You are on page 1of 100

1.

kinetics
Chemical
Kinetics is the study of the rates of chemical processes in an effort to
understand what it is that influences these rates and to develop theories
which can be used to predict them. A knowledge of reaction rates has many
practical applications, for example in designing an industrial process, in
understanding the complex dynamics of the atmosphere and in
understanding the intricate interplay of the chemical reactions that are the
basis of life.
At a more fundamental level we want to understand what happens to the
molecules in a chemical reaction – that is what happens in a single reactive
encounter between two reagent molecules. By understanding this we may
be able to develop theories that can be used to predict the outcome and rate
of reactions.

1.1 Books
Any general physical chemistry text (such as P W Atkins Physical
Chemistry, OUP, any edition) will have several chapters on chemical
kinetics. There is a nice small book about kinetics in the Oxford Chemistry
Primers series: B G Cox Modern Liquid Phase Kinetics, OUP, 1994. For
more detail (beyond the scope of this course), Reaction Kinetics by M
J Pilling and P W Seakins (OUP 1995) is a good source.

2. Rates, rate laws and rate constants


In this Section we will introduce the language and terms used to describe
the rates of chemical reactions. At this stage we will not be concerned with
the theory of reactions or mechanisms, but just stand back and describe the
overall rates.

2.1 Rate of reaction


The rate is defined as
concentration, c

change in concentration, c, in time t


t c

We can talk about the rate of formation or loss of any species – reactant, t

intermediate or product. It is, however, important to specify which species time, t


we are talking about. The rate can be positive or negative: a positive rate
means that the concentration is increasing with time e.g. a product; a
negative rate means that the concentration is falling with time e.g. a
reactant.
The rate may vary with time (and concentration), so it is usual to define
the rate over a very small time, t. We think of the rate as the derivative of
concentration

concentration with respect to time

rate  dconcentration
dt
time
This derivative is the slope of a graph of concentration against time, taken
The rate is the instantaneous
at a particular time. slope, and this varies with time.

1
From its definition it is clear that the units of a rate are concentration per
unit time, for example mol dm –3 s–1. There are other measures of
concentration, for example in the gas phase pressure is proportional to
concentration, so a rate can be expressed in torr min –1 (1 torr = 1 mm Hg, a
measure of pressure). It is also common to express concentration not in
moles per unit volume but in molecules per unit volume, so the rate would
be expressed in molecules dm–3 s–1 or molecules cm–3 s–1.
Example 1
Exercise 1
Symbolically, concentration is often indicated by square brackets. So
[A] means the concentration of species A, and [Br2] means the
concentration of bromine, and so on.

2.1.1 Relation to stoichiometric equations


Consider the reaction whose stoichiometric equation is
O2  2H2  2H2O
The stoichiometric equation shows how the number of moles of reactants
and products are related; it must be balanced.
This equation says that to form two moles of water, one mole of oxygen
and two moles of hydrogen must react. It follows that the rate of
consumption of hydrogen is twice that of oxygen.
d H 2  d O 2 
dt  2 dt
The rate of formation of water is twice the rate of loss of oxygen, as two
moles of water are formed from one mole of oxygen. As water is the
product, the rate of change of its concentration is positive i.e. the
concentration is increasing with time; however, the rates of change of the
concentrations of both hydrogen and oxygen are negative as these
concentrations are decreasing with time. In derivative form:
d H 2 O  d O 2 or, 1 dH 2O d O 2 
 2 equivalently  

dt dt 2 dt dt
Example 2
Exercise 2

A warning: not everyone or


every book is careful about
using this convention, and you
will find that the "rate" is often
used in a loose sense, rather
than being defined in the
precise way given here. Be
prepared to be flexible.

2
With all these reaction no matter which species is considered.
different "rates",
different for each
species, how can we
define the "rate of
the reaction"?
The usual
procedure is to
include the
stoichiometric
coefficients in the
definition of the
rate. These
coefficients are the
numbers in front of
the species in the
balanced chemical
equation. So the
stoichiometric
coefficient is 1 for
oxygen and 2 for
both hydrogen and
water.
The rate of reaction,
r, is defined for any
species A as

rd

[1]

where A is the
stoichiometric
coefficient of
species A in the
balanced chemical
equation. In
addition the
stoichiometric
coefficients are
given negative signs
for reagents and
positive signs for
products. This
definition ensures
that the rate is
always positive and
the same for a given
3
2.2 Rate laws and rate constants
Experimentally it is found that rates depend on the concentrations of the
species involved in the reaction equation (and sometimes on the
concentrations of species which do not appear at first sight to be involved!).
The relation between the rate and these concentrations can often be
expressed mathematically in the form of an equation called a rate law.
Some rate laws are very simple and some are very complicated. A rate law
may be determined experimentally (Section 4) or may be the result of a
theoretical prediction, or both.
Often reactions are found to have rate laws of the form
r  k A  B …
a b

where k is constant, characteristic of a particular reaction, called the rate


constant or the rate coefficient. The powers a, b ... are also constants: a is
the order with respect to A, b is the order with respect to B. The overall
order is (a + b + ...). Orders are commonly integers, but they do not have to
be.
For example, the solution reaction

N N
Ph Br N+ Br–
N

Ph
PhCH2Br DABCO
is found to have the rate law
r  k PhCH2 Br DABCO
The reaction is second order overall, and first order with respect to both
PhCH2Br and DABCO.
Some reactions have very much more complex rate laws. For example,
in the gas phase, the reaction
H2  Br2  2HBr
has the rate law (Section
6.2)
k H Br  32

r  Br   k HBr
a 2 2

2 b

For this reaction, no order with respect to Br2, and no overall order, can be
defined.
An enzyme, E, catalysing the conversion of a substrate, S, to products is
often found to obey the rate law (Section 5.5.4)

kcat[E]0 S
r
S  KM
Again, no order with respect to S can be defined.
2.2.1 Simple rate laws
A first order rate law is one in which the rate is proportional to the
concentration raised to the power 1 (hence "first")
r  k1st
A 1
k1st is called a first order rate constant. The rate equation, and the reaction it
describes, is said to be first order in A or it is said that the order with
respect to A is one. As the units of r are concentration time–1 and the units
of [A] are concentration, the units of k1st are found as

r  k1st A1 so rconc. time–1


k1st    time–1
A conc.
Hence first order rate constants have units of time–1.
A second order reaction has the concentration raised to the power of 2
r  k2nd A
2

k2nd is called a second order rate constant. The rate equation, and the
reaction it describes, is said to be second order in A, or the order with
respect to A is two. The units of k2nd are found to be conc.–1 time–1.
Example 3
Exercises 3 & 4 2.2.2 Importance of rate laws
1. If we know the rate law and the constants in it we can use this to predict
the rate for any set of conditions (concentrations). The rate law is thus a
very succinct and practical way of expressing the rate. You might use this,
for example, in a model of the atmosphere or in predicting the rate of an
enzyme catalysed reaction.
2. The form of the rate law can tell us something about the mechanism of
the reaction. This is a point which we will consider in more detail below.
3. Knowing the rate law enables us to separate the concentration
dependence from the underlying, fundamental effect which is the size of the
rate constant.

2.3 Effect of temperature


Rate constants, and hence reaction rates, and are often found to depend
strongly on temperature. It is therefore important to quote the temperature
at which any rate constant is determined. Most commonly the rate goes up
with temperature, but this is not always the case. For example the rate
constant for the reaction 2NO + O2  2NO2 decreases with increasing
temperature.
Experimentally, a very large number of rate constants are found, to some
level of approximation, to vary with temperature according to the
Arrhenius equation:

kT    Ea ⎞
A exp  
⎝ RT ⎠ [2]
where we have written the rate constant as k(T) to emphasize that it depends
on temperature.
R is the gas constant, which has dimensions energy temperature–1 mol–1, Ea/RT must be dimensionless
so the product RT has dimensions of energy mol–1. Ea has units of energy for us to be able to take the
exponential of it.
mol–1 and is called the activation energy; it is usually expressed in kJ mol–1.
Typical values for Ea are between 10 and 200 kJ mol–1, for example:

2N2O5(g)  4NO2(g) + O2(g) Ea = 160 kJ mol–1


EtO– + MeI  EtOMe + I– (EtOH solution) Ea = 82 kJ mol–1
Cl(g) + O3(g)  ClO(g) + O2(g) Ea = 2.1 kJ mol–1
CH(g) + CH4(g)  C2H4(g) + H(g) Ea = –1.7 kJ mol–1

A is called the A factor or the pre-exponential factor. It is often found to


be independent of temperature, or at least only weakly dependent on
temperature. Since the exponential term is dimensionless, A must have the
same dimensions and units as k.
The Arrhenius equation predicts that, for a positive activation energy, the
A
rate constant increases with temperature. The equation can be manipulated increasing Ea
into a straight line form by taking natural logarithms of both sides:
k
  Ea ⎞
k T   A exp 
⎝ RT ⎠ T
E 1⎞
ln kT  ln A  a  of the form y  mx  c The rate constant, k, plotted as
R ⎝T⎠ a function of temperature for
three values of the activation
So a plot of ln k against 1/T should be a straight line; such a plot is called an energy in the ratio 1:5:10. Note
that the temperature scale does
Arrhenius plot. The slope of the plot is –Ea/R (dimensions temperature, e.g. not start at zero.
K), and the intercept with the vertical axis, when 1/T goes to zero, is ln A.
We see that A can be identified as the rate constant at infinite ln A
temperature (or, more precisely, when RT >> Ea) i.e. the maximum possible slope = –Ea/R
rate constant. In Section 3 we will see what the significance of A and Ea are
ln k

at the molecular level.


It should be noted that many physical and chemical processes show
approximate Arrhenius behaviour, even if the processes themselves turn out the 0

to be rather complex. This is partly because experimental data is often not stoichiomet An
Arrhenius
of sufficient quality to detect minor deviations from the Arrhenius equation. ric plot
equation Exercise 5

2.4 Relation to equilibrium constants that these


Chemical equilibrium is a dynamic state, not one in which all reaction has will be the
stopped. The forward and back reactions continue just as they do when a rate laws).
system is away from equilibrium, but when equilibrium has been reached At
these forward and backward rates are equal. As a result, there is no change equilibrium
in the concentration of any of the species, even though the reactions are still
going on.
Suppose we have a reaction of the form
A+B C+D
Also suppose that rate of the forward reaction is kf[A][B] and that the rate
of the back reaction is kr[C][D] (these are assumptions, we cannot say from
1/T

these two rates are equal

kf A eq Beq  kr Ceq Deq


where the "eq" subscripts have been added to emphasize that this relation is
only true at equilibrium.
Rearranging gives
kf Ceq D eq

kr A eq B eq
The quantity of the right is the equilibrium constant, Keq, so it follows that
k
K  f
eq
kr
The equilibrium constant is therefore the ratio of the rate constants of the
forward and backward reactions.

3. Theories about reaction rates


We would like to be able to develop a theory which will enable us to
calculate – almost certainly with the aid of a computer – the rate constant of
any reaction we care to specify. To develop such a theory will be first and
foremost a test of our understanding of the fundamental processes involved.
If the theory produces predictions which are in accord with experiment,
then we may have some faith in the theory.
Then, if the theory proves to be successful and our calculations reliable,
we could go on to use the theory to predict the rates of unknown or little-
studied reactions. We might want to do this to avoid experiments, which
are not always easy or possible.
It turns out that although the theory is now quite well understood the
calculations needed to predict values of rate constants are very challenging.
With the currently available super-computers it is feasible to calculate rates
for simple gas phase reactions (e.g. F + H2  HF + H and H2 + OH  H2O
+ H). Reactions in solution present an even greater challenge as we have to
consider the role of a large number of solvent molecules.
Even though we will not be able to actually make any calculations of
rate constants ourselves, a great deal of insight into chemical reactions is
obtained by studying the model on which the theory is based and looking at
the general features we expect from such a model. This will lead to an
interpretation of why rate laws have the form they do, the factors that
influence the size of the rate constant, and how the Arrhenius equation
arises.

3.1 What happens in a reaction?


Suppose that two molecules A and B are going to react. When they are a
large distance apart we imagine that each molecule is largely independent
of the other. In A the atoms are held together in a certain geometry by
electrons which we imagine as occupying MOs.
As A and B approach one another, the electrons and nuclei from one
molecule start to interact with those from the other. We usually describe
this by saying that the MOs begin to interact and, as we saw in Reactions
and Mechanisms in Organic Chemistry course, often the strongest
interaction is that between the HOMO of one molecule and the LUMO of
another.
As a result of this interaction, bonds begin to be broken and new bonds
begin to be formed; new MOs are generated. As the rearrangement
progresses further, the new molecules (the products) begin to gain their
identity and move apart. The atoms and electrons (and MOs) can now be
identified as belonging to one molecule or another; the reaction is complete.
An example of this process is the familiar SN2 reaction:

Nu X Nu
X Nu X

HOMO/LUMO interaction products

Chemical reactions are therefore seen at a fundamental level to involve


electronic rearrangements and orbital interactions. The theory needed to
understand them is the same theory needed to study the electronic structure
and bonding in molecules i.e. quantum mechanics, in the form of MO
theory.

3.1.1 Potential energy surfaces


Using MO theory, we can calculate (at least in principle) the energy of a
molecule, or come to that any arbitrary arrangement of nuclei. All that we
need to do is to find the energies of the MOs, assign the electrons to these,
and hence compute the overall electronic energy the system.
As two molecules A and B come together and start to interact, we can
imagine computing the total energy of the system. When A and B are far
apart, the energy will simply be the sum of the energies of A and B.
However, as they approach the energy will depend on the details of how the
atoms and orbitals in A and B are interacting; eventually we will reach the
transition state whose energy we can also calculate. Finally, as the product
molecules move apart, the energy will tend to the sum of the energies of the
products.
So, we can imagine that the total energy of the system varies
continuously throughout the reaction.
The problem is that the energy will depend on the precise arrangement
of atoms, and if A and B have any complexity at all there will be an almost
mind-boggling different number of ways in which the two molecules can
approach one another and interact.
In principle we can calculate the energy of each of these arrangements of
the atoms in A and B as they interact. What we end up with is the idea of a
potential energy surface (PE surface or PES) which gives the energy (the
potential energy) as a function of the positions of all the atoms in the
system.
For all but the simplest molecules, such a surface is a function of a very
large number of distances and angles. It cannot simply be visualized as
energy as a function of the x- and y-coordinates; rather it is multi-
dimensional.
Difficult though it may be a visualize or draw, the PE surface is key to
our understanding of reactions. We imagine the atoms as "moving" on this
potential energy surface. They start out at one position which corresponds
to reactants, move along some path over the surface as they rearrange
themselves, and then end up at a position which corresponds to products.

alternative products
potential energy

reagents

products

Note that the stable molecules, the products and reactants, exist in potential
energy minima. There may be several such minima on the surface. For
example, the reaction between H and ClBr may give HCl + Br or HBr + Cl;
the alternative products will correspond to different minima on the surface.

3.1.2 The lowest energy pathway and the transition state


As the reactants and products are in potential energy minima, it will always
be the case that extra energy has to be put in order to traverse the path
between them. The amount of extra energy will depend on the details of the
potential energy surface and the path taken. Of the many possible paths
which lead from reactants to products, the one which involves the least
expenditure of energy turns out to be the one which is most favoured.
The point of highest energy on this pathway is called the transition state;
it corresponds to an arrangement of the atoms of the reacting molecules in
which there are partially made and partially broken bonds. Just exactly
where the transition state is depends on the details of the PE surface.
Why is the lowest energy pathway the one which the reaction takes?
The answer lies in understanding the way in which energy is distributed
amongst the molecules. To move from reactants to products by the
minimum energy pathway still needs a certain amount of energy; to be
precise the extra energy, E, needed is the difference between the energy of
the transition state, and the energy of the reactants.
Only a fraction of the molecules will have this extra energy, and this
fraction is, according to the Boltzman n d i stribution,
exp  E a ⎞
⎝ RT ⎠
As this fraction is an exponential function of the energy, increasing E by
just a small amount (1 or 2 kJ mol–1) reduces greatly the number of
molecules which have the required extra energy. Thus, very few molecules
are able to take pathways which are higher in energy than the lowest energy
pathway and so, for all practical purposes, we can safely assume that all the
reactants which go to products do so via this minimum energy route.

3.1.3 The transition state


The transition state is not a molecule in the conventional sense. Molecules stable molecule

exist in potential energy minima; this means then when they distort there is
a force which pushes them back to their equilibrium geometry. The PE
surface at the transition state is a minima in some directions, but along the
transition state
path that leads from reactants to products it is at a maximum. This means
that when the molecule distorts along this path there is no restoring force
and the distortion continues; the transition state then "falls apart" either to
products or back to reactants.
As it exists at a PE maximum the transition state only has a fleeting
existence, perhaps of the order of a the period of a molecular vibration,
~ 10–13 s. The recent advent of lasers capable of producing femto-second
pulses has made it possible – just – to study something as short-lived as a
transition state.
The calculations which are needed to map out the PES are time
consuming, even for the most powerful computers. A number of surfaces
have been mapped out in great detail, including those for the reaction
between H + H2, F + H2 and H2 + OH  H2O + H.
It is essential to distinguish the transition state from an intermediate. An
intermediate is a molecule like any other, existing in a PE minimum, and
detectable by conventional means; it may be highly energetic compared to
reactants and products, but it is not unstable in the sense that a transition
state is.

3.1.4 The reaction coordinate and the reaction profile


The minimum energy pathway from reactants to products is called the
reaction coordinate. Moving along this pathway involves changing the
positions of all of the atoms involved in the reaction in a complex way, so
the "coordinate" is in fact a composite of many motions.
We can plot the energy against the reaction coordinate to give a plot
which is know as the reaction profile, which you have encountered before.
potential energy
transition state

Ea

products

reactants

reaction coordinate

The transition state exists at the potential energy maxima, and the
difference in energy between this maxima and the reactants is called the
activation energy, Ea.
The fraction of molecules coming together with sufficient energy to
reach the transition state is exp(–Ea/RT). For a typical value of the
activation energy of 50 kJ mol–1, this means that only about 1 in 10 9 of the
molecules which attempt to reach the transition state actually have enough
energy to make it. We can imagine the molecules all queuing up like eager
climbers at Everest base camp (the reactants); many set off towards the
summit (the transition state), but few have the energy to make it, and so
slide back down the slope.
As we commented before, actually calculating the form of the potential
energy surface is a challenging task. However, this discussion has allowed
us to develop an understanding of the origin of the energy barrier to
reaction, and to introduce the important concepts of the transition state and
the reaction profile.

3.2 Collision theory


Although we have not said so explicitly, we have assumed that reactions
take place when molecules come together in a collision, and we have also
assumed that the energy needed to overcome the activation barrier comes
from the energy of the collision. This leads to the idea that we might be able
to understand something about the rates of reactions by analysing molecular
collisions. In its simplest form collision theory is rather straightforward to
develop and apply but, as we shall see, the values of the rate constants that
it predicts do not compare well with experiment. Nevertheless, the simple
theory does aid our understanding of the nature of reactions at a
microscopic level.

3.2.1 Gas kinetic theory


This is a simple theory about the motions and energies of the particles
(molecules and atoms) which make up gases. The theory is based entirely
on classical mechanics (i.e. Newton's Laws).
The theory treats the particles as objects whose size is much less than the
typical distance between them. These particles have kinetic energy which
causes them to move around in random directions, colliding with one
another and with the walls of the container. All collisions are assumed to
be elastic, which means that energy is conserved.
With these constraints, it can be shown that for a macroscopic sample
the energies and speeds of the particles are distributed according to the
Maxwell distribution. All speeds are possible, but very high and very low
speeds are very improbable. The mean speed, c , is given by
 8k T ⎞ 12
c B

⎝ m ⎠
where kB is Boltzmann's constant, T is the temperature and m is the mass of
the particle (in SI units, the mass must be in kg).
When considering molecular collisions it turns out that we need to know
the mean relative speed, crel , of two molecules, A and B; this makes sense
as collisions just depend on the relative motion of the two molecules. This
mean relative speed is
1

c  8Bk2

rel   
⎝  ⎠
where  is the reduced mass, given by  is the reduced mass of A and
B calculated using this formula;
 do not confuse this with the
  m
⎞ A mB reduced mass of a diatomic

molecule.
 m m 
⎝ A B ⎠
Example 4
3.2.2 Calculation of the collision rate
If we want to calculate the rate of a reaction we need to calculate the rate at
which the reactant molecules collide. However, we know that the
orientation of molecules is important for reaction, so just calculating the
number of collisions will not be sufficient. For example an S N2 reaction
requires that the nucleophile attacks from the rear; an attack from the front
is much less likely to be successful.

– –
Nu X Nu X

In its simplest form, gas kinetic theory cannot cope with this subtlety; we
simply have to assume that the molecules are structureless spheres. We
will see that this is a major defect of the theory.
It is relatively easy to compute the collision rate, ZAB, between a
molecule A, which is represented by a sphere of radius rA, and a molecule
B, represented by a sphere of radius rB. The details are given in the
appendix; here, we will simply quote the result†


You are not required to know how to prove this expression nor remember its precise
form.

ZAB
 cB r  r  rel
2

cA A B
c 1

    8k T ⎞ 2

r  B 
cA cB rA 
B
2  ⎠



where cA and cB are the concentrations of A and B, respectively, in
molecules per unit volume. In SI units the radii would be in m, the reduced
mass in kg (not mass units) and the concentrations in molecules m–3, hence
the units of ZAB will be collisions per m3 per second (m–3 s–1).
The quantity rA  rB  is an area and is often called the collision cross-
2
Typical values for  (in nm2)
are 0.24 for collisions between
H2 molecules, 0.43 for N2, 0.30
for CO, 0.21 for He, 0.91 for Cl2
section, . Using this we can re-write the collision rate as
and 0.40 for O2.  8k T⎞ 1
Z  c c  B  2

AB AB
 
Example 5 ⎝  ⎠
Exercise 6
3.2.3 Calculation of rate constants
We will now assume that a collision between A and B will result in a
reaction provided that the energy of the collision is sufficient to overcome
the energy barrier for reaction which was described in section 3.1.4.
The fraction of collisions with energy in excess of Ea, the activation
energy, is given by exp(–Ea/RT) (see section 3.1.4). So the number of
successful collisions per unit volume per unit time is ZABexp(–Ea/RT). Each
of these collisions leads to a molecule of product, so the number of moles of
product formed per unit time per unit volume is ZABexp(– Ea/RT)/L, where
L is Avogadro's number. This is the rate expressed in moles per unit volume
per unit time. So:
1
r ZAB
L exp Ea RT 
1  8k T ⎞ 12
 c   B  exp RT 
c E a
 [3]
⎝ 
AB
L
If the reaction takes place when A and B collide, as we have assumed,
we expect it to be first order in A and in B; the rate law is

rk
cA cB
2nd
L
L [4]
where we have divided the concentrations, ci, by Avogadro's number so that
they are in the more usual units of moles per unit volume. Thus the rate
will be in moles per unit volume per unit time.
Then, comparing Eq. [3] and Eq. [4] we can find an expression for the
second order rate constant as
exp E RT 
1
B
k  8k L 2

2nd      a
⎝  ⎠ [5]
3 –1 –1
k2nd will be in m mol s .
3.2.4 Comparison with the Arrhenius equation
The expression for the rate constant, Eq. [5], has a similar form to the
Arrhenius equation (Section 2.3)

 8k 2 L exp RT 
1
B
collision theory k
  E
2nd   a
⎝  ⎠
Arrhenius k  A exp RT 
Ea

Comparing these two shows that the A-factor is related to the collision rate
–to be precise it is the collision rate per unit concentration of A and B
   1
 8k T ⎞ 2
Acoll theor 2  B 
rA rB  
 ⎠ [6]

L

The activation energy in the Arrhenius expression can be identified with the
extra energy needed to reach the transition state.
Example 6
Collision theory predicts that the A-factor depends on temperature, Exercise 7
1
varying as T 2 . However, the exponential factor has a much stronger
temperature dependence, and this tends to swamp the weak temperature
dependence of the A-factor.

3.2.5 Comparison with experiment


In order to test our expression, Eq. [6], we need to find values for the radii
+
and the activation energy. The radii can be obtained from measurements of
properties of gases, such as viscosity and effusion. Using gas kinetic
theory, these measurements of bulk properties can be related back to the
size of the atoms or molecules involved.
As collision theory does not predict the value of the activation energy,
we will simply compare experimental and predicted values of the pre-
exponential factor, A.
*Diels-Alder reaction between
reaction T/K Aexp Acoll theor Aexp two cyclopentadiene
(gas phase) / dm3 mol–1 s–1 / dm3 mol–1 s–1 /A molecules.
coll theor

CH3+CH3C 300 2.4  10 10


1.1  10 11
0.22
2NOCl2NO + Cl2 470 9.4  109 5.9  1010 0.16
H2+C2H4C 800 1.24  106 7.3  1011 1.7  10–6
Diels-Alder reaction* 500 1.5  106 3.0  1011 5  10–6
K+Br2KBr+Br 600 1.0  1012 2.1  1011 4.8

The general picture is clear: with one exception collision theory


overestimates the value of the A-factor. The overestimate varies from a
factor of 5 for a rather simple reaction such as CH 3+CH3C, to a factor
of more that 106 for the Diels-Alder reaction.

3.2.6 Steric factors


It is not surprising that collision theory predicts A-factors to be too large as
in the derivation we have assumed that all collisions with sufficient energy
lead to reaction. Clearly this assumption is not valid as we know that
orientation of the reactants is important. By assuming that the molecules
are hard spheres, simple collision theory ignores orientational effects
entirely.
The fraction of sufficiently energetic collisions which lead to reaction is
called the steric factor, p. It can be given a value by comparing the
experimental and predicted A-factors
Aexperiment
p A
coll theor

Typically p is much less than 1, but there are a few reactions which have p
greater than 1. These are reactions in which it appears that the molecules
interact over greater distances than the radii found from gas kinetic theory
would indicate.

4. Experimental determination of rate laws and rate


constants
We need to determine, by experiment, the form of the rate law, and hence
the associated rate constants, for a number of reasons:
1. To extend our knowledge of the rates of chemical processes, so that
predictions can be made.
2. To compare the measured rate constants with those predicted by theory.
3. To study the mechanisms of chemical reactions.
The last point is one we will consider in more detail in Sections 5 and 6
where we will find that a knowledge of the form of the rate law can be used
to find out something about the mechanism of a complex reaction.
The basic measurements we can make are concentration as a function of
time. We then use various methods to determine the rate law from this raw
data. Section 4.1 first considers how this data can be analysed and the
Section 4.2 considers how these data are obtained.

4.1 Fitting data to rate laws


There are a very large number of possible rate laws, so there is really no
way that we can look at some data of concentration against time, and
determine the rate law directly. The way we proceed is to propose a rate
law and then see if the data can be made to fit it. If the fit is acceptable to
within the errors of the experimental data we say that the proposed rate law
is consistent with the data. If the fit is not good enough, another law will
have to be proposed and tested against the data.
Rate laws are essentially differential equations, and so need to be
integrated (solved) in order to see if the data fits the law. A few simple rate
laws can be solved "by hand", but most can only be solved numerically
using a computer program. There are many computer algorithms available
for tackling this problem.
We will investigate just the simplest rate laws which can be integrated
by hand.
4.1.1 First order rate law
For a reaction of the form A  products, a first order rate law takes the
form
r  k1st
A1
writing the rate in the differential form this becomes
dA  A [7]
 k
1st
dt
We need the minus sign as A is a reactant and so its concentration
decreases with time i.e. the slope of a graph of [A] against time will be
negative.
The variables [A] and t in Eq. [7] can be separated to give:
1
dA  k dt
A 1st

which can then be integrated


1
 dA
  k 1st dt

 lnA k1stt  const.


The constant of integration can be removed by supposing that at t = 0 [A] [A]0
[A] = [A]0, the initial concentration of A; so: Ln[A]o = const
Substituting this gives the final form
lnA  k1st t  lnA 0 [8]
t
This can also be expressed as Plot of [A] against time for a

A  A exp 1st t


0
[9] first order reaction. [A] falls
exponentially from [A]0 towards
k  zero.

4.1.1.1 Straight line graph for first order kinetics


Equation [8] is of the form y = mx + c, so a graph of ln[A] against time will ln [A]0
be a straight line, with slope –k1st and intercept ln[A]0. slope = –k1st
ln [A]

We do not need to know the absolute concentration in order to determine


the first order rate constant, just some quantity which is proportional to
concentration. The argument goes like this. Suppose that the measured t
quantity, I, is proportional to concentration: I =  [A], where  is the
constant of proportion whose value we do not know. [A] and [A]0 can be
written in terms of I: [A] = I/ and [A]0 = I0/. Equation [8] becomes
ln I    k1st t  ln I 0  
– ln   ln I  k1stt  ln   ln I0 is a very useful and
ln I  k1stt  ln I0 unique feature of
analysing first order
The unknown constant  cancels, and so a plot of ln I against t has slope data.
–k1st, just as before. The need for only a relative measure of concentration
Exercise 8
4.1.1.2 More complex first order rate laws
Equation [7] expresses the rate in terms of the reactant. What happens if
we want to express the rate in terms of the appearance of a product? For
example, if the reaction has the stoichiometric equation
AB
we can express the rate law
as
dB A [10]

k
1st
dt
Note that the derivative is positive as the concentration of B grows as the
reaction proceeds. We cannot solve this differential equation as it stands
since both [A] and [B] vary with time.
However, we can find [B] at any time by expressing it in terms of [A].
This can be done by recognizing that, due to the stoichiometry of the
reaction:
[B] = [A]0 – [A] [11]
where [A]0 is the initial concentration of A. In words, each mole of [B] that
is formed must be as a result of the loss of a mole of [A]. Equation [11]
can be rearranged to
[A] = [A]0 – [B] [12]
[A]0 and this can be substituted into Eq. [9]
B A  A0exp [9]
conc.

to give k1st t 
A
A  B  A exp k t 
0 0 1st
t

Growth of the product [B] and


B  A01  exp [13]
k1stt 
decay of the reactant [A] as a
function of time. At all times As expected, this predicts that [B] rises from zero and tends towards [A]0 at
[A] + [B] = [A]0. long times.
If the stoichiometric equation were different, for example
A  2B
the relationship between the concentrations is different as two moles of B
are produced from each mole of A consumed. Then
Exercises 9, 10 & 11
1 1
[A] = [A]0 – 2 [B] or [A] + 2 [B] = [A]0

4.1.2 Second order rate law


For a reaction of the form A  products, a second order rate law takes the
form
dA
 2nd A [14]
k 2
dt
As before, the variables [A] and t in Eq. [14] can be separated to give
1
dA  k2nddt
A 2
which can then be integrated
dA
1

 A
 k 2nd
dt
1
  k
2nd t  const.
A
The constant of integration can be found by supposing that at t = 0
[A] = [A]0; so
1
  const.
A
Which gives the final 0
form
1 1
k t
A 2nd [15]

4.1.2.1 Graphs for second order rate laws


A plot of [A] against time shows a hyperbolic decrease starting from [A] 0 slope = k2nd
and falling asymptotically towards zero at very long times. Equation [15] is

1/[A]
of the form y = mx + c, so a graph of 1/[A] against time will be a straight
line, with slope k2nd and intercept 1/[A]0. 1/[A]0
In contrast to the case of a first order process, we need to know absolute 0 t
concentrations in order to find a value for the rate constant.

4.1.2.2 More complex second order rate laws


Consider a reaction with the stoichiometry
A+BC
and rate law

dA
 2nd A
k
dt
As [A] and [B] are related by the stoichiometry of the reaction it is possible
to solve this differential equation, but it is not that easy. Any more complex
rate laws really are impossible to solve "by hand".
If at time zero the concentrations of A and B are equal, then at all
subsequent times they will remain equal because of the 1:1 stoichiometry of
the reaction: [A] = [B]. Then the rate law becomes
dA 4.1.3 Is ethod
 ol
k at
dt io
which is one we have already solved. n
m
2nd

A 2
As rate laws become more complex, it rapidly becomes difficult to integrate
them by hand. A way out of this is to use the isolation method as way of
simplifying the rate laws. Suppose we have a reaction
A + B  products
and we suspect that the rate law is
dB  A  2  B
 k
3rd
dt
The idea is to make measurements with one reagent in great excess,
typically more than 50 times the concentration of the other. If A is in
excess, we can say that its concentration will not change very much during
the course of the reaction, i.e. as [B] drops from its initial value to zero.
The rate law becomes
dB  A  2  B
 k
3rd 0
dt
where [A]0 is the initial excess concentration of [A]. In this situation,
k3rd [ A]02 is constant and can be written as keff; the rate law becomes

dB
 ef B where keff  k3rd A 0
2

k

f
dt
The rate law is now looks like that of a first order reaction – we usually call
this pseudo first order because it only appears to be so; underneath lurks
greater complexity. Likewise keff is called a pseudo first order rate constant
as it is not really a rate constant (it depends on concentration).
Having found keff we can find k3rd, which is what we really want to know,
from keff = k3rd[A]02. If several measurement have been made at different
excess concentrations of A, the value of k3rd can be found from a plot of keff
against [A]02.
Exercise 12
Equally well we could have chosen to put B in excess, in which case the
rate law would become
dA
 3rd A  2 B 
k 0
dt
which, as k 3rd is effectively constant, can be written
B 0
dA
 eff,2 A2 where
eff,2   B 0
k k k3rd
dt
Now keff,2 is a pseudo second order rate constant.
There are problems with the isolation method. Firstly, putting one
reagent in such large excess may make the rate inconveniently fast.
Secondly, such conditions may alter the mechanism of a complex reaction.
Nevertheless, the isolation method is a popular approach to analysing
complex rate laws. In practice one would put all suspected reagents but one
in excess so that the order and the pseudo-rate constant for that species
could be determined. The same procedure would then be repeated for all
the other reagents.

4.1.4 Differential method


If the rate law is of the form
r
[16]
k A 
n

then an appealing method of finding the value of the order, n, is to plot the
log of the rate against the log of [A]
ln r  ln k  n lnA
Such a graph will have slope n. This method, called the differential
method, is convenient as, in contrast to the method of Sections 4.1.1and
4.1.2 we do not have to make any assumptions about the value of n.
The drawback is that rather than plotting a function of concentration we
have to plot rates, and rates are much harder to measure than
concentrations. The rate is the slope of a graph of concentration against
time, and as such a graph is usually curved taking an accurate slope is not at
all easy.
More complex laws can, of course, be manipulated into the form of Eq.
[16] by using the isolation method described in Section 4.1.3

4.1.5 Half lives


The half life of a reaction is defined as the time it takes for the
concentration of a specified reagent to fall to half of its initial value. For a
first order reaction it turns out that the half life is independent of the initial
concentration:
To prove this we start with the integrated rate law, Eq. [8]
lnA  k1st  lnA0
[8]
At time t = 0 the concentration is [A]0; after a half life, t = t 12 , the [A] [A]0

concentration is [A]0/2. Substituting these values into Eq. [8] we find [A]0/2
lnA0 2  k1stt21  lnA0

t 12
 k1s lnA
1  2
0 t1/2 t
t
0 lnA0 
t1 
1 A0
ln
2
k1s A 0 2
t
t 1  1 ln
2

k1s 2
t

We have shown that the half life is independent of the initial concentration. This means that for a
first order reaction the time taken for the concentration to halve from a
particular value is constant throughout the reaction.
In reality, measuring the half life is not a very accurate way of finding
the rate constant; we would do much better to plot the data as suggested in
Section 4.1.1.1. However, it may be useful when appraising kinetic data
roughly to keep in mind this property of the half life and its relation to the
rate constant.
We can use a similar method to find the half life of a second order
reaction, starting from Eq. [15]. The result is:

Exercise 13

t1  1 1
2

k2nd A 0
In contrast to the first order case, this half life does depend on the initial
concentration.
The half life is a useful way of specifying the extent of a reaction. If a
reaction has been taking place for several half lives we can say that a
significant amount of reaction has taken place and the process is well on the
way to completion or equilibrium. On the other hand, a process which has
been observed for less than a half life since it started has not taken place for
long enough to be characterised very well.

4.1.6 Comparing rate constants


As the half life of a first order reaction is independent of concentration it is
easy simply to look at the rate constant and get an immediate feel for how
fast it is. For example, the reaction
2N2O5  4NO2 + O2
has a first order rate constant of about 2000 s–1 at 400 K; the half life is
therefore 0.3 ms, which makes the reaction quite fast.
On the other hand the second order reaction
2NOBr  2NO + Br2 (gas phase)
has a rate constant at 280 K of 0.8 mol-1 dm3 s–1. Is this "faster" or "slower"
than the first order reaction above? To answer this we need to know the
concentration of the reactants. Suppose that the pressure of NOBr is 0.1
atmospheres, corresponding to a concentration of 4.4  10–3 mol dm–3; using
the above expression we find that the half life is 280 s – so, under these
conditions, the reaction is much slower than the decomposition of N2O5.
The key point to understand here is that we cannot compare directly the
rates of reactions of different orders by comparing their rate constants;
concentration has to be taken into account.

4.2 Obtaining kinetic data


The raw data needed for the analysis described in Section 4.1 is
concentration as a function of time. In this Section we will investigate how
we go about measuring such data. This is the raw material of kinetic
studies.
Kinetic measurements represent quite a challenge to the experimentalist.
Firstly, reactions proceed on a vast range of different timescales – varying
from the almost geological to sub nano-second. We need all sorts of
different strategies for making measurements over this range. Secondly,
many reactions involve complex mixtures, perhaps with the species in
vastly different concentrations; we want to be able to measure the
concentrations of all these species individually. Thirdly, we want to be able
to do all this without interfering with the reaction mixture – this points to
the use of physical methods of measuring concentration, which are non-
invasive. Finally, it would be nice to be able to automate taking
concentration readings.
First we will look at methods of measuring concentration, and then
consider the special difficulties of making measurements on fast reactions.

4.2.1 Light absorbance


As you have seen in earlier lectures, molecules absorb light at characteristic
frequencies. These frequencies are associated with transitions between
energy levels, for example, those associated with vibrations or electrons.
For kinetic studies, the commonest kinds of transitions to use are those
in the visible or ultra-violet (UV) part of the spectrum (1000 – 200 nm).
These transitions come about when the absorbed photon causes the
molecule to move to an excited electronic energy level.
All molecules absorb in the UV to some extent, but it takes the presence
of special groups to cause strong absorptions in the visible region. Here are
some examples of species that have been monitored by UV/vis absorption
in kinetic experiments:
COO–
SCN
H2O
CH3 N
OH2
Br2 Fe
H2O

MeO O
NO2 OH2 OH2

400 nm 495 nm
490 nm 216 nm 278 nm
The extent to which light is absorbed is related directly to the
concentration of the absorbing species by the Beer-Lambert Law.
II l
0 exp cl
[17] I0 I
where I0 is the intensity of the light entering the medium, I is the intensity
of light exiting the medium, l is the path length through which the light
passes, c is the concentration and  is the extinction coefficient. The value
of the extinction coefficient depends on the species absorbing the light and
the wavelength. The path length is related to the physical dimensions of the
container holding the sample. Typically, a glass or quartz "cuvette" of
known dimensions (a 1 cm path length is common) is used for solutions.
Equation [17] can be rearranged to give the concentration
ln I  ln I0  cl
I
ln 0  cl
I
The quantity ln(I0/I) is called the absorbance, A, and it is common for
spectrophotometers to read this out directly. Absorbance is directly
proportional to concentration.
A  cl
A typical value for the extinction coefficient for a transition which absorbs "strongly" is 10,000 mol–1 dm3 cm–1; a "weak" absorption
might have an extinction
coefficient three orders of To turn measured absorbances into absolute concentrations, we need to
magnitude lower.
know l. Usually this is done by measuring the absorbance of a series of
Example 7
solutions of known concentration and then simply plotting A against c; the
slope is l. If we are studying first order processes then only a relative
measure of concentration is needed, so it is sufficient just to know the
absorbance (Section 4.1.1.1).
An example of the use of light absorption to follow a reaction is the
bromination of acetone (under acid conditions)
O O
Br2 Br–
Br

Of all the species involved, only Br2 absorbs significantly at 495 nm and so
measuring absorbance at this wavelength enables us to measure its
concentration.
Absorbance measurements are very convenient; they are non-invasive,
rapid and can easily be automated. As different species absorb at different
wavelengths it is possible to study different molecules in a reaction
mixture. However, UV/vis absorptions tend to be rather broad (especially in
solution), so it is possible that more than one species will absorb at a given
wavelength. The same principle applies to IR measurements.
Proton NMR can also be used to monitor concentrations, as under the
right conditions the peak height is proportional to concentration. It has the
advantage that individual chemical species can easily be followed, but
NMR is not very sensitive so as an analysis method it is not very fast.

4.2.2 Conductivity measurements


Ions in solution conduct electricity and the resistance of such a solution
depends on the concentration of the ions and their identity. In fact the
conductance, which is the reciprocal of resistance, is directly proportional
to the concentration of the ions.
If a reaction involves a change in the number of charged species it is
possible to monitor the progress in the reaction by measuring the change in
conductance. An example is the following reaction, in which neutral species
react to give charged ones; the reaction can therefore be monitored by the
increase in conductance.

N N
Ph Br N+ Br–
N

Ph
The contribution that an ion
makes to the conductance The following reaction (in aqueous solution) involves no change in the
depends on the type of ion. In number of ions, but as the identity of the ions change there is still a
water, H3O+ and OH– ions are
very mobile and are thus good
measurable change in conductance
conductors.
O O
+ – EtOH
OH +
OEt –
O
Exercise 14
The relationship between conductance and concentration can be rather
involved, so it is really only convenient to use this method for relative
measurements of concentration i.e. appropriate for first order kinetics,
Section 4.1.1.1 It is relatively easy to measure conductance using some
kind of bridge or more modern electronic device; the measurement is non-
invasive and rapid.

4.2.3 Gas pressure


The pressure, p, that a gas exerts is related to the volume, V, the number of
moles n, and the temperature, T, by the ideal gas equation
pV  nRT
This can be rearranged to give

n p
V  RT
the fraction n/V is a measure of concentration; in SI it would be in moles m–
3
. The relationship tells us that at constant temperature pressure is directly
proportional to concentration.
In a reaction we will generally have a mixture of gases, and so we can
talk about the partial pressure, pi, of each gas i. The partial pressure of a
gas is the pressure that the gas would exert if it occupied the volume on its
own. Dalton's Law states that the total pressure, pT, is the sum of the partial
pressures:
PT = P1 + P2 + P3 +…..
Each partial pressure follows an ideal gas law

ni pi
V  RT
where ni is the number of moles of gas i. Therefore the partial pressure of
species i is a measure of its concentration.
In a reaction involving gases the only measurable pressure is the total
pressure exerted by the system. However, to make kinetic measurements we
need to know the concentration, that is the partial pressure, of the individual
species. We need, therefore, to find some way of relating these two. An
example of how this is done is considered here.
Consider the reaction between NO and O2 in the gas phase
2NO + O2  2NO2
Overall three moles of gas go to two, so as the reaction proceeds there is a
measurable decrease in the total pressure. The total pressure is the sum of
the partial pressures
pT  pNO  pNO2  pO 2
There are three unknown quantities on the right, and only one
measurable quantity, the total pressure; to solve this we are going to have to
introduce some extra restrictions.
Suppose that at time zero there is no NO2 present, and that the NO and
O2 are mixed in a 2:1 ratio, just as the stoichiometry of the reaction. If the
initial partial pressure of O2 is pO ,02
, then the initial partial pressure of NO
is 2 pO ,0 , and so the initial total pressure, p0, is 3 pO ,0 . This is the situation
2 2

set out on line 1 of the Table

pO2 pNO pNO pT


2

1 pO ,0 2 pO ,0 0 3pO 2,0  p0
2 2

2 pO2,0  p O2 2 pO2 ,0  p O2 2pO2 3 pO2 ,0  p O2

3
3 1
p0  pO2 2
3
p0  pO 2 2pO2 p0  pO2
4 pT  32 p0

Suppose after time t the partial pressure of O2 has fallen by pO2 ; due to
the stoichiometry of the reaction the partial pressure of NO will fall by
2pO2 and the partial pressure of NO2 will increase by 2pO2 . This is the
situation set out on line 2 of the Table.
Our aim now is to find pO2 in terms of measurable quantities. There are
two steps. In going from line 2 to line 3 in the Table we have used the fact
that p0  3 pO 2
,0 to substitute pO ,0 . From line 3 we then see that
for 2
pT  p0  pO 2
from which it follows that pO2  p0  pT . This value of
pO2 is used in going from line 3 to 4.
So, finally, on line 4 we have an expression for pO2 in terms of
measurable quantities: pO 2  pT  32 p0 .
Exercises 15 & 16
There are many devices capable of measuring pressure, from simple
manometers (columns of mercury) to various electronic devices, some of
which read absolute pressures. If concentration is measured in pressure
units, it is common to express rate constants in pressure units too. So, for
example, a second order rate constant might be expressed in mmHg–1 s–1.

4.2.4 Electrochemistry
As we have seen, the EMF (voltage) generated by an electrochemical cell
depends on the concentration of the species in the cell. By constructing a
cell with an appropriate electrode it is possible to measure the concentration
of a wide variety of ions in solution. Such electrodes can give both absolute
and relative measurements of concentration; they are non-invasive,
sensitive to particular ions (selective) and have a rapid response. They are
widely used when studying reactions of ions in solution.
An example of the use of electrochemistry is in studying the oxidation of
formic acid by bromine in aqueous solution:
Br2 + HCOOH  2Br– + 2H+ + CO2
The calomel electrode is there
to complete the cell (we need Two electrodes are dipped into the solution: a calomel reference electrode
two electrodes); it plays no role and a platinum electrode. At the platinum electrode the redox process is
in the reaction.
Br2 + 2e– 2Br–
and so the voltage produced by the Pt electrode depends on the
concentration of the bromine and the bromide. It can be shown that the
voltage produced by the cell, E, is

EE0
RT  –
ln Br 2 
2F Br2 
where E0 is the standard EMF (voltage) produced by the cell (a known
quantity), R is the gas constant and F is the Faraday, 96485 C mol–1. If the
reaction is run with an excess of Br – then the voltage produced by the cell
depends solely on the concentration of Br2; the progress of the reaction can
therefore be monitored.

4.2.5 Chemical methods


Concentration can be measured by "classical" methods of chemical
analysis, such as titrations. Such methods are rather slow, and so it is
usually necessary to extract some of the reaction mixture (an "aliquot") and
then stop (quench) the reaction so that no further reaction takes place during
the analysis.
Examples of such quenching procedures might be sudden cooling or
chemical removal of one of the reagents. For example, the alkaline
hydrolysis of an ester can be followed by taking samples of the reaction
mixture, quenching them in a know excess of acid to stop the reaction and
then titrating the remaining acid against standard alkali. In such a way the
concentration of alkali in the reaction mixture can be determined.
A somewhat more sophisticated chemical analysis uses a gas
chromatogram (GC), possibly in conjunction with a mass spectrometer. A
gas chromatogram separates the components of a mixture by using an inert
carrier gas to sweep them down a column (typically about a metre long)
which is packed with some sort of porous solid material. Different
components travel at different speeds down the tube and so are detected at
different times after injecting the sample; their concentrations can thus be
measured separately.
The detector may be a simple one which detects a change in the thermal
conductivity of the carrier gas resulting from the presence of other
molecules, or it may be a mass spectrometer which can not only detect the
presence of molecules but also identify them. A mass spectrometer can also
be used on its own to measure and identify species directly – that is without
the use of a GC.
An example of using this method of analysis is in studying the
photolysis of acetaldehyde (photolysis means a reaction, usually
destructive, brought about by irradiation with light). There are a number of
possible products resulting from what turns out to be rather a complex
reaction
CH3CHO + h  CH4 + CO + C2H6 {not a balanced equation}
The rate of production of all three possible products can be monitored by
passing small amounts of the reaction mixture into a GC at known times
and measuring the heights of the three separate peaks which appear.
Another example of analysis using a GC is the pyrolysis of ethane
(pyrolysis means destruction by heating)
C2H6  CH2=CH2 + H2 + other products {not a balanced equation}
As a further example, a mass spectrometer was used to monitor the rates
of the following two reactions which are both important in the atmospheric
oxidation of ethane
C2H5 + O2  C2H5O2
C2H5 + O2  C2H4 + HO2
The two alternative products, C2H5O2 and C2H4 have different masses are
so can be detected separately by the mass spectrometer.

4.2.6 Studying fast reactions


Whatever method we choose for monitoring the concentration of the
reacting species we also have to consider how the reagents are going to be
mixed and the reaction initiated. This is a serious problem for fast
reactions. It is no good if it takes 1 second to mix the reagents if the
reaction is over in 100 ms! Further, our method of measuring concentration
has to be fast enough to make measurements over the time scale of the
reaction.
A great deal of effort has been put into developing methods for
measuring such fast reactions; we will consider some of these here.

Continuous flow
reagent from
reservoir
mixing zone observe

waste
d
reagent from
reservoir

The idea here is to flow the reagents together (they can be gases or
solutions), and after mixing let them continue to flow down the reaction
tube. Different distances down the tube correspond to different times after
mixing and initiating the reaction. As the reagents are constantly
replenished, we can take our time in making observations at any distance
(and hence time). Essentially the method is a way of achieving rapid
mixing and then being able to monitor the reaction at leisure, rather than in
"real time".
Mixing can be complete in under 1 ms, so reactions with half-lives on
the ms timescale can studied. A typical flow velocity is 10 m s–1, so a
distance of 1 cm corresponds to 1 ms of reaction.
Usually the reaction is monitor by spectrophotometry, which is a very
rapid technique. The main disadvantage of this method is that you needs
lots of reagents, particularly if you are trying to observe a fast reaction
where the flow rate needs to be high.
An example of using this method in solutions is studying the
complexation reaction between Fe2+ and thiocyanate, SCN–, in aqueous solution. The resulting
complex absorbs strongly in the visible part of the spectrum and is thus
easily monitored by spectrophotometry.
The method is perhaps used more in the gas phase; in this case the flow
is created by pumping on the "waste" end of the tube and supplying gases at
the other.
For example, the reactions between oxygen atoms and other species can
be studied by generating O atoms in an electric discharge and then flowing
them down a tube. Reagents can be introduced into the tube and the rate of
their reaction with O atoms studied.
The problem is how to measure the O atom concentration in the flow
tube. One way to do this is as follows. A small amount of NO is introduced The green glow is called the air
into the flow tube. The NO reacts very quickly with O to produce afterglow and was first
observed in the early part of the
electronically excited NO2, denoted NO2*; this then rapidly emits a photon 20th century.
(in the green part of the spectrum) which is detected using a
photomultiplier. Finally, the NO2 the reacts with another O atom
regenerating the NO.
O + NO  NO2*
NO2*  NO2 + h (detected)
O + NO2  NO + O2
As all of these reactions are fast and the NO is regenerated, the intensity of Exercise 17
the green emission is directly proportional to the O atom concentration.

Stopped flow

trigger initiates observation


reagent when plunger hits stop
observe

plunger moves back until hits stop


reagent

This apparatus is similar in conception to the continuous flow method, in


that the reagents are flowed together to ensure rapid mixing. However, the
difference is that the reacting mixture pushes a plunger back until it hits a
stop; then the flow is halted and measurement, at a fixed point, is initiated
(again, usually using spectrophotometry). Observations are made at a point
at which mixing is complete. The stopped flow method is only used for
solutions.
In contrast to continuous flow, stopped-flow requires measurements in
real time. However, it has the advantage of not requiring large amounts of
reagents (as little as 1 cm3 may be sufficient). Generally the time resolution
is a little faster then for continuous flow.
Two examples, both taken from studies on enzymes:
(1) The enzyme myeloperoxidase (found in leukocytes) catalyses the
formation of HOCl from H2O2 and Cl–; the HOCl is though to play a part in
the antimicrobial activity of these cells. The reaction was studied using a
stopped flow system not only to find out about the overall rate of the
reaction but also to monitor two transient intermediates which are formed
along the reaction pathway. These absorb at different wavelengths, so their
build-up and decay could be monitored separately.
(2) The enzyme lactate dehydogenase catalyses the oxidation of lactate to
pyruvate; the oxidizing agent is to the cofactor NAD+
OH
O
+
CO2– NAD NADH
CO2–
lactate pyruvate
Fluorescence is when light of
one wavelength is emitted by a
The enzyme binds NADH rather strongly, and in elucidating the details of
molecule when it is irradiated the mechanism it was necessary to find out the rate constant for the
by light of another (usually
shorter) wavelength. The
dissociation of the NADH-enzyme complex. This was studied using the
irradiation causes electrons to stopped flow technique, and the reaction was monitored by a change in
move to higher energy levels
and these then emit light
fluorescence from the NADH when it binds to the enzyme.
(fluoresce) when they drop
back down to lower levels. Flash photolysis
The technique of flash
photolysis was pioneered by
The idea of this method is to use a short, intense flash of light (now often
Norrish and Porter, working in from a laser) to generate some reactive species, typically free radicals.
this Department in the 1950s;
in 1967 they received a Nobel
Then, the reaction of these species with other reagents present in the system
Prize for their work. is monitored, usually spectrophotometrically.
This is a useful method for studying the fast reactions of radicals; there
is no mixing problem as the reagents are generated in situ. However,
measurements do have to be made in "real time", so rapid detection is
needed.
Two examples:
If two radicals come together
by definition they have
(1) The recombination of two CH3 radicals, to give ethane, was studied by
sufficient energy to fall apart generating the methyl radicals by photolysis of acetone (a pulse a laser light
again. How then, do radicals
recombine? Somehow they
of duration tens of ns was used). The concentration of the CH 3 radicals was
have to lose the excess energy followed spectrophotometrically.
before they fall apart. Kinetic
studies, on such things as (2) The reaction, in solution, of a carbocation was followed after it has been
methyl radical combination, are generated using a laser pulse
used to unravel how this
excess energy is disposed of. laser pulse
Ph Ph + nucleophiles, Nu– Ph
20 ns, 248 nm
Cl Nu
Ph Ph Ph
Second order rate constants of up to 109 mol–1 dm3 s–1 have been measured.

5. Analysing the kinetics of complex reactions


We have already come across the idea that many reactions take place in a
series of steps, often involving the formation of intermediates. This set of
steps is called the mechanism. For example, the alkaline hydrolysis of
some esters is thought to proceed by the following mechanism
O HO O– O

OH EtO–
OEt OEt OH
Another example is the set of steps believed to be responsible for the
destruction of ozone in the upper atmosphere
Step 1: Cl + O3  ClO + O2
Step 2: ClO + O  Cl + O2
The Cl used up in step 1 is regenerated in step 2, and so can go on to
destroy more ozone.
Each of the reactions which comprise the mechanism is called an
elementary step; it is elementary in the sense that we believe that it takes
place in a single reactive encounter between the species involved. The
elementary steps are thus the basic building blocks of a complex reaction
and cannot be broken down further. A mechanism can be defined as a
proposed set of elementary steps which account for the overall features of
the reaction.
One of the aims of studying chemical kinetics is to be able to determine
the mechanisms of complex reactions, and this Section is concerned with
just how kinetics can help in such a study. However, a study of chemical
kinetics alone is unlikely to be able to determine a mechanism, especially a
complex one. We need to use other information, such as the direct
detection of intermediates (for example, by spectroscopy), trapping and
labelling experiments, and above all our chemical intuition. However,
chemical kinetics does play an important role in such studies.
The way we proceed with a complex reaction is first to propose a
mechanism and then proceed to analyse its chemical kinetics. We can then
compare this with experiment and see if the two agree. If they do, the
proposed mechanism can be said to be consistent with the kinetic evidence;
if they do not agree, we have to go back and modify the proposed
mechanism. There is no way that we can "determine" the mechanism just
from kinetic data, all we can do is propose a mechanism and see if it fits the
data.
The details of the mechanism we propose will depend on other
information available to us and our chemical intuition.

5.1 Rate laws for elementary reactions


We can always write down the rate law for an elementary reaction just by
inspecting the stoichiometric equation. This is because we believe the
reaction to take place in a single encounter between the species involved.
For example, for the reaction
Cl + O2
 ClO + O
k

we can write down straight away the rate of loss of reagents or gain of
products:
dCl dO 2 
 kClO  k ClO 
2
dt 2


dt
dO dClO
 kClO  kClO 
2 2
dt

dt
Example 8
Exercise 18 As before, the rate has to be negative for a reagent, as it is lost in the
reaction, but it is positive for a product as it is gained.

5.2 Rate laws for complex reactions


Once we specify a reaction mechanism we can easily write down the rate
laws for all the steps. As an example we will consider the reaction of HBr
with an alkene (in solution). The overall reaction is
Br
HBr
R R
but it is not believed that this takes place in a single encounter between the
reactant molecules. Rather, it is thought to take place by two sequential
elementary steps, each of which is a single reactive encounter. The
mechanism is thought to be:

H Br Br
k1 k2 Br

R R
R
A C+ P
The rate of formation of the product, P, is
dP
 k C Br   
2
dt
This depends on the concentration of the intermediate carbocation, C +.
Now this species is created in step 1, and destroyed in step 2, so its overall
rate of change depends on the rates of both steps
d C  
     
 

k1 A HBr k2 C Br
dt
The first term, which is positive, is the rate of step 1 in which C + is formed;
the second term, which is negative, is the rate of step 2 in which C+ is
destroyed. Using this approach we can write down expressions for the rate
of change of all species involved in a given mechanism.
Even for this simple reaction the mathematics has become a bit involved.
To find out exactly how [P] depends on time – which we would need to do
to compare the predictions of the mechanism with experiment – we need to
integrate these differential equations, just as we did in Section 4.1. This is
not easy as the variables from one equation appear in the other; that is the
equations are coupled. In fact this pair can be solved relatively easily, but if
we add a few more steps to the mechanism we rapidly end up in a situation
where the solution becomes impossible by hand.
It is possible to solve such coupled differential equations numerically
using a computer algorithm. Indeed, a huge amount of effort has gone into
this approach as it is essentially the only way of handling complex kinetic
systems such as the atmosphere, combustion and so on. For our purposes,
however, we need to find some simpler approaches which, although they
may not be universally applicable, will enable us to get a handle on the
kinetics of complex reaction schemes without resorting to computers.
All of these methods remove some of the complex time dependence by
making assumptions about the relative rates of the processes involved.

5.3 Sequential reactions


Suppose that we have two sequential reactions which for simplicity we will 1.0

concentration
assume are both first order C
0.8

A k1  B 1
B k2  C 2 0.6

We can imagine two extreme (limiting) cases:


0.4
(1) Rate constant of process [1] is very much greater than that of process
[2]: B
0.2
k1 >> k2
A
In this case A is converted rapidly into B, leading to an accumulation of B 0.0
0 1 2 3 4 5
which is more slowly converted into C (see the graph opposite); the t/s
production of C mirrors the fall in B. The rate at which the product C is Concentration against time for
formed depends on the rate of process [2] – it is said that step [2] is the rate two sequential reactions. Here
k1 = 4 s–1 and k2 = .5 s–1.
determining step.
(2) Rate constant of process [2] is very much greater than that of process
[1]:
k2 >> k1.
In this case A is converted slowly into B, and then B reacts to form C
almost as soon as B is formed (see graph below); the production of C
mirrors the fall in A. The rate at which the product C is formed depends on
the rate of process [1] – step [1] is now the rate determining step. In this
limit there is very little B present at any time as the moment it is produced
it reacts. We will refer again to this Section 5.5.
If either of these limiting cases applies then the kinetics is greatly 1.0
concentration

simplified, as the overall rate of production of the product depends only on C


the step with the smallest rate constant. The same is true for more complex 0.8

schemes of sequential reactions


0.6

A k1  B k2  C k3  D k4  E


If the step from B to C has the smallest rate constant, then the rate of 0.4

formation of the product E depends only on k2.


If the reactions are not first order things become more complex as the 0.2
A
B
relative rates of the steps depend not only on the rate constants but also on
0.0
the concentration of the species involved. Consider the following 0 1 2 3 4 5
t/s
sequential reactions
Concentration against time for
A  B   C   D
k1 k2 two sequential reactions. Here
k1 = .5 s–1 and k2 = 4 s–1
The first step is second order, and so its rate relative to the second step can
be altered by varying the concentrations of A or B.
Things get more complex if the reactions are not simply sequential; to Exercises 19 and 20

handle these we need some further tools introduced in the next two
sections.
5.4 The pre-equilibrium hypothesis
In many mechanisms the species directly involved in the rate limiting step
are in equilibrium with the reagents. A typical example is the acid
catalysed hydrolysis of esters:

H OH2 H 2O
O k1 OH

OEt
OEt
E k–1
+
EH

HO k2 OH
H2O
H2O
OEt RDS OEt
+
EH
The carbonyl group of the ester (E) is protonated on the oxygen and it is
this species (EH+) which is subject to nucleophilic attack by water in the
rate determining step. As the attack by water is slow, it is quite likely that
EH+ will lose its proton and return to ester before it is attacked by H 2O – the
first reaction is reversible.
In the pre-equilibrium hypothesis we assume that (1) E, H3O+ and EH+
are always in equilibrium which implies that (2) the reaction of EH + with
H2O is slow enough that the equilibrium is not perturbed. The latter reaction
is therefore rate determining.
The reaction scheme is
E  H O  k1  EH  
H O3 2 1
EH   H O k 1  E 
2 3 -1
H O
EH   H O k 2  rest of 2
2
reaction
For the equilibrium hypothesis to apply the rate of process [–1] must be
much faster that the rate of process [2], so that equilibrium is established
between E, H3O+ and EH+. If this is so, then the rates of [1] and [–1] are
equal, as this is the definition of equilibrium (see Section 2.4):
Ea,–1
Ea,2 at equilibrium : rate of [1] = rate of [–1]
  
k E H O+  k EH+ H  [18]
O
EH++H2O 1 3 –1 2
Ea,1
Alternatively, we could write an equilibrium constant for the equilibrium

E+H3O + products
EH EH  H2O
O 3
Energy profile for the case EH H
 2
[19]
Keq 
O
where pre-equilibrium applies.
The intermediate has a lower 3
energy barrier (Ea,–1) to
EH O 
returning to reactants than it By comparing Eq. [18] and Eq. [19] we find, as expected from Section 2.4,
does to go on to products
(Ea,2). Note that there is no that
particular requirement on the k
size of Ea,1. K  1
eq
k 1
The rate of formation of products is assumed to be controlled by the rate of
process [2] i.e. this is the rate limiting step.

rate  k2 EH  H 2O  [20]
+
An expression for [EH ] can be found from Eq. [19]
 E  H 3 O 
   
EH  Keq
H 2 O
Substituting this into Eq. [20] we find 
E H O  
rate k 2 H 2O K eq 3

H 2O
 k K E H O   or k k1
EH O [21]
2
  3
2 eq 3 k1
The final form of the rate equation is easy to handle as it only involves the
reacting species and no intermediates; the differential equation is simple
enough that it can be integrated in the ways described in Section 4.1.
Let us assume that in an experiment it is found that the rate law is of the
form
rate  k
exp EH3 O 
Comparison of this with the theoretical rate law of Eq. [21] enables us to can assume that the
identify kexp as k1k2/k–1. In other words the "rate constant" kexp is not really a concentration of B is
rate constant but a composite of rate constants of three elementary constant.
processes. For different esters we could then interpret changes in kexp in Mathematically this
terms of changes in either Keq (i.e. the basicity of the carbonyl), or k2 (i.e. condition is expressed
the reactivity of the protonated carbonyl), or both. as
This pre-equilibrium approach is often successful for intermediates
which involve simple protonation or de-protonation, as such processes are
generally fast compared to breaking or making bonds to heavier atoms.

5.5 The steady-state hypothesis


When we were discussing sequential reactions in Section 5.3
A k1  B k 2  C
we discovered that if the rate constant of process [2] is much greater that
that of process [1] the intermediate B reacts the moment it is produced. We
noted that in this situation the concentration of B will always be low.
It is also the case that once the reaction has got under way, the
concentration of B will change very little during the reaction (see graph
opposite). During the majority of the time that the reaction is proceeding we
1.0

concentration
A
0.8

0.6 C

0.4

0.2

B ( 5)
Example 9 0.0
0 1 2 3 4 5
t/s

dB Concentration against time for


dt SS two sequential reactions (see
0 Section 5.3). Here k1 = .2 s–1
and k2 = 10 s–1; the
This approximation – that the concentration of an intermediate is not concentration of B is shown
changing – is called the steady state approximation. When the reaction is using a five-fold scale
expansion.
proceeding with the intermediate or intermediates at constant concentration
the reaction is said to be reacting "in the steady state".
Use of the steady state approximation greatly simplifies solving the
differential equations as it removes the time dependence. We will show
how this comes about for the simple scheme
A k1
1
BB
2
k2 
C
The rate of change of B has two terms
dB
 k A k B
1 2
dt
as before, the first term is positive as process [1] leads to the production of
B whereas the second is negative as process [2] destroys B. If we assume
that B is in the steady state then
dB  k A k B  0
1 2 SS
dt SS
where the subscript SS is to remind us that the concentration of the
intermediate is calculated assuming the steady state. This equation can be
rearranged to give an expression for [B]SS

BSS

k1A
k
2

In the steady state we can the used this expression to calculate the rate of
formation of products
dC  k B
2 SS
dt SS
k1A

k k
2 2

 k1 A
The result is hardly a surprise. For B to be in the steady state it must be
react the moment it is formed. This is situation (2) described in Section 5.3,
in which the first step is rate determining and the reaction rate just depends
on the rate of process [1]. Our final expression for the rate of formation of
products says just this – the rate depends only on kl.
Exercises 21 & 22
5.5.1 Validity of the steady-state approximation
The steady state approximation must be applied with care – it is not
appropriate for all of the species in a reacting mixture. It would not, for
example, be appropriate to apply it to reactants or products as these decay
away and accumulate during the reaction. Generally, it is appropriate for
intermediates which are high in energy, often called "reactive
intermediates". Such species are relatively slow to form, but once formed
their high
reactivity
means
that they
are
consumed
at once by
further
reactions.
Rate
laws
derived
using the
steady-
state
approxim
ation will
only be
valid once
the
reaction
has
reached
the steady
state; they
will not
be valid
in
the initial phase (sometimes called induction phase) of the reaction or when
the reaction is close to completion.
You can get some idea of how valid the steady-state and pre-equilibrium
hypotheses are by using the applet which you will find at
http://www.ch.cam.ac.uk/magnus/kinetic.html You can compare
the exact and approximate solutions for different combinations of rate
constants (check "k–1 = 0" to simulate the simple A  B  C scheme).

5.5.2 Example: acid catalysed hydrolysis of esters


We will return to the acid catalysed hydrolysis of esters, using the scheme
analysed in Section 5.4:
E  H O  k1  EH   1
H O3 2

EH   H O k1  E  -1
2 3
H O
EH   H O k2  rest of  2
2
reaction
This time, however, we will assume that EH+ is in the steady state – which
requires that it reacts via steps [–1] or [2] as soon as it is formed. The
expression for the rate of change of the concentration of [EH+] is
d EH+    +
 +
   +
   


 
k1 E H3O
k–1 EH H2O k2 EH H2O
dt
the first term is positive, as EH+ is formed in step [1] but the second and
third terms are negative as EH+ is destroyed in steps [–1] and [2]. Applying
the steady-state approximation and rearranging gives us an expression of
[EH+]SS

d EH+     +
 +
   +
   



 H2O
dt SS k1 E H3O k–1 EH SS k2 EH SS H2O 0
  +
 
k1E H3 O+ 
S
EH S
k –1 H 2 O  k 2 H 2 O
The rate of formation of products is just k2[EH+][H2O]
rate  k H O EH+
2 2
 
SS

 k 2 H 2O
k E H O+
1
 3

k –1 H 2O  k 2 H 2O


k1k 2E H3 O+ 
k–1  k2 [22]
This is our steady-state expression for the rate of the reaction. It is different
from the one derived using the pre-equilibrium hypothesis, Eq. [21], but not
that different.
In order for EH+ to be in the steady state it must react the moment that it
is formed either by returning to reactants (step [–1]) or by going on to
products (step [2]). We can imagine three different situations according to
the relative size of the rate constants of these two processes.
(1) rate of process [2] is much faster than rate of process [–1].
Ea,–1
In words this means that once the protonated ester is formed it is more
Ea,2 likely to go on to products than be deprotonated. Mathematically this is
Ea,1
EH++H2O expressed as
 
k 2 EH+ H2 O k–1 EH+ H2 O  
+
E+H3O cancelling the common terms on left and right this becomes k >> k . The
products 2 –1
Energy profile for case (1) denominator in Eq. [22] can then be approximated to k2 so the rate law
where the rate constant for step
2 is greater than that for step becomes
 
–1 i.e. the activation energy for
2 is less than that for –1. k1k 2 E  H 3O 
rate
k2


 k1E H3 O 
We can see that this makes sense. We have assumed that step [2] is much
faster than [–1] and so in effect the reaction is just two sequential steps,
with the first one being the slower. It is not surprising, therefore, that the
overall rate law just depends on k1, as [1] is the RDS.
(2) rate of process [–1] is much faster than rate of process [2].
Ea,–1
In words this means that the protonated ester is more likely to be
Ea,2
deprotonated and return to reactants rather than going on to products.
Ea,1 EH++H2O Mathematically this is expressed as
k–1
EH H O EH H O
+ 2 2 + 2

k
+
E+H3O
products cancelling the common terms on left and right this becomes k–1 >> k2. The
Energy profile for case (2) denominator in Eq. [22] can then be approximated to k–1 so the rate law
where the rate constant for step
2 is less than that for step –1 becomes
 
i.e. the activation energy for 2
is greater than that for –1. k k E  H O 
rate 1 2 k 3
1


 K eqk 2E H 3 O 
where we have used Keq = k1/k–1, as in Section 2.4. This expression is
identical to the one derived assuming that steps [1] and [–1] formed a pre-
equilibrium (Section 5.4). Assuming, as we have done here, that step [–1]
is faster than step [2] does of course ensure that such an equilibrium is set
up and that step [2] is a slow "bleed off" of EH +. Step [2] is rate
determining, as is predicted by the rate law.
(3) rate of processes [–1] and [2] comparable
No simplifications to the rate law, Eq. [22], are appropriate.
We have seen, therefore, that even for this simple reaction scheme three
different rate laws are possible, depending on the relative rates of the
different processes involved. All three rate laws can be determined from
the single steady state expression.
5.5.3 Example: decomposition of N2O5
Our second example is the thermal decomposition of N 2O5 in the gas phase.
The overall reaction is
2N 2 O 5  4NO 2  O 2
and experimentally the rate law is found to be
dN 2O5 
dt  k obs N 2O5 
At first, it was thought that this was an example of elementary first order
reaction, but careful investigations showed that several steps and
intermediates are involved. Using the steady state hypothesis we shall be
able to show that this complex mechanism gives a simple rate law.
The proposed mechanism involves four elementary steps:
N O k1  NO  NO
2 5 2 3 [1]
NO  NO k1  N
2 3 25 [–1]
O
[2]
NO 2  NO3 k2  NO 2 O  2
[3]
NO NO 2 N5 O  2 3NO k3

Our approach will be first to identify those species which are


intermediates and then to put their concentrations in the steady state. The
obvious targets are NO and NO3; all the other species are reagents or
products. The rate of change of each can be written down just by inspecting
the reaction scheme.
To avoid clutter, we have not
As NO3 as is generated in step [1] and lost in steps [–1], [2], it follows that added the subscript "SS" to the
dNO3  steady-state concs. of NO and
NO  k  NNOO   k1 2 
NO
5 –1 2   3k2
NO 2
NO3.
dt
3

 k1 N2 O5  –1  k2 2 NO3 [23]

k NO 
 0 in the steady state
Likewise, NO is generated in step [2] and lost in step [3]:
dNO NO  k NON O  [24]
k
NO 2 2 3 3 25
dt
 0 in the steady state
The rate of change of the reactant N2O5 involves three terms: N2O5 is
consumed in steps [1] and [3], and regenerated in step [–1]
dN 2 O5 
 k N O   k NO NO   k NON O 
2 5
dt
1 2 5 1 2 3 3 [25]
This expression involves the concentration of intermediates and we will
attempt to eliminate these with the aid of the steady state expressions for
NO3 and NO, Eqs. [23] and [24].
The expression for [NO3] in the steady state is easily found from Eq.
[23]:
0  k1 N 2O 5   k –1  k 2 NO 2
[26]
NO3 
k1N2O 5
hence NO3  

k –1  k 2 NO 2 
An expression for [NO] in the steady state can be found from Eq. [24]
0  k 2 NO 2 NO 3   k3 NON 2O 5 
k NO 2 NO 3 
hence NO  2
k3 N2O5
This expression includes [NO3], an intermediate whose concentration we
are trying to eliminate, so we will substitute for it from Eq. [26]
k 2 NO 2 NO3 
NO  k3N2O5
k 2 NO 2  k 1 N 2 O 5 

k3N 2 O 5  k –1  k 2 NO 2 
k1k2

k3 k –1  k 2 
Finally, substituting this expression for [NO] and Eq. [26] for [NO3] into
Eq. [25] we find an expression for the overall rate
dN 2 O5 
 k N O  k NO NO  k NON O 
1 2 5 1 2 3 3 2 5
dt
 k N O
 NO  k1 N 2 O 5 
k 
2
  k1k2 N O 
k3 k–1  k 2
3
k–1k NO 2
1 2 5 1 2 5
k2 
 N O
k  k k

 k k †
–1 1 12
 
2 5

1
k–1   k–1  k2 
 2k k † k2
 N O 
25
⎝ k–1  k2 
1 2
[27]
 
Our analysis has produced a rate law of the correct form, and we can
identify kobs as the quantity in brackets on the last line. The surprise is that
such a relatively complex mechanism can produce such a simple rate law.
An interesting feature of the rate law is that the rate constant for process
[3], k3, does not appear. We can rationalise this by noting that the only fate
of the NO generated in step [2] is to be consumed in step [3]. So, we can
argue that the rate constant for step [3] is immaterial as there is no
competing process for NO.
Example 10
Experimentally, it has been determined that the rate constant of process
[–1] is much faster than of process [2] i.e. k–1 >> k2. Under these conditions
the denominator in the final expression for the rate law, Eq. [27], can be
approximated to k–1, resulting in overall rate law
dN 2 O5  k1
 2 k N O 
2 25
dt k–1
We recognise k1/k–1 as an equilibrium constant.
Exercise 23

5.5.4 Enzyme kinetics: the Michaelis-Menten equation the aid of catalysts call
In biological systems almost all chemical processes are brought about with enzymes. These
catalysts are remarkable in the specificity they show – that is an individual
enzyme catalyses a particular reaction between particular molecules – and Enzymes are proteins, which
are long-chain polymers made
for their efficiency. Enzymes can achieve chemical transformations under up of different amino acids.
mild conditions and with a speed which chemists can only dream of O
achieving at the bench.
H 2N
OH
Enzymes work by binding to the reagent molecules and then providing a R
reaction pathway which is lower in energy than the pathway in the absence There are about 20 different
naturally occurring amino acids,
of the enzyme. At the end of the reaction the products are released and the and a typical enzyme may have
enzyme can be used to catalyse further reaction. The molecules which bind between 50 and several
hundred individual amino acids
to the enzyme, that is the reagents, are called substrates. in the chain.
For many enzymes it is observed that, for a fixed quantity of enzyme, as O R1 O
H
N
the amount of substrate increases the rate of the catalysed reaction first N N
increases linearly and then levels off to a limiting maximum value (see H
O R
2 H
graph opposite). This behaviour is called "saturation kinetics". The chain folds round on itself
in a very specific way due to
We can reproduce this form using a simple kinetic scheme, named after the formation of hydrogen
its inventors Michaelis and Menten, by an analysis using the steady state bonds and other non-covalent
interactions. The catalytic
approximation. The overall scheme is power is very much associated
with the shape of the protein.
E+S ES  E + P
where E is the enzyme, S the substrate, ES the enzyme-substrate complex
linear
and P is the products saturation

rate or velocity
The three elementary steps are:
E  S k1 
 1
ES ES k 1
-1
 E  S ES
2
kcat  P 
E
Steps [1] and [–1] involve the substrate binding to and being released,
[substrate]
unchanged, from the enzyme. Step [2] leads to products and the release of
Typical behaviour of the
the enzyme, ready for it to act again. velocity or rate of an enzyme
We assume that ES is in the steady state and, as it is formed in step [1] catalysed reaction as a function
of substrate concentration.
and lost in steps [–1] and [2], it follows that
dES
 k ES ES ES
k k
1 –1 cat
dt [28]
 0 in steady state
We know the initial concentration of enzyme, [E]0, but once the reaction is
under way the free enzyme concentration will be less than this as some has
become ES. At any time the total amount of free and bound enzyme is
equal to [E]0 so
[E]0 = [ES] + [E]
hence [E] = [E]0 – [ES] [29]
We can substitute this expression for [E] into Eq. [31], leaving [ES] as the
only unknown:
k1 E S  k –1 ES  kcat ES  0
k1 E 0  ESS  k –1 ES  kcat ES  0
k1 E 0 S
hence ES
k 1S k –1  kcat [30]

Biochemists tend to use the term "velocity" rather than rate, so in this
section we will follow them and use V for velocity rather than r for rate.
The velocity of formation of product is given by
V  kcat ES
Substituting for [ES] from Eq. [30] gives
kcat k1 E0 S
V
k1 S  k 1  k cat
It is usual to rearrange this by dividing top and bottom by k1 to give
kcat [E]0 S
V k k ⎞
S  1 k cat 
⎝ 1 ⎠
kcat [E]0

 S [31]
S  KM

Equation [31] is the Michaelis-Menten equation; KM, the Michaelis
constant, is defined as
k1  kcat
KM 
k1
KM has dimensions of concentration.
The dependence of V on [S] is fairly complex, so it is usual to measure V
in the initial rate, which means for times short enough that [S] has not
changed much from its initial value. We can then take [S] as its initial
value when analysing data using Eq. [31] to find kcat and KM.

Interpreting the Michaelis-Menten equation


First, we will examine the final expression, Eq. [31], to see if it has the
correct behaviour. At low substrate concentrations, such that [S] << KM,
the denominator of Eq. [31] can be approximated to KM and so
kcat[E]0 S
Vlow [S] 
KM
The velocity is now first order in both [S] and [E]0, and this corresponds to
the linear region of the plot given above. In this limit the ratio kcat KM is
the second order rate constant for the reaction, and its value determines the
rate when there is more free enzyme than enzyme bound to substrate. In
this limit the rate is limited by the substrate binding to the enzyme. There is
plenty of enzyme "waiting" for substrate to process.
Typical values for kcat K M are in the range 101 to 108 mol–1 dm3 s–1; the
upper limit is close to the diffusion controlled limit, which is the fastest
possible rate[S]
At high substrate concentrations, that>>a reaction can proceed of
KM, the denominator in solution.
Eq. [31]
can be approximated to [S] and so
Vmax  kcat[E]0
In this limit the velocity is independent of [S] and is at its maximum,
usually called Vmax or Vsat; this is the saturation behaviour described above.
This limit is interpreted as being when virtually all the enzyme is present as
enzyme-substrate complex. The system is then working at its maximum
capacity – adding more substrate has no effect on the rate. The rate limiting
step is ES going to products.
Equation [31] can be rewritten by replacing kcat[E]0 by Vmax to give Vmax

Vmax S
V V
S  KM
Using this form we can find an interpretation of the value of KM. Consider Vmax
the substrate concentration, , which results in a velocity of half the 2

S 
2
1

maximum:
V Vmax S1
max
2
2 S1 2 KM KM [S]

hence S12  KM KM is equal to the substrate


concentration which gives a
velocity of half its maximum
So, in words, KM is the concentration of substrate which gives a velocity of value.
half the maximum possible.
A large value of KM implies that there is not a strong affinity between the
enzyme and the substrate, and so a large concentration of substrate is
needed to achieve half the maximum rate. A small KM implies that there is
a strong affinity and so little substrate is needed to reach of velocity of
Vmax/2.
Typical values for KM are in the range 10–6 to 10–2 mol dm–3. Note that
KM is not an equilibrium constant. Typical values of kcat are in the range 10–1
to 105 s–1.

Graphical analysis
The Michaelis-Menten equation, Eq. [31], can be rearranged into a straight
line plot by inverting both sides
k [E]0 S
V  cat
S  KM
1
hence  S KM
V k [E] S
cat 0
1 KM 1
 
kcat[E]0 kcat[E]0 S
Hence a plot of 1/V against 1/[S] should be a straight line with slope KM/
(kcat[E]0) and intercept 1/(kcat[E]0). Provided that [E]0 is known, we can
determine KM and kcat.
1/V As kcat[E]0 = Vmax the equation of the straight line can be written
1 1 KM 1
slope =  
1/Vmax
KM /Vmax V Vmax Vmax S
Now the slope is KM/Vmax and the intercept is 1/Vmax. The intercept with the
1/[S]
x-axis, when 1/V = 0 is –1/KM.
–1/KM
In practice, this kind of "double reciprocal plot" is not very suitable as
A Lineweaver-Burke plot, used the position of the line, and hence the slope and intercept, is overly
to determine kcat and KM for
Michaelis-Menten kinetics. influenced by the values at low [S] (high 1/[S]). A computer fitting direct
to the Michaelis-Menten equation would be a better option.
Example 11
Exercise 24 Discussion
The Michaelis-Menten scheme is just the simplest one which explains
saturation kinetics. Many enzymes show a more complex behaviour, for
example if the catalysis involves several steps. In addition, as the reaction
is assumed to proceed in the steady-state, the experiment does not give
individual values for k1 and k–1; to find these we would need to perform
experiments outside the steady state region.

5.6 Arrhenius parameters for complex reactions


In Section 2.3 we noted that many rate constants are found experimentally
to obey the Arrhenius equation
  Ea ⎞
k
A exp 
⎝ RT ⎠
and, as was seen in Section 3, the same dependence is expected for
elementary reactions.
For reactions involving several steps we have seen that apparently
simple kinetics can arise. For example, in section 5.5.2 we analysed the
following scheme for the hydrolysis of an ester, E,
E  H O  k1  EH   1
H O3 2

EH   H O k 1  E 
2 3 -1
H O
EH   H O k2  rest of 2
2
reaction
and found the rate law to be

rate  k1k2 E H3 O


+
 
k–1  k2
If we assume that the rate of process [–1] is much faster than rate of process
[2], this rate law simplifies to

rate 
k1k2 E H O+
 [32]
k–1
 kobs
EH O+ 
If we assume that each elementary rate constant obeys an Arrhenius law,
with its own value of A and Ea:

k1  A1 exp RT k –1  A–1 exp Ea,– RT k2  A2 exp RT 


Ea,1 1 Ea,2
then the temperature dependence of kobs can be found by substituting these
into Eq. [32]:

kobs  A1 exp Ea,1 RT  A2 exp Ea,2 RT 


A1 A2 A–1 exp Ea,–1 RT 
 exp E  E   RT 
E
a,1 a,2 a,–1
A–1
The expression looks like an Arrhenius law with pre-exponential factor
A1A2/A–1 and activation energy (Ea,1 + Ea,2 – Ea,–1). In other words the 1–1 2

apparent activation energy is a composite of the activation energies of the Ea,appEH++H2O


three elementary steps.
From the reaction profile opposite, it is clear that the apparent activation
E+H3O+
energy, Ea,app, is the energy gap between the reactants and the highest products
energy transition state on the reaction pathway. Further, as step [–1] is the Reaction profile for the case
discussed in the text. To avoid
reverse of step [1], (Ea,1 – Ea,–1) is simply the energy change, E, between E clutter, the arrow representing
+ H3O+ and EH+ + H2O. So we could write Ea,app = (E + Ea,2). the activation energies of the
three reactions have simply
In the case that process [2] is much faster than process [–1] a similar been labelled with the reaction
numbers.
analysis shows that the Ea,app = Ea,1; again, the apparent activation energy is
the energy gap between the reactants and the highest energy transition state
on the reaction pathway.

6. Chain reactions
In the multistep reactions we have looked at so far the reagents basically
undergo a series of reactions, one leading on from the other. In chain
reactions the elementary steps are arranged in rather a different way. This
is best illustrated by an example. Consider the two reactions (which are
involved in the overall reaction H2 + Br2  2HBr in the gas phase)
Br  H 2  HBr 1
 H H  Br2  2
HBr  Br
Suppose we start with a bromine atom (we will discover later where this
might come from). It can react with H2 in step [1] generating product HBr
and a hydrogen atom. This hydrogen atom can then react with Br2 in step
[2] generating more HBr and a bromine atom. The bromine atom can then
react again in step [1], and the resulting H in step [2] generates two more
HBr molecules. This can go on and on, over and over again. Thus, as the
two steps taken together do not destroy Br, one Br atom can result in the
formation of many HBr molecules.
These two steps are said to form a chain in that the output of the second
step in the input to the first. The Br and H atoms are called chain carriers
as they are the species which carry forward the reaction. Reactions [1] and
[2] are called chain propagation steps as they are the steps which carry on
the chain and lead to the formation of product.
These chain reactions have the remarkable property than only a small
number of chain carriers are needed to produce very many product
molecules. Indeed, if steps [1] and [2] were the only ones involved, a single
bromine atom could result in the consumption of infinite amounts of H2 and
Br2!
Many reactions involve chains like this; they are particularly common
for reactions involving reactive species such as free radicals. The chemistry
of the atmosphere involves many such chain reactions, and we will see
examples of this below. Combustion in flames and polymerization, such as
those used to form plastics, are also examples of chain reactions.

6.1 Terminology for chain reactions


We have already seen what the terms chain carrier and propagation mean.
The reaction which generates the chain carrier is called the initiation step.
In the case of Br2, a collision with some unspecified molecule, M, is usually
sufficient to generate Br atoms
M could be another molecule of
Br2, or H2 or an inert "buffer Br2  M  2Br  M
gas" deliberately added to the
mixture to dilute the reactants. Another frequently encountered initiation step is photodissociation, where
the absorption of sufficiently energetic photons (usually in the UV or
visible parts of the spectrum) leads to bond breaking
Cl2 h  2Cl
Processes which destroy chain carriers are called termination steps. E.g.
Br  Br  M  Br2  M
A third body, M, is needed to carry off the excess energy that the newly
formed Br2 molecule has; if this energy is not dissipated, the molecule will
fall apart immediately. M might be the walls of the container.
Another kind of step is an inhibition step; this is one which consumes
product and so leads to a reduction in the overall rate of the reaction. In the
H2 + Br2 reaction the following is thought to be an important inhibition step
HBr  H  H 2  Br
This step does not destroy chain carriers, but does destroy the product, HBr.
Some reactions have chain branching steps; these are steps in which one
chain carrier reacts in such a way as to give rise to two (or more) carriers.
For H2 + O2 reaction the following branching reaction is thought to occur.
H  O2 OH  O
OH and O are both chain carriers. The presence of chain branching
reactions can lead to an ever increasing rate of reaction – an explosion.
Example 12
6.2 H2 + Br2  2HBr (gas phase)
This reaction is perhaps the best understood chain reaction. It was studied
by Bodenstein in the early part of the 20th century and, remarkably, he
found by experiment that the rate law was
dHBr k H Br 2 3

dt  Bra  2 k HBr
2

2 b
This complex rate law gives us some clues to the mechanism. Firstly, it has
to be a fairly complex one to produce this kind of rate law, and, by
experience, the fractional power is indicative that radicals (atoms) are
involved. Secondly, the presence of the term [HBr] on the bottom of the
fraction means that HBr inhibits the reaction i.e. the reaction goes slower as
the product builds up; we will need to include a step or steps in our
proposed mechanism which accounts for this.
When the reaction has not been going on for long, the amount of HBr is
small and the term in [HBr] in the denominator can be ignored. Then the
rate law reduces to
dHBr k Br [33]
H
a 2
1 2 2

dt initial

We will start out proposing a mechanism which gives this rate law.

6.2.1 Reaction in the initial phase – low [HBr]


The heart of the reaction scheme will be the two propagation steps:
Br  H 2 
k
2
 HBr  H 2
H  Br k3  HBr  Br 3
2

We need an initiation step, which we will choose to be


Br2
 M k1  2Br 1
M
To get a steady reaction we must have at least one termination step
otherwise the number of chain carriers would go on increasing and
increasing, and as a consequence the rate would go on increasing. We
choose the termination step to be
Br  Br  M k 4  Br [4]
2
M
Why have we chosen these steps out of the many reactions we could have
chosen? The simplest reason to give is that these are the steps which will
give the rate law we want! This is not such an absurd circular argument as
it sounds. What it means is that although other reactions may be taking
place, such as
Br  H  M  HBr
 M H 2  M  2H 
M
these are not taking place at a rate which makes them competitive with the
set of reactions [1]–[4] that we have chosen.
To analyse our reaction scheme we will assume that the reactive
intermediates, H and Br, are in the steady state. We start by writing down
the rate of change of these two species, and setting them to zero in the
steady state:
dBr  2k M Br  k Br H   k H Br   2k M Br  2  0
2 2 2 3 2 4
dt SS 1 [34]
dH  k  k HBr   0
Br2H
2 3 2
dt SS [35]
The rate of formation of HBr is
simply
dHBr  k HBr 
 k BrH
2 2 3 2
dt [36]
We need, therefore, expressions for [H] and [Br] in terms of the reacting
species H2 and Br2. The expression for [H] is obtained by rearranging Eq.
[35]
k 2 Br H 2   k3 H Br2   0

hence
k 2 Br H 2
H 
[37]
k3 Br2
However this expression includes a term [Br], which we need to find
another expression for. Ultimately, this will come from Eq. [34]; but there
is a trick which greatly simplifies this substitution. Notice that the quantity
sandwiched between the lines in Eq. [34] is zero as it is precisely the
expression (times –1) which is equal to zero in Eq. [35].
So, straight from Eq. [34] we can write
 2k1 M Br2   2k4 M Br   0
2

2k M Br2 
hence Br   1
2

2k4 M

Br k1
Br2
k4
This expression for [Br] can be substituted into Eq. [37]
k2 H 2 k 1
H Br 
3
k Br
2
k 4 2
k2 k3 k1 H 2 

k 4  Br2 2
1

Then this expression for [H], along with the one for [Br], can be substituted
into Eq. [36] to give the final rate equation from which the concentration of
the all the intermediates has been eliminated
dHBr
 k BrH  k HBr 
2 2 3 2
dt
k1
k Br H  k 2 H 2
k


Br  Br 
 k3  1
2 2 2 k Br  k 2  2
k
 2k 4 ⎝ 3 2 4 
k 1 k 4 H Br 1 2
2 2 2

Comparing this to Eq. [33] we see that it is of the correct form, with
k1 k4
ka  2k2
[38]

As before, we see that the experimental rate constant is in fact a composite


of several rate constants of elementary reactions.
6.2.2 Data on the elementary reactions
The following data on various elementary reactions associated with the
H2+Br2 reaction will be needed in the subsequent discussion.

Reaction rate constant Ea E


at 500K / kJ mol –1
/ kJ mol–1
1 M + Br2  Br + Br + M 3.8  10–8 193
2 Br + H2  HBr + H 960 82 70
3 H + Br2  HBr + Br 9.6  1010 4 –173
4 M + Br + Br  Br2 + M 4.2  10–13 *
5 H + HBr Br + H2 7.2  109 12 –70
6 Br + HBr  H + Br2 173
H2  H + H 436
HBr  H + Br 366
All rate constants are bimolecular and given in dm3 mol–1 s–1 unless
otherwise noted.
*
third order rate constant in dm6 mol–2 s–1.
Exercise 25
6.2.3 Effect of an inhibition step
The experimental rate law for the H 2 + Br2 reaction shows that the reaction
is inhibited by product HBr. Our rate law derived in Section 6.2.1 does not
predict this as we included no inhibition steps. In this Section we will
extend the reaction scheme to include such steps and show that with this
modification the experimental rate law can be predicted.
There are two possible candidates for inhibition steps which consume
products:
H  HBr k5  H 5
 Br 2

Br  HBr k6  Br  6
2
H
It turns out that inclusion of step [5] gives a rate law in agreement with
experiment, but inclusion of step [6] does not. How can we understand
this? What the observation is telling us is not that "reaction [6] does not
take place and reaction [5] does" but that reaction [5] competes effectively
with other reactions of H whereas reaction [6] does not compete effectively
with other reactions of Br.
The data given in the table above show that whereas reaction [5] is
exothermic and has a small activation energy, reaction [6] is substantially
endothermic, and so necessarily has an activation energy at least equal to
this endothermicity. The difference in E values for reactions [5] and [6]
can be attributed to the fact that in [5] a much stronger bond (H–H) is being
formed than in [6] (Br–Br).
Following through the same kind of analysis as above (see Exercise 26), Exercise 26
but with the inclusion of step [5] gives the rate law
 k ⎞ 1
 2   23
dHBr 2k
 2  k H 2 Br2
1

dt  ⎝ 4 ⎠ k ⎞
Br   5  HBr
2  k3 ⎠

Comparison with the experimental rate law allows us to identify ka as
before and kb as k5/k3.

6.2.4 Chain length


As was commented on above, one chain carrying Br atom can lead to the
production of many HBr molecules as the carrier is cycled through the
chain propagation steps. In fact a Br atom will go on producing HBr until
the atom is removed in a termination step, such as step [4].
This leads to the concept of chain length, l, which is the average number
of times that the closed cycle of steps which produce products is repeated
per chain carrier. In this case, steps [2] and [3] (the propagation steps) form
the closed cycle which generates product HBr and regenerates the chain
carrier Br. The number of chain carriers depends on the rate of initiation.
The chain length can be written in terms of rates:

rate of overall reaction


l rate of initiation
which is equivalent to the definition above. In the limit of low [HBr], the
rate of the overall reaction is
dHBr
 2k k1 k4 H Br  1 2
2 2 2
dt
and the rate of initiation is the rate of step [1]
rate of init.  2k1 M Br2 
So the chain length is
k1
2k H Br  1 2
2 2 2
k4
l
2k1 M Br2 
k
 k k2 H 2 
1 2
Br  1
M 
2

4 1 2
Exercises 27 – 29
Under typical conditions ([M] = 1 atm, [Br 2] = [H2] = 0.1 atm, 500 K) the
chain length is of the order of 10 13, showing that one Br atom produces a
very large number of HBr molecules before it is terminated.

6.2.5 Comments
The H2 + Br2 reaction is probably the best understood of all chain reactions.
The same principles used to set up and analyse this reaction can be used for
other chain reactions; examples are given in the problems.
6.3 H2 + O2 – a branched chain reaction
At some temperatures and pressures the reaction between hydrogen and
oxygen proceeds smoothly, reaching a steady state as does the H2 + Br2
reaction, but at other temperatures and pressures the reaction mixture
explodes. The reasons for this different behaviour lie in the shifting
balance in the rates of different elementary steps.
The overall reaction scheme is rather complex, so we will not look at it
in detail, but concentrate on those reactions which are responsible for
controlling whether or not there is an explosion.
The initiation step involves the formation of two OH radicals
H2 + O2  2OH [1]
and the propagation step
is
H2 + OH  H + H2O [2]
The H atoms produced in this step are involved in branching reactions,
where two radicals are produced from one
H + O2  OH + O [3]
H2 + O  OH + H [4]
Unchecked, these reactions, together with [2], lead to an increase in the
number of H and OH radicals. The increase in number of chain carriers
leads to an explosion.
At low pressures (approx. 10–3 atm.), however, it is found that the
reaction is steady. It is thought that this comes about because at these low
pressures the branching reactions are slow, and so other reactions which
destroy H atoms (termination reactions) are competitive. The explosive
growth of chain carriers is thus controlled. At these low pressures, the
termination reactions mainly involve the radicals diffusing to the walls of
the vessel and recombining there.
As the pressure is raised, the rates of reactions [3] and [4] increase
(remember that they depend on [O2] and [H2]) whilst the rate of termination
by diffusion to the walls remains largely unaffected. Eventually an
explosion results at the point where the branching reactions overwhelm the
termination reactions.
As the pressure is raised still further, another reaction becomes more and
more important: this is the termolecular reaction leading to destruction of H
H + O2 + M  HO2 + M [5]
where M is any gas. The rate of this reaction goes at [O 2][M], so it
responds quadratically to pressure and becomes more and more important
as the pressure rises (contrast steps [3] and [4] whose rates vary linearly
with pressure of H2 or O2). The HO2 radical is not very reactive and tends
to diffuse to the walls where it is lost. Overall, then, at higher pressures
step [5] effectively removes H atoms, thereby reducing the rate of chain
branching and hence inhibiting the explosion; the result is once more a
steady reaction.
If the pressure is raised still further the reaction once more becomes
explosive. What is happening is that a series of three reactions is
reconverting the unreactive HO2 radicals into chain carrying OH radicals
HO2 + H2  H + H2O2
2 HO2  H2O2 + O2
H2O2  2OH
The full story is rather complex!

7. Appendix: Calculation of the collision rate


Our aim is to calculate the rate of collisions between molecule A, which is
represented by a sphere of radius rA, and molecule B, represented by a
sphere of radius rB. It is easiest first to calculate the number of collisions
with a single B molecule, which we imagine to be stationary. A collision
will take place if the trajectory of molecule A takes its centre within (rA +
rB) of the centre of B

A B

hit hit miss

rA+rB

Molecule A hits B if the trajectory takes the centre of A within (rA+rB) of the centre of B. Three
trajectories are shown above; the position of A at is closest approach to B is show in a dotted outline.
At the bottom of the diagram is shown the geometry of the most distant approach which still results in
a collision.
Now imagine what happens in 1 second. On average, all of the A
molecules in a tube of radius (rA + rB) and length c will hit the stationary B
molecule placed at one end of the tube
A

rA+rB

This argument needs a little refinement. Firstly, the speed c we need to


use in this calculation is the mean relative speed of A and B; this is needed
because we assumed that B was static. The relative speed, crel , is given by
 8k T ⎞ 12
crel   B
 
⎝  ⎠
where  is the reduced mass, given by
 m m ⎞
   m A Bm 
⎝ A B⎠

where mA and mB are the masses of A and B.


Secondly, we do not literally mean that all the particles in this tube will
hit the B molecule at the end. Many will hit other molecules before they
reach B and will be deflected into new trajectories; other molecules will
move into the tube as a result of other collisions. What we really mean is
that on average this is the number of molecules that will hit B.
With these provisos the calculation can proceed. The volume of the tube
is
cross - sectional area  length  r  
2
rel
rA B c
so if there are cA molecules of A per unit volume, the number of molecules
in the tube is
volume concentration 
r 2  cA
rA
rel
c B

All of these molecules hit the stationary B molecule, so this is the number
of collisions per second. When we take into account that there is not one B
molecule but cB per unit volume, the total number of collisions between A
and B molecules, ZAB, is

ZAB
 cB r  r  rel
2

cA A B c 1

    8k T ⎞ 2
r  B 
cA cB rA 
B
2  ⎠



3

You might also like