than the activation energy. Increasing the temperature has exactly
that effect - it changes the shape of the graph.
In the next diagram, the graph labelled T is at the original
temperature. The graph labelled T+t is at a higher temperature.

If you now mark the position of the activation energy, you can see
that although the curve hasn't moved very much overall, there has
been such a large increase in the number of the very energetic
particles that many more now collide with enough energy to react.

Remember that the area under a curve gives a count of the number
of particles. On the last diagram, the area under the higher
temperature curve to the right of the activation energy looks to
have at least doubled - therefore at least doubling the rate of the
reaction.

converted by W eb2PDFConvert.com

Summary
Increasing the temperature increases reaction rates because of
the disproportionately large increase in the number of high energy
collisions. It is only these collisions (possessing at least the
activation energy for the reaction) which result in a reaction.

Questions to test your understanding


You will find questions about all the factors affecting rates of reaction on the
page about catalysts at the end of this sequence of pages.

Where would you like to go now?


To the rates of reaction menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002 (modified October 2013)

converted by W eb2PDFConvert.com

THE EFFECT OF CATALYSTS ON


REACTION RATES
This page describes and explains the way that adding a catalyst
affects the rate of a reaction. It assumes that you are already
familiar with basic ideas about the collision theory of reaction
rates, and with the Maxwell-Boltzmann distribution of molecular
energies in a gas.
Note: If you haven't already read the page about collision
theory, you should do so before you go on.
Use the BACK button on your browser to return to this page,
or come back via the rates of reaction menu.

Note that this is only a preliminary look at catalysis as far as it


affects rates of reaction. If you are looking for more detail, there is
a separate section dealing with catalysts which you can access via
a link at the bottom of the page.

The facts
What are catalysts?
A catalyst is a substance which speeds up a reaction, but is
chemically unchanged at the end of the reaction. When the
reaction has finished, you would have exactly the same mass of
catalyst as you had at the beginning.

Some examples
Some common examples which you may need for other parts of
your syllabus include:

reaction

catalyst

Decomposition of hydrogen peroxide

manganese(IV)
oxide, MnO2

Nitration of benzene

concentrated
sulphuric acid

Manufacture of ammonia by the Haber


Process

iron

Conversion of SO2 into SO3 during the


Contact Process to make sulphuric acid

vanadium(V)
oxide, V2O5

Hydrogenation of a C=C double bond

nickel

Note: You can find details of these and other catalytic


reactions by exploring the menu for the main section on
catalysis. You will find a link at the bottom of this page.

converted by W eb2PDFConvert.com

The explanation
The key importance of activation energy
Collisions only result in a reaction if the particles collide with a
certain minimum energy called the activation energy for the
reaction.
Note: What follows assumes you have a reasonable idea
about activation energy and its relationship with the MaxwellBoltzmann distribution. This is covered on the introductory
page about collision theory.
If you aren't confident about this, follow this link, and use the
BACK button on your browser to return to this page.

You can mark the position of activation energy on a MaxwellBoltzmann distribution to get a diagram like this:

Only those particles represented by the area to the right of the


activation energy will react when they collide. The great majority
don't have enough energy, and will simply bounce apart.

Catalysts and activation energy


To increase the rate of a reaction you need to increase the number
of successful collisions. One possible way of doing this is to
provide an alternative way for the reaction to happen which has a
lower activation energy.
In other words, to move the activation energy on the graph like this:

converted by W eb2PDFConvert.com

Adding a catalyst has exactly this effect on activation energy. A


catalyst provides an alternative route for the reaction. That
alternative route has a lower activation energy. Showing this on an
energy profile:

A word of caution!
Be very careful if you are asked about this in an exam. The correct
form of words is
"A catalyst provides an alternative route for the reaction with
a lower activation energy."
It does not "lower the activation energy of the reaction". There is a
subtle difference between the two statements that is easily
illustrated with a simple analogy.
Suppose you have a mountain between two valleys so that the only
way for people to get from one valley to the other is over the
mountain. Only the most active people will manage to get from one
valley to the other.
Now suppose a tunnel is cut through the mountain. Many more
people will now manage to get from one valley to the other by this
easier route. You could say that the tunnel route has a lower
activation energy than going over the mountain.
But you haven't lowered the mountain! The tunnel has provided an
alternative route but hasn't lowered the original one. The original
mountain is still there, and some people will still choose to climb it.
In the chemistry case, if particles collide with enough energy they
can still react in exactly the same way as if the catalyst wasn't
there. It is simply that the majority of particles will react via the
easier catalysed route.

converted by W eb2PDFConvert.com

Questions to test your understanding


If this is the first set of questions you have done, please read the
introductory page before you start. You will need to use the BACK BUTTON
on your browser to come back here afterwards.
These questions cover all the various factors which affect rates of
reaction, not just catalysts.
questions on factors affecting rates of reaction
answers

Where would you like to go now?


To the more detailed catalysis menu . . .
To the rates of reaction menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002 (modified October 2013)

converted by W eb2PDFConvert.com

ORDERS OF REACTION AND RATE


EQUATIONS
Changing the concentration of substances taking part in a reaction
usually changes the rate of the reaction. A rate equation shows this
effect mathematically. Orders of reaction are a part of the rate
equation. This page introduces and explains the various terms you
will need to know about.
Note: If you aren't sure about why changing concentration
affects rates of reaction you might like to follow this link and
come back to this page afterwards - either via the rates of
reaction menu or by using the BACK button on your browser.

Rate equations
Measuring a rate of reaction
There are several simple ways of measuring a reaction rate. For
example, if a gas was being given off during a reaction, you could
take some measurements and work out the volume being given off
per second at any particular time during the reaction.
A rate of 2 cm3 s-1 is obviously twice as fast as one of 1 cm3 s-1.
Note: Read cm3 s-1 as "cubic centimetres per second".

However, for this more formal and mathematical look at rates of


reaction, the rate is usually measured by looking at how fast the
concentration of one of the reactants is falling at any one time.
For example, suppose you had a reaction between two
substances A and B. Assume that at least one of them is in a form
where it is sensible to measure its concentration - for example, in
solution or as a gas.

For this reaction you could measure the rate of the reaction by
finding out how fast the concentration of, say, A was falling per
second.
You might, for example, find that at the beginning of the reaction,
its concentration was falling at a rate of 0.0040 mol dm-3 s-1.
Note: Read mol dm-3 s-1 as "moles per cubic decimetre (or
litre) per second".

This means that every second the concentration of A was falling by


0.0040 moles per cubic decimetre. This rate will decrease during
the reaction as A gets used up.
Summary
For the purposes of rate equations and orders of reaction, the rate
of a reaction is measured in terms of how fast the concentration of
one of the reactants is falling. Its units are mol dm-3 s-1.

converted by W eb2PDFConvert.com

Orders of reaction
I'm not going to define what order of reaction means straight away
- I'm going to sneak up on it!
Orders of reaction are always found by doing experiments. You
can't deduce anything about the order of a reaction just by looking
at the equation for the reaction.
So let's suppose that you have done some experiments to find out
what happens to the rate of a reaction as the concentration of one
of the reactants, A, changes. Some of the simple things that you
might find are:
One possibility: The rate of reaction is proportional to the
concentration of A
That means that if you double the concentration of A, the rate
doubles as well. If you increase the concentration of A by a factor
of 4, the rate goes up 4 times as well.
You can express this using symbols as:

Writing a formula in square brackets is a standard way of showing


a concentration measured in moles per cubic decimetre (litre).
You can also write this by getting rid of the proportionality sign and
introducing a constant, k.

Another possibility: The rate of reaction is proportional to the


square of the concentration of A
This means that if you doubled the concentration of A, the rate
would go up 4 times (22). If you tripled the concentration of A, the
rate would increase 9 times (32). In symbol terms:

Generalising this
By doing experiments involving a reaction between A and B, you
would find that the rate of the reaction was related to the
concentrations of A and B in this way:

converted by W eb2PDFConvert.com

This is called the rate equation for the reaction.


The concentrations of A and B have to be raised to some power to
show how they affect the rate of the reaction. These powers are
called the orders of reaction with respect to A and B.
For UK A' level purposes, the orders of reaction you are likely to
meet will be 0, 1 or 2. But other values are possible including
fractional ones like 1.53, for example.
If the order of reaction with respect to A is 0 (zero), this means that
the concentration of A doesn't affect the rate of reaction.
Mathematically, any number raised to the power of zero (x0) is
equal to 1. That means that that particular term disappears from
the rate equation.
The overall order of the reaction is found by adding up the
individual orders. For example, if the reaction is first order with
respect to both A and B (a = 1 and b = 1), the overall order is 2.
We call this an overall second order reaction.
Some examples
Each of these examples involves a reaction between A and B, and
each rate equation comes from doing some experiments to find
out how the concentrations of A and B affect the rate of reaction.

Example 1:

In this case, the order of reaction with respect to both A and B is 1.


The overall order of reaction is 2 - found by adding up the
individual orders.
Note: Where the order is 1 with respect to one of the
reactants, the "1" isn't written into the equation. [A] means
[A]1.

converted by W eb2PDFConvert.com

Example 2:

This reaction is zero order with respect to A because the


concentration of A doesn't affect the rate of the reaction. The order
with respect to B is 2 - it's a second order reaction with respect to
B. The reaction is also second order overall (because 0 + 2 = 2).

Example 3:

This reaction is first order with respect to A and zero order with
respect to B, because the concentration of B doesn't affect the
rate of the reaction. The reaction is first order overall (because 1 +
0 = 1).
What if you have some other number of reactants?
It doesn't matter how many reactants there are. The concentration
of each reactant will occur in the rate equation, raised to some
power. Those powers are the individual orders of reaction. The
overall order of the reaction is found by adding them all up.

The rate constant


Surprisingly, the rate constant isn't actually a true constant! It
varies, for example, if you change the temperature of the reaction,
add a catalyst, or change the catalyst.
The rate constant is constant for a given reaction only if all you are
changing is the concentration of the reactants. You will find more
about the effect of temperature and catalysts on the rate constant
on another page.
Note: If you want to follow up this further look at rate
constants you might like to follow this link. Alternatively, you
could visit it later via the rates of reaction menu.

Calculations involving orders of reaction


You will almost certainly have to be able to calculate orders of
reaction and rate constants from given data or from your own
experiments.
There are all sorts of ways of doing these sums, and it is important
that you practice the methods that your syllabus wants. Check your
syllabus and past exam papers to see what sort of examples you
need to be able to work out.
Note: For UK A'level students, if you haven't got copies of
your syllabus and past papers follow this link to find out how
to get hold of them.

Many text books make these sums look really difficult. In fact for A'
level purposes, the calculations are usually fairly trivial. You will find
them explained in detail in my chemistry calculations book.

converted by W eb2PDFConvert.com

Note: There are several reasons why there are very few
calculations on this site. It is much easier to learn to do
sums from a carefully organised book than from a website; I
would be in breach of my contract with my publishers if I
included material similar to what is in the book; and I need to
sell a few books to generate some income!
If you are interested in my chemistry calculations book you
might like to follow this link.

Questions to test your understanding


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

Where would you like to go now?


To a simple look at how orders of reaction are related to
reaction mechanisms . .
To the rates of reaction menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002 (modified October 2013)

converted by W eb2PDFConvert.com

ORDERS OF REACTION AND


MECHANISMS
This page looks at the relationship between orders of reaction and
mechanisms in some simple cases. It explores what a mechanism
is, and the idea of a rate determining step. It also explains the
difference between the sometimes confusing terms "order of
reaction" and "molecularity of reaction".
Note: If you aren't sure about orders of reaction you ought to
read the introductory page before you go on. You will find a
link back to here at the bottom of the introductory page.

Reaction mechanisms
What is a reaction mechanism?
In any chemical change, some bonds are broken and new ones
are made. Quite often, these changes are too complicated to
happen in one simple stage. Instead, the reaction may involve a
series of small changes one after the other.
A reaction mechanism describes the one or more steps involved
in the reaction in a way which makes it clear exactly how the
various bonds are broken and made. The following example
comes from organic chemistry. It doesn't matter in the least if it is
unfamiliar to you!
This is a reaction between 2-bromo-2-methylpropane and the
hydroxide ions from sodium hydroxide solution:

The overall reaction replaces the bromine atom in the organic


compound by an OH group.
The first thing that happens is that the carbon-bromine bond in a
small proportion of the organic compound breaks to give ions:

Carbon-bromine bonds are reasonably strong, so this is a slow


change. If the ions hit each other again, the covalent bond will
reform. The curly arrow in the equation shows the movement of a
pair of electrons.
If there is a high concentration of hydroxide ions present, the
positive ion stands a high chance of hitting one of those. This step
of the overall reaction will be very fast. A new covalent bond is
made between the carbon and the oxygen, using one of the lone
pairs on the oxygen atom.

Because carbon-oxygen bonds are strong, once the OH group has


attached to the carbon atom, it tends to stay attached.
The mechanism shows that the reaction takes place in two steps
and describes exactly how those steps happen in terms of bonds
being broken or made. It also shows that the steps have different
rates of reaction - one slow and one fast.
converted by W eb2PDFConvert.com

Note: If you are interested in exploring more organic reaction


mechanisms you will find probably more than you will want to
know about by following this link!

The rate determining step


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. In the example above, the hydroxide ion can't combine with
the positive ion until that positive ion has been formed. The second
step is in a sense waiting around for the first slow step to happen.
The slow step of a reaction is known as the rate determining
step.
As long as there is a lot of difference between the rates of the
various steps, when you measure the rate of a reaction, you are
actually measuring the rate of the rate determining step.

Reaction mechanisms and orders of reaction


The examples we use at this level are the very simple ones where
the orders of reaction with respect to the various substances
taking part are 0, 1 or 2. These tend to have the slow step of the
reaction happening before any fast step(s).
To try to explain how fractional orders of reaction can arise is
beyond the scope of UK A' level courses.

Example 1
Here is the mechanism we have already looked at. How do we
know that it works like this?

By doing rate of reaction experiments, you find this rate equation:

The reaction is first order with respect to the organic compound,


and zero order with respect to the hydroxide ions. The
concentration of the hydroxide ions isn't affecting the overall rate of
the reaction.
If the hydroxide ions were taking part in the slow step of the
reaction, increasing their concentration would speed the reaction
up. Since their concentration doesn't seem to matter, they must be
taking part in a later fast step.
Increasing the concentration of the hydroxide ions will speed up the
fast step, but that won't have a noticeable effect on the overall rate
of the reaction. That is governed by the speed of the slow step.

converted by W eb2PDFConvert.com

Note: If you decreased the concentration of hydroxide ions


enough, you will eventually slow down the second step of this
reaction to the point where both steps have similar rates. At
that point, the concentration of the hydroxide ions will matter.
At normal concentrations, the rates of the two steps differ
widely, and so this problem doesn't arise.

In a simple case like this, where the slow step of the reaction is the
first step, the rate equation tells you what is taking part in that slow
step. In this case, the reaction is first order with respect to the
organic molecule - and that's all.
This gives you a starting point for working out a possible
mechanism. Having come up with a mechanism, you would need
to find more evidence to confirm it. For example, in this case you
might try to detect the presence of the positive ion that is formed in
the first step.

Example 2
At first sight this reaction seems identical with the last one. A
bromine atom is being replaced by an OH group in an organic
compound.

However, the rate equation for this apparently similar reaction turns
out to be quite different. That means that the mechanism must be
different.

The reaction this time is first order with respect to both the organic
compound and the hydroxide ions. Both of these must be taking
part in the slow step of the reaction. The reaction must happen by
a straightforward collision between them.

The carbon atom which is hit by the hydroxide ion has a slight
positive charge on it and the bromine a slight negative one
because of the difference in their electronegativities.
As the hydroxide ion approaches, the bromine is pushed off in one
smooth action.
Note: If you are interested in understanding these
mechanisms in more detail, you could follow this link. For the
purposes of this page, all that matters is that the rate
equations show that the two apparently similar reactions
happen by different mechanisms.

Molecularity of a reaction
You may come across an older term known as the molecularity of
a reaction. This has largely dropped out of UK A' level syllabuses,
but if you meet it, it is important that you understand the difference
between this and the order of a reaction. The terms were
sometimes used carelessly as if they mean the same thing - they
don't!

Order of reaction
The important thing to realise is that this is something which can
converted by W eb2PDFConvert.com

only be found by doing experiments. It gives you information about


which concentrations affect the rate of the reaction. You cannot
look at an equation for a reaction and deduce what the order of the
reaction is going to be - you have to do some practical work!
Having found the order of the reaction experimentally, you may be
able to make suggestions about the mechanism for the reaction at least in simple cases.

Molecularity of a reaction
This starts at the other end! If you know the mechanism for a
reaction, you can write down equations for a series of steps which
make it up. Each of those steps has a molecularity.
The molecularity of a step simply counts the number of species
(molecules, ions, atoms or free radicals) taking part in that step.
For example, going back to the mechanisms we've been looking
at:

This step involves a single molecule breaking into ions. Because


only one species is involved in the reaction, it has a molecularity of
1. It could be described as unimolecular.
The second step of this mechanism, involves two ions reacting
together.

This step has a molecularity of 2 - a bimolecular reaction.


The other reaction we looked at happened in a single step:

Because of the two species involved (one molecule and one ion),
this reaction is also bimolecular.
Unless an overall reaction happens in one step (like this last one),
you can't assign it a molecularity. You have to know the
mechanism, and then each individual step has its own
molecularity.
There's nothing the least bit complicated about the term
molecularity. The only confusion is that you may sometimes find it
used as if it meant the same as order. It doesn't!

More about reaction mechanisms and orders of


reaction
Relating orders of reaction to mechanisms is relatively easy where
the slow step is the first step of the reaction mechanism. It isn't so
easy when it is one of the later steps. I want to look at this in a bit
more detail in case your syllabus requires it.
We will revisit the simple case first where the slow step is the first
step of the mechanism.

Cases where the slow step is the first step in the mechanism

converted by W eb2PDFConvert.com

Suppose you had a reaction between A and B, and it turned out


(from doing some experiments) to be first order with respect to
both A and B. So the rate equation is:
Rate = k[A][B]
Which of these two mechanisms is consistent with this
experimental finding?

Mechanism 1

Mechanism 2

Remember that in simple cases, where the slow step is the first
step of the mechanism, the orders tell you what is taking part in the
slow step.
In this case, the reaction is first order with respect to both A and B,
so one molecule of each must be taking part in the slow step. That
means that mechanism 2 is possible.
However, mechanism 1 must be wrong. One molecule of A is
taking part in the slow step, but no B. The rate equation for that
would be:
Rate = k[A]

Note: Be careful! You can't be sure that mechanism 2 is


correct - it may be, but you can't be sure. A and B could, for
example, react to give some sort of intermediate, which went
on to turn into D and E by one or more fast steps. The rate
equation would be the same, because it is governed by the
same slow step, but you can't be sure about what happens
after that.
All you can be sure of is that mechanism 1 is inconsistent
with the rate equation, and so is wrong.

Cases where the slow step isn't the first step in the
mechanism
This is much more difficult to do and explain. I'm going to start with
as simple example as possible

Example 1
Suppose the mechanism for a reaction between A and B looks
like this:

This time the slow step is the second step. Notice that the first
(fast) step is reversible.
I need to assume that the fast step is much faster than the slow
step - for reasons that I'm not going to explain. We are looking well
beyond A level here!

converted by W eb2PDFConvert.com

The rate of the reaction will be governed by the slow step, and so
the rate equation might look like this:

Rate = k[A][X]
. . . except, of course, that X isn't one of the things you are starting
with!
At an introductory level, the flawed discussion often goes
something like this:
X is made from one molecule of A and one molecule of B,
and so its concentration will depend on the concentrations of
A and B.
That means that you can replace [X] by [A][B].
That gives a rate equation:
Rate = k[A][A][B] = k[A]2[B]
As it happens, that gives the right answer in this case, but this
simplistic view doesn't work in all cases - as I will show below.

So let's start again and do it better!


The rate of the reaction will be governed by the slow step, and so
the rate equation might look like this:

Rate = k[A][X]
. . . except, of course, that X isn't one of the things you are starting
with!
You need to be able to express the concentration of X in terms of
[A] and [B], and you can do that because the first step is an
equilibrium.
The equilibrium constant for the first reaction is:

You can rearrange this to give an expression for [X]:

. . . and then substitute this value into the rate expression we


started with:

If you sort this out, and combine the two different constants into a
new rate constant, you get the rate expression:

So the reaction is second order with respect to A and first order


with respect to B.

Example 2
I am going to take another similar-looking example now, chosen to
be deliberately awkward. This is to try to show that you can't
reliably work these problems out just by looking at them.
This relates to the following reaction:

The mechanism for this is:

converted by W eb2PDFConvert.com

From the slow step of the reaction, the rate equation would like like
this:

Rate = k[C][X]
Up to now, this looks just the same as the previous example - but it
isn't!
Let's do it properly, and think about the equilibrium expression for
the first step in order to find a value for [X] that we can substitute
into the rate equation.

The problem is that we have now got an extra variable in this


equation that wasn't there before - we have got the concentration
of Y. What do we do about that?
Well, for every mole of X formed, a mole of Y will be formed as
well.
Provided that there was no X or Y in the mixture to start with, that
means that the concentration of Y is equal to the concentration of
X, and so we can substitute a value of [X] in place of the [Y], giving:

Rearranging that to get an expression for [X] gives:

And now you can substitute this into the rate equation:

. . . and then tidy it up by combining the two constants into a single


new rate constant:

The order is 0.5 with respect to both A and B, and 1 with respect to
C.

There is no way you could have come up with that answer if you
had just looked at the equations and assumed that the
concentration of X is proportional to the concentrations of A and B.
The presence of the extra Y makes a serious difference to the
result - but just looking at the equations, that isn't at all obvious.

converted by W eb2PDFConvert.com

Note: Syllabuses often aren't at all clear about what the


examiners actually want with regards to this. You have to
look at past papers and mark schemes. Of the UK-based
exams, I know that CIE have asked a question relating to a
mechanism where the slow step was either the second or
third in a series of three steps. But they were simply
expecting students to just look at the equations, and not
work it out properly.
Personally, I think this is misleading and wrong.

Questions to test your understanding


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

Where would you like to go now?


To the rates of reaction menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002 (last modified October 2013)

converted by W eb2PDFConvert.com

ENERGY PROFILES FOR SIMPLE


REACTIONS
This page takes a closer look at simple energy profiles for
reactions, and shows how they are slightly different for reactions
involving an intermediate or just a transition state. Both of those
terms are explained as well.

Types of Energy Profile


What is an energy profile?
If you have done any work involving activation energy or catalysis,
you will have come across diagrams like this:

This diagram shows that, overall, the reaction is exothermic. The


products have a lower energy than the reactants, and so energy is
released when the reaction happens.
It also shows that the molecules have to possess enough energy
(called activation energy) to get the reactants over what we think of
as the "activation energy barrier".
In this example of a reaction profile, you can see that a catalyst
offers a route for the reaction to follow which needs less activation
energy. That, of course, causes the reaction to happen faster.
Note: If you aren't very happy about this, read the page
about catalysts before you go on.
Use the BACK button on your browser to return to this page,
or come back via the rates of reaction menu.

Diagrams like this are described as energy profiles. In the


diagram above, you can clearly see that you need an input of
energy to get the reaction going. Once the activation energy
barrier has been passed, you can also see that you get even more
energy released, and so the reaction is overall exothermic.
If you had an endothermic reaction, a simple energy profile for a
non-catalysed reaction would look like this:

converted by W eb2PDFConvert.com

Unfortunately, for many reactions, the real shapes of the energy


profiles are slightly different from these, and the rest of this page
explores some simple differences. What matters is whether the
reaction goes via a single transition state or an intermediate. We
will look at these two different cases in some detail.

Energy profiles for reactions which go via a single transition


state only
This is much easier to talk about with a real example. The equation
below shows an organic chemistry reaction in which a bromine
atom is being replaced by an OH group in an organic compound.
The starting compound is bromoethane, and the organic product is
ethanol.

During the reaction one of the lone pairs of electrons on the


negatively charged oxygen in the -OH group is attracted to the
carbon atom with the bromine attached.
That's because the bromine is more electronegative than carbon,
and so the electron pair in the C-Br bond is slightly closer to the
bromine. The carbon atom becomes slightly positively charged
and the bromine slightly negative.

As the hydroxide ion approaches the slightly positive carbon, a


new bond starts to be set up between the oxygen and the carbon.
At the same time, the bond between the carbon and bromine starts
to break as the electrons in the bond are repelled towards the
bromine.
At some point, the process is exactly half complete. The carbon
atom now has the oxygen half-attached, the bromine half-attached,
and the three other groups still there, of course.

And then the process completes:

converted by W eb2PDFConvert.com

Note: These diagrams have been simplified in various ways


to make the process clearer. For example, the true
arrangement of the lone pairs of electrons around the oxygen
in the first diagram has been simplified for clarity. The
bromine also has 3 lone pairs as well as the bonding pair, but
they play no part. And, of course, the other groups attached
to the carbon have been left out in order to concentrate on
what is important.

The second diagram where the bonds are half-made and halfbroken is called the transition state, and it is at this point that the
energy of the system is at its maximum. This is what is at the top of
the activation energy barrier.

But the transition state is entirely unstable. Any tiny change in


either direction will send it either forward to make the products or
back to the reactants again. Neither is there anything special about
a transition state except that it has this maximum energy. You can't
isolate it, even for a very short time.
The situation is entirely different if the reaction goes through an
intermediate. Again, we'll look at a specific example.

Energy profiles for reactions which go via an intermediate


For reasons which you may well meet in the organic chemistry part
of your course, a different organic bromine-containing compound
reacts with hydroxide ions in an entirely different way.
In this case, the organic compound ionises slightly in a slow
reaction to produce an intermediate positive organic ion. This then
goes on to react very rapidly with hydroxide ions.

converted by W eb2PDFConvert.com

Note: If you haven't come across the use of curly arrows in


organic chemistry yet, all you need to know for now is that
they show the movement of a pair of electrons. In the first
equation, for example, the bonding pair of electrons in the CBr bond moves entirely on to the bromine to make a bromide
ion. In the second equation, a lone pair on the hydroxide ion
moves towards the positive carbon to form a covalent bond.

The big difference in this case is that the positively charged


organic ion can actually be detected in the mixture. It is very
unstable, and soon reacts with a hydroxide ion (or picks up its
bromide ion again). But, for however short a time, it does have a
real presence in the system. That shows itself in the energy profile.

The stability (however temporary and slight) of the intermediate is


shown by the fact that there are small activation barriers to its
conversion either into the products or back into the reactants
again.
Notice that the barrier on the product side of the intermediate is
lower than that on the reactant side. That means that there is a
greater chance of it finding the extra bit of energy to convert into
products. It would need a greater amount of energy to convert back
to the reactants again.
I've labelled these peaks "ts1" and "ts2" - they both represent
transition states between the intermediate and either the reactants
or the products. During either conversion, there will be some
arrangement of the atoms which causes an energy maximum that's all a transition state is.

And finally . . .
It is perfectly possible to get reactions which take several steps going through a number of different intermediates and transition
states. In cases like this, you would end up with a whole "mountain
range" of peaks, some of which might be simple transition states,
and others with the little dips which hold intermediates. You
wouldn't expect to come across problems like this at levels
equivalent to UK A level.

Questions to test your understanding


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

Where would you like to go now?


converted by W eb2PDFConvert.com

To the rates of reaction menu . . .


To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2010 (modified October 2013)

converted by W eb2PDFConvert.com

FINDING ORDERS OF REACTION


EXPERIMENTALLY
This page is an introduction to some of the experimental methods
that can be used in school labs to find orders of reaction.
There are two fundamentally different approaches to this - you can
either investigate what happens to the initial rate of the reaction as
you change concentrations, or you can follow a particular reaction
all the way through, and process the results from that single
reaction. We will look at these two approaches separately. Don't
expect full practical details.
This is going to be a very long page. I wouldn't really recommend
that you try to read it all in one go.

Initial rate experiments


How initial rate experiments work
An outline of the experiments
The simplest initial rate experiments involve measuring the time
taken for some easily recognisable event to happen very early on
in a reaction.
This could include the time taken for, say, 5 cm3 of gas to be
produced. Or it could be the time taken for a small measurable
amount of precipitate to be formed. Or you could measure the time
taken for some dramatic colour change to occur. We will look at
examples of all these below.
You then change the concentration of one of the components of the
reaction, keeping everything else constant - the concentrations of
other reactants, the total volume of the solution and the
temperature and so on. Then you find the time taken for the same
event to take place with that new concentration.
This is repeated for a range of concentrations of the substance you
are interested in. You would need to cover a reasonably wide
range of concentrations, taking perhaps 5 or so different
concentrations varying from the original one down to half of it or
less.
Obviously, you could then repeat the process by changing
something else - the concentration of a different substance, or the
temperature, for example.

Understanding the results


We will take a simple example of an initial rate experiment where
you have a gas being produced. This could be a reaction between
a metal and an acid, for example, or the catalytic decomposition of
hydrogen peroxide.
If you plotted the volume of gas given off against time, you would
probably get the first graph below.

converted by W eb2PDFConvert.com

A measure of the rate of the reaction at any point is found by


measuring the slope of the graph. The steeper the slope, the faster
the rate. Since we are interested in the initial rate, we would need
the slope at the very beginning.
If you then look at the second graph, enlarging the very beginning
of the first curve, you will see that it is approximately a straight line
at that point. That is only a reasonable approximation if you are
considering a very early stage in the reaction. The further into the
reaction you go, the more the graph will start to curve.
Measuring the slope of a straight line is very easy. The slope in this
case is simply V/t.
Now suppose you did the experiment again with a different (lower)
concentration of the reagent. Again, we will measure the time
taken for the same volume of gas to be given off, and so we are
still just looking at the very beginning of the reaction:

The initial rates (in terms of volume of gas produced per second)
are:
experiment 1

experiment 2
Now suppose you didn't actually know what the volume V was.
Suppose, for example, that instead of measuring the time taken to
collect 5 cm3 of gas, you just collected the gas up to a mark which
you had made on the side of a test tube. Does it matter?
If you are simply wanting to compare initial rates, then it doesn't

converted by W eb2PDFConvert.com

matter. If you look at the expressions in the table above, you should
recognise that the initial rate is inversely proportional to the time
taken. In symbols:

In experiments of this sort, you often just use 1/t as a measure of


the initial rate without any further calculations.
You can then plot 1/t as a measure of rate against the varying
concentrations of the reactant you are investigating. If the reaction
is first order with respect to that substance, then you would get a
straight line. That's because in a first order reaction, the rate is
proportional to the concentration.

If you get a curve, then it isn't first order. It might be second order but it could equally well have some sort of fractional order like 1.5
or 1.78.
The best way around this is to plot what is known as a "log graph".
The maths of this might not be familiar to you, but you may find that
you are asked to do this as a part of a practical exam or practical
exercise. If it is an exam, you would probably be given help as to
how to go about it.
The maths goes like this:
If you have a reaction involving A, with an order of n with respect to
A, the rate equation says:

rate = k [A]n
If you take the log of each side of the equation, you get:

log(rate) = log k + n log[A]


If you plotted log(rate) agains log[A], this second equation would
plot as a straight line with slope n. If you measure the slope of this
line, you get the order of the reaction.
So you would convert all the values you had for rate into log(rate).
Convert all the values for [A] into log[A], and then plot the graph.
This should be a straight line. If it isn't, then you have done
something wrong! Measure the slope to find the order, n.
Note: Don't worry if you don't understand logs (logarithms),
or how I got from the first equation to the second one! I
suspect that in the unlikely event of you needing it in an
exam at this level, it would be given to you.
All you need to do is find the log button on your calculator
and use it to convert your numbers. Practise to start with by
trying to find log 2. It should give a value of 0.3010(etc). You
probably have to enter 2 and then press the log button, but on
some calculators it might be the other way around. If you do
it the wrong way around, you will just get an error message.

Some sample reactions


converted by W eb2PDFConvert.com

The catalytic decomposition of hydrogen peroxide


This is a simple example of measuring the initial rate of a reaction
producing a gas.

A simple set-up to do this might be:

The reason for the weighing bottle containing the catalyst is to


prevent introducing errors at the beginning of the experiment.
Since this is the part of the reaction you are most interested in,
introducing errors here would be stupid!
You have to find a way of adding the catalyst to the hydrogen
peroxide solution without changing the volume of gas collected. If
you added it to the flask using a spatula, and then quickly put the
bung in, you might lose some gas before you got the bung in.
Alternatively, as you pushed the bung in, you might force some air
into the measuring cylinder. Either way, it makes your results
meaningless.
To start the reaction, you just need to shake the flask so that the
weighing bottle falls over, and then continue shaking to make sure
the catalyst mixes evenly with the solution.
You could also use a special flask with a divided bottom, with the
catalyst in one side, and the hydrogen peroxide solution in the
other. They are easy to mix by tipping the flask.
If you use a 10 cm3 measuring cylinder, initially full of water, you
can reasonably accurately record the time taken to collect a small
fixed volume of gas.
You could, of course, use a small gas syringe instead.
If you were looking at the effect of the concentration of hydrogen
peroxide on the rate, then you would have to change its
concentration, but keep everything else constant.
The temperature would have to be kept constant, so would the total
volume of the solution and the mass of manganese(IV) oxide. You
would also have to be sure that the manganese(IV) oxide used
always came from the same bottle so that its state of division was
always the same.
You could, of course, use much the same apparatus to find out
what happened if you varied the temperature, or the mass of the
catalyst, or the state of division of the catalyst.

The thiosulphate-acid reaction


If you add dilute hydrochloric acid to sodium thiosulphate solution,
you get the slow formation of a pale yellow precipitate of sulphur.

There is a very simple, but very effective, way of measuring the


time taken for a small fixed amount of precipitate to form. Stand
the flask on a piece of paper with a cross drawn on it, and then
look down through the solution until the cross disappears.

converted by W eb2PDFConvert.com

So . . . you put a known volume of sodium thiosulphate solution in a


flask. Then you add a small known volume of dilute hydrochloric
acid, start timing, swirl the flask to mix everything up, and stand it
on the paper with the cross on. Time how long it takes for the cross
to disappear.
Then repeat using a smaller volume of sodium thiosulphate, but
topped up to the same original volume with water. Everything else
should be exactly as before.
If you started with, say, 50 cm3 of sodium thiosulphate solution, you
would repeat the experiment with perhaps, 40, 30, 20, 15 and 10
cm3 - each time made up to a total of 50 cm3 with water.
The actual concentration of the sodium thiosulphate doesn't have
to be known. In each case, you could record its relative
concentration. The solution with 40 cm3 of sodium thiosulphate
solution plus 10 cm3 of water has a concentration which is 80% of
the original one, for example. The one with 10 cm3 of sodium
thiosulphate solution plus 40 cm3 of water has a concentration
which is 20% of the original one.
When you came to plotting a rate against concentration graph, as
we looked at further up the page, you would plot 1/t as a measure
of the rate, and volume of sodium thiosulphate solution as a
measure of concentration. Alternatively, you could plot relative
concentrations - from, say, 20% to 100%. It doesn't actually matter
- the shape of the graph will be identical.
You could also look at the effect of temperature on this reaction, by
warming the sodium thiosulphate solution before you added the
acid. Take the temperature after adding the acid, though, because
the cold acid will cool the solution slightly.
This time you would change the temperature between
experiments, but keep everything else constant. To get reasonable
times, you would have to use a diluted version of your sodium
thiosulphate solution. Using the full strength solution hot will
produce enough precipitate to hide the cross almost instantly.

Iodine clock reactions


There are several reactions which go under the name "iodine
clock". They are all reactions which give iodine as one of the
products. This is the simplest of them, but only because it involves
the most familiar reagents.
The reaction we are looking at is the oxidation of iodide ions by
hydrogen peroxide under acidic conditions.

The iodine is formed first as a pale yellow solution darkening to


orange and then dark red, before dark grey solid iodine is
precipitated.
There is a very clever way of picking out a when a particular very
small amount of iodine has been formed.
Iodine reacts with starch solution to give a very deep blue solution.
If you added some starch solution to the reaction above, as soon
as the first trace of iodine was formed, the solution would turn blue.
That doesn't actually help!
However, iodine also reacts with sodium thiosulphate solution.

If you add a very small amount of sodium thiosulphate solution to


your reaction mixture (including the starch solution), it will react with
the iodine that is initially produced, and so the iodine won't affect
converted by W eb2PDFConvert.com

the starch, and you won't get any blue colour.


However, when that small amount of sodium thiosulphate has been
used up, there is nothing to stop the next lot of iodine produced
from reacting with the starch. The mixture suddenly goes blue.
Note: There is a neat piece of video on YouTube showing an
iodine clock reaction (not necessarily the one I am talking
about here, but it doesn't matter).
It shows three reactions side by side:
The right-hand one is done at room temperature.
The left-hand one is also done at room temperature,
but at three times the concentration of one of the
reagents.
The central one is colder than room temperature, and
presumably at the same concentration as the righthand one.
The blue colours appear in exactly the order you would
predict.

In our example, you could obviously look at the effect of changing


the hydrogen peroxide concentration, or the iodide ion
concentration, or the hydrogen ion concentration - each time, of
course, keeping everything else constant.
That would let you find the orders with respect to everything taking
part in the reaction.

Following the course of a single reaction


Rather than doing a whole set of initial rate experiments, you can
also get information about orders of reaction by following a
particular reaction from start to finish.
There are two different ways you can do this. You can take
samples of the mixture at intervals and do titrations to find out how
the concentration of one of the reagents is changing. Or (and this
is much easier!) you can measure some physical property of the
reaction which changes as the reaction continues - for example,
the volume of gas produced.
We need to look at these two different approaches separately.

Sampling the reaction mixture


Bromoethane reacts with sodium hydroxide solution as follows:

During the course of the reaction, both bromoethane and sodium


hydroxide will get used up. However, it is relatively easy to
measure the concentration of the sodium hydroxide at any one
time by doing a titration with some standard acid - for example,
with hydrochloric acid of a known concentration.
You start with known concentrations of sodium hydroxide and
bromoethane, and usually it makes sense to have them both the
same. Because the reaction is 1:1, if the concentrations start the
same as each other, they will stay the same as each other all
through the reaction.
So all you need to do is to take samples using a pipette at regular
intervals during the reaction, and titrate them with standard
hydrochloric acid in the presence of a suitable indicator.

converted by W eb2PDFConvert.com

That is a lot easier said than done!


The problem is that the reaction will still be going on in the time it
takes for you to do the titration. And, of course, you only get one
attempt at the titration. By the time you take another sample, the
concentration of everything will have changed!
There are two ways around this.
You can slow the reaction down by diluting it, adding your sample
to a larger volume of cold water before you do the titration. Then do
the titration as quickly as possible. That's most effective if you are
doing your reaction at a temperature above room temperature.
Cooling it as well as diluting it will slow it down even more.
But if possible (and it is possible in the case we are talking about)
it is better to stop the reaction completely before you do the
titration.
In this case, you can stop it by adding the sample to a known
volume (chosen to be an excess) of standard hydrochloric acid.
That will use up all the sodium hydroxide in the mixture so that the
reaction stops.
Now you would titrate the resulting solution with standard sodium
hydroxide solution, so that you can find out how much hydrochloric
acid is left over in the mixture.
That lets you calculate how much was used up, and so how much
sodium hydroxide must have been present in the original reaction
mixture.
This sort of technique is known as a back titration. These
calculations can be quite confusing to do without some guidance. If
you are interested, you will find back titrations discussed on pages
72-75 of my chemistry calculations book.

Processing the results


You will end up with a set of values for concentration of (in this
example) sodium hydroxide against time. The concentrations of
the bromoethane are, of course, the same as these if you started
with the same concentrations of each reagent.
You can plot these values to give a concentration-time graph which
will look something like this:

Now it all gets pretty tedious!


You need to find the rates of reaction at a number of points on the
graph, and you do this by drawing tangents to the graph, and
measuring their slopes.

converted by W eb2PDFConvert.com

You would then draw up a simple table of rate against


concentration.
The quickest way to go on from here is to plot a log graph as
described further up the page. You would convert all your rates into
log(rate), and all the concentrations into log(concentration). Then
plot log(rate) against log(concentration).
The slope of the graph gives you the order of reaction.
In our example of the reaction between bromoethane and sodium
hydroxide solution, the order would turn out to be 2.
Note: This all takes ages to do - not just the practical which

Notice that thiswould


is theprobably
overalltake
order
of theanreaction
the
at least
hour, but -allnot
thejust
graph
order with respect
to the
whose
concentration
you were
drawing,
andreagent
processing
the results
from the graphs.
There is
obvious
way this could
be asked
in any normal
measuring. Thenorate
of reaction
was falling
because
the practical or
at this
level. were falling.
concentrations written
of bothexam
of the
reactants
I can see that it is just possible that you might be asked in
principle how you would do it, but actually doing it could only
reasonably be a part of a coursework exercise.

Following the course of the reaction using a physical


property
An example where a gas is given off
A familiar example of this is the catalytic decomposition of
hydrogen peroxide that we have already looked at above as an
example of an initial rate experiment.

This time, you would measure the oxygen given off using a gas
syringe, recording the volume of oxygen collected at regular
intervals.
So the practical side of this experiment is straightforward, but the
calculation isn't.
The problem is that you are measuring the volume of product,
whereas to find an order of reaction you have to be working in
terms of the concentration of the reactants - in this case, hydrogen
peroxide.
That means that you will have to work out the concentration of
hydrogen peroxide remaining in the solution for each volume of
oxygen you record. To do this, you have to be happy with
calculations involving the ideal gas law, and also basic mole
calculations.
Having got a table of concentrations against time, you will then
process them in exactly the same way as I described above. Plot
the graph, draw tangents to find rates at various concentrations,
and then plot a log graph to find the order.

converted by W eb2PDFConvert.com

Note: It seems to me fairly unlikely that you could ever be


asked to do this in an exam situation. And, at this level, you
would almost certainly be given some guidance with the
calculations
In a practical exam, few schools could provide a class set of
the very expensive gas syringes accurate enough to produce
meaningful results, and the time taken to process the results
would be far greater than was available in any normal exam.
That is equally true of a theory paper.
If you know that you can follow the course of a reaction which
produces a gas using this method, that is probably all you
will need. But check your syllabus, and past papers and
mark schemes.

Colorimetry
In any reaction involving a coloured substance (either reacting or
being produced), you can follow the course of the reaction using a
colorimeter.

All of this is contained in one fairly small box.


The colour of the light can be changed by selecting a particular
coloured filter (or using some more sophisticated device like a
diffraction grating). The colour is chosen so that it is the frequency
of light which is absorbed by the sample.
Taking copper(II) sulphate solution as a familiar example, you
would choose to use a red filter, because copper(II) sulphate
solution absorbs red light. The more concentrated the solution is,
the more of the red light it will absorb.
Note: For an explanation of why absorbing red light makes
copper(II) sulphate solution blue, see the first part of the page
about the colours of complex metal ions. You don't need to
read about the origin of the colour for now.

converted by W eb2PDFConvert.com

A commonly quoted example of the use of colorimetry in rates of


reaction is the reaction between propanone and iodine in the
presence of an acid catalyst.

The solution of iodine in propanone starts off brown, and then


fades through orange to yellow to colourless as the iodine is used
up.
A colorimeter lets you measure the amount of light which is
absorbed as it passes through a solution - recorded as the
absorbance of the solution.
It is common to plot a calibration curve for a colorimeter by
making up solutions of the coloured substance of known
concentration and then measuring the absorbance of each under
the same conditions as you will do the experiment. You then plot a
graph of absorbance against concentration to give your calibration
curve.
During your rate of reaction experiment, you read the absorbance
from the meter at regular intervals, and then use your calibration
curve to convert those values into concentrations.
Then you are faced with the same graphical methods as before.
Note: In truth, these days, you are more likely to plug your
colorimeter into a computer with the right software to do it all
for you!

Two other methods

pH measurements
If you have a reaction in which hydrogen ions are reacting or being
produced, in principle you should be able to follow changes in their
concentration using a pH meter.
You may be aware that pH is a measure of hydrogen ion
concentration, and it isn't difficult to calculate an actual hydrogen
ion concentration from a pH.
However, if you are measuring pH over a fairly narrow range of
hydrogen ion concentrations, the pH doesn't change all that much.
For example, the pH of a solution containing 0.2 mol dm-3 H+ has
a pH of 0.70. By the time that the concentration has fallen to 0.1
mol dm-3, the pH has only increased to 1.00.
Whether it is feasible to use a pH meter obviously depends on how
accurate it is. If the pH meter only recorded to 0.1 pH units, your
results aren't going to be very good.

Conductivity measurements
The electrical conductivity of a liquid depends on the number of
ions present, and the nature of the ions. For example, we looked at
this reaction much further up the page:

During the course of the reaction, as hydrogen ions and iodide


ions get used up, the conductivity of the mixture will fall.

converted by W eb2PDFConvert.com

Note: I am not giving any more detail on this, because


conductivity measurements aren't a part of any of the
syllabuses that I am tracking. CIE expect you to know that it
is possible to use conductivity measurements to follow the
course of a reaction involving changes in the ions present, but
not how you would actually carry out the experiments or
process the results.

Questions to test your understanding


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

Where would you like to go now?


To the rates of reaction menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2011 (modified October 2013)

converted by W eb2PDFConvert.com

RATE CONSTANTS AND THE


ARRHENIUS EQUATION
This page looks at the way that rate constants vary with
temperature and activation energy as shown by the Arrhenius
equation.
Note: If you aren't sure what a rate constant is, you should
read the page about orders of reaction before you go on. This
present page is at the hard end of the rates of reaction work
on this site. If you aren't reasonably confident about the basic
rates of reaction work, explore the rates of reaction menu
first.

The Arrhenius equation


Rate constants and rate equations
You will remember that the rate equation for a reaction between
two substances A and B looks like this:

Note: If you don't remember this, you must read the page
about orders of reaction before you go on. Use the BACK
button on your browser to return to this page.

converted by W eb2PDFConvert.com

The rate equation shows the effect of changing the concentrations


of the reactants on the rate of the reaction. What about all the other
things (like temperature and catalysts, for example) which also
change rates of reaction? Where do these fit into this equation?
These are all included in the so-called rate constant - which is only
actually constant if all you are changing is the concentration of the
reactants. If you change the temperature or the catalyst, for
example, the rate constant changes.
This is shown mathematically in the Arrhenius equation.

The Arrhenius equation

What the various symbols mean


Starting with the easy ones . . .

Temperature, T
To fit into the equation, this has to be meaured in kelvin.

The gas constant, R


This is a constant which comes from an equation, pV=nRT,
which relates the pressure, volume and temperature of a
particular number of moles of gas. It turns up in all sorts of
unlikely places!

Activation energy, EA
This is the minimum energy needed for the reaction to occur.
To fit this into the equation, it has to be expressed in joules
per mole - not in kJ mol-1.

Note: If you aren't sure about activation energy, you should


read the introductory page on rates of reaction before you go
on. Use the BACK button on your browser to return to this
page.

And then the rather trickier ones . . .

e
This has a value of 2.71828 . . . and is a mathematical
number, a bit like pi. You don't need to worry exactly what it
means, although if you have to do calculations with the
Arrhenius equation, you may have to find it on your calculator.
You should find an ex button - probably on the same key as
"ln".

The expression, e-(EA / RT)


For reasons that are beyond the scope of any course at this
level, this expression counts the fraction of the molecules
present in a gas which have energies equal to or in excess
of activation energy at a particular temperature. You will find

converted by W eb2PDFConvert.com

a simple calculation associated with this further down the


page.

The frequency factor, A


You may also find this called the pre-exponential factor.
A is a term which includes factors like the frequency of
collisions and their orientation. It varies slightly with
temperature, although not much. It is often taken as constant
across small temperature ranges.
By this time you've probably forgotten what the original Arrhenius
equation looked like! Here it is again:

You may also come across it in a different form created by a


mathematical operation on the standard one:

"ln" is a form of logarithm. Don't worry about what it means. If you


need to use this equation, just find the "ln" button on your
calculator.

Using the Arrhenius equation

The effect of a change of temperature


You can use the Arrhenius equation to show the effect of a change
of temperature on the rate constant - and therefore on the rate of
the reaction. If the rate constant doubles, for example, so also will
the rate of the reaction. Look back at the rate equation at the top of
this page if you aren't sure why that is.
What happens if you increase the temperature by 10C from, say,
20C to 30C (293 K to 303 K)?
The frequency factor, A, in the equation is approximately constant
for such a small temperature change. We need to look at how e-(EA
/ RT) changes - the fraction of molecules with energies equal to or
in excess of the activation energy.
Let's assume an activation energy of 50 kJ mol-1. In the equation,
we have to write that as 50000 J mol-1. The value of the gas
constant, R, is 8.31 J K-1 mol-1.
At 20C (293 K) the value of the fraction is:

By raising the temperature just a little bit (to 303 K), this increases:

You can see that the fraction of the molecules able to react has
almost doubled by increasing the temperature by 10C. That
causes the rate of reaction to almost double. This is the value in
the rule-of-thumb often used in simple rate of reaction work.

converted by W eb2PDFConvert.com

Note: This approximation (about the rate of a reaction


doubling for a 10 degree rise in temperature) only works for
reactions with activation energies of about 50 kJ mol-1 fairly
close to room temperature. If you can be bothered, use the
equation to find out what happens if you increase the
temperature from, say 1000 K to 1010 K. Work out the
expression -(EA / RT) and then use the ex button on your
calculator to finish the job.
The rate constant goes on increasing as the temperature
goes up, but the rate of increase falls off quite rapidly at
higher temperatures.

The effect of a catalyst


A catalyst will provide a route for the reaction with a lower
activation energy. Suppose in the presence of a catalyst that the
activation energy falls to 25 kJ mol-1. Redoing the calculation at
293 K:

If you compare that with the corresponding value where the


activation energy was 50 kJ mol-1, you will see that there has been
a massive increase in the fraction of the molecules which are able
to react. There are almost 30000 times more molecules which can
react in the presence of the catalyst compared to having no
catalyst (using our assumptions about the activation energies).
It's no wonder catalysts speed up reactions!

Note: If you read this carefully, you should notice that I am


not saying that the reaction will be 30000 times faster. There
may well be 30000 times more molecules which can react,
but it is highly likely that the frequency factor will have
changed in the presence of the catalyst. And the rate
constant k is just one factor in the rate equation. You won't
just have the original reactants present as before. The
catalyst is bound to be involved in the slow step of the
reaction, and a new rate equation will have to include a term
relating to the catalyst.
Nevertheless, the catalysed reaction is still going to be a lot
faster than the uncatalysed one because of the huge increase
in sufficiently energetic molecules.

Other calculations involving the Arrhenius equation


If you have values for the rate of reaction or for the rate constant at
different temperatures, you can use these to work out the activation
energy of the reaction. Only one UK A' level Exam Board expects
you to be able to do these calculations. They are included in my
chemistry calculations book, and I can't repeat the material on this
site.

converted by W eb2PDFConvert.com

Note: There is no way of making this sufficiently different


from what is in the book to avoid being in breach of contract
with my publishers if I included it on this site.
If you are interested in my chemistry calculations book you
might like to follow this link.

Questions to test your understanding


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

Where would you like to go now?


To the rates of reaction menu . . .
To the Physical Chemistry menu . . .
To Main Menu . . .

Jim Clark 2002 (modified October 2013)

converted by W eb2PDFConvert.com

Das könnte Ihnen auch gefallen