You are on page 1of 21

Section 16a

RATES OF REACTION AND REACTION MECHANISM


Chemical kinetics (kinetics): The study of the rates of chemical reactions and the steps by
which they occur.
Rate: Rapidity of change, a property involving time.
Some chemical reactions are very fast and appear instantaneous to the eye on mixing the
reactants, while others are very slow and the products only form over minutes, hours or even
longer.
Reaction rate: The change in concentration of a reactant or product divided by the time it
takes for the change to occur. Common units are mol L1 s1.

In this example the numerical value of the rate depends on which reactant or product
concentration is being measured. To overcome this it is possible to express the rate of
reaction defined by its chemical equation.

Most chemical reactions take place in a series of elementary reactions (steps).


Elementary reaction: A one step reaction. There are three classes of elementary reaction.
Unimolecular reaction: An elementary reaction involving only one particle
Example:

Bimolecular reaction: An elementary reaction involving only two particles (the most
common type).
The particles may be the same or different
Examples:

Termolecular reaction: An elementary reaction involving only three particles.


(Common in gas phase reactions where a third body is required to disperse the energy
released from bond formation when two species combine to give one
Example:
The molecule of chlorine initially present gains energy released by the formation of the
new Cl-Cl bond.)
Molecularity: The number of particles (molecules, ions or radicals) taking part in an
elementary reaction;
1 for unimolecular, 2 for bimolecular and 3 for termolecular.
Reaction mechanism: The sequential elementary steps of a reaction.
A+B
X
X+C
P+Q
The overall reaction is A + B + C
P+Q
The species X is called a reaction intermediate.

Reaction intermediate: A species produced in one step of a chemical reaction but which is
consumed in a later step. It is not a product of the overall reaction, and hence does not
appear in the overall equation.
Rate-determining step: The slowest elementary step which effectively determines the
overall rate of a reaction.
Rate equation (law): A mathematical equation relating the instantaneous rate of the reaction
to the concentrations of reactants, products, or any other permanent species (such
as catalysts) taking part in the reaction at that instance
Example: for the reaction
H2O2(aq) + 2H+(aq) + 2I(aq)
I2(aq) + 2H2O at pH of 7
the rate law is found by experiment to be: v = k[H2O2][I]
Rate coefficient (constant): Constant of proportionality in a rate law.
Example: k in the above rate law.
Reaction order:
(a) With respect to a single species - the power to which the concentration of a single species
is raised in a rate law.
Example: the above reaction is first-order in H2O2 and first-order in I.
(b) Overall - the sum of the powers of the species in the rate law.
Example: the above reaction is second-order overall . k is a second-order rate constant.
For a reaction between two species to occur they must collide, and most reactions take place
in a series of bimolecular collisions, i.e. by a series of bimolecular elementary steps. If the
step involves two different species, A and B, the rate of collision, and therefore the rate of
reaction, is proportional to the concentrations of A and B. Thus for the elementary bimolecular
reaction
A+B
C+D
rate of reaction = k[A][B] where k is called a bimolecular rate constant. A bimolecular rate
constant is clearly a 2nd order rate constant. For elementary reactions the order of the
reactions are the same as the molecularity. As the units for rate of reaction are mol L 1 s1 it
follows from the rate laws that the units of 1st order, 2nd order and 3rd order rate constants
are respectively s1, L mol1s1 and L2 mol2 s1 .

As bonds are often broken in elementary reactions not all collisions between reactant species
lead to reaction, only those which collide with sufficient energy to overcome the activation
energybarrier.
Activation energy: Symbol Ea. The minimum energy needed for colliding species to react, or
cross the barrier. The magnitude of k depends on the height of the barrier, i.e. on the

activation energy. The greater the Ea, the smaller the k, and the slower the reaction. The
activation energy, Ea, is determined from an Arrhenius plot.
Arrhenius plot: A plot of ln k or ln rate against 1/T where T is the absolute temperature
(temperature in kelvin). The slope of an Arrhenius plot equals -Ea/R where R is the universal
gas constant. This follows from the Arrhenius equation.
Arrhenius equation: k = Aexp(Ea/RT)
pre-exponential factor.

where A is called the Arrhenius

The kinetic energy of molecules, and hence the fraction of collisions with energy in excess
of Eaincreases with increasing temperature. Therefore the rate of a reaction, and the
magnitude of k, increase with temperature. The greater the Ea, the more sensitive the rate
and rate constant to changes in temperature.
Unlike equilibrium constant expressions, the rate law for a reaction cannot be derived from
inspection of the overall chemical equation but must be determined experimentally. Only for
elementary reactions (i.e. one step reactions) can the rate law be deduced directly from the
equation (as explained above).
Information on reaction mechanisms can be obtained from experimentally obtained rate laws.
While a rate law for a reaction can be derived for a postulated mechanism (because the rate
law for each elementary step can be written), agreement between the experimental rate law
and postulated mechanism does not prove the mechanism. Different mechanisms can lead to
the same rate law.
Integrated rate equation: A mathematical equation which relates the concentration of a
species at a given time to its initial concentration and the time, and which involves the rate
constant
Example: For the 1st order reaction A
Products [A]t = [A]oekt where [A]t and [A]o are
the concentrations of reactant A at time t and at the beginning of the reaction respectively.
Half-life: Symbol t, the time for the amount of a reactant in limiting amount (i.e. not in
excess) to fall to half its original value. Commonly applied to radioactive isotopes
Example: The half-life of cobalt-60, 60Co, is 5.3 years. Thus a sample of radioactive Co-60
would have lost half of its radioactivity at the end of 5.3 years.
Mean life-time: Symbol , the time for the amount of a reactant in limiting amount (i.e. not in
excess) to fall to 0.368 (1/e) of its original value. This is the average time a molecule must
wait before undergoing reaction.
Chain reaction: A reaction in which an intermediate formed in an initiation step produces
further reactive intermediates in the propagation steps until removed in a termination step.
Chain length: The average number of propagating cycles between initiation and termination.

Catalyst: A substance which increases the rate of a reaction without being consumed in the
overall reaction. It provides an alternative pathway or mechanism which increases the rate of
the overall reaction. The catalyst is involved in the new pathway but if it is consumed in one
step it is regenerated in a later step.
Homogeneous catalysis: The catalyst is in the same phase as the reactants.
Examples:
Sulfuric acid in the esterification of an organic acid and alcohol.
The catalysis of the decomposition of hydrogen peroxide by potassium iodide.
Heterogeneous catalysis: The catalyst is in a different phase from the reactants. A solid
catalyst adsorbs reacting species onto its surface where the reaction takes place.
Example:
Solid V2O5 in the reaction of gaseous SO2 and O2 to form SO2.
An important step in the manufacture of sulphuric acid.

Types of Elementary Reactions


The molecularity of a reaction refers to the number of reactant particles involved in the reaction. Because there can only be discrete numbers of particles, the
molecularity must take an integer value. Molecularity can be described as unimolecular, bimolecular, or termolecular. There are no known elementary reactions involving
four or more molecules.1

Unimolecular Reaction
A unimolecular reaction occurs when a molecule rearranges itself to produce one or more products. An example of this is radioactive decay, in which particles are
emitted from an atom. Other examples include "cis-trans isomerization, thermal decomposition, ring opening, and racemization."1 The rate at which a substance
decomposes is dependent on its concentration. Unimolecular reactions are often first-order reactions as explained by Frederick Alexander Lindemann. 3 To read further
about Lindemann and why unimolecular reactions are not second-order,

Bimolecular Reaction
A bimolecular reaction involves the collision of two particles. Bimolecular reactions are common in organic reactions such as nucleophilic substitution. The rate of
reaction depends on the product of the concentrations of both species involved, which makes bimolecular reactions second-order reactions.

Termolecular Reaction
A termolecular reaction requires the collision of three particles at the same place and time. This type of reaction is very uncommon because all three reactants must
simultaneously collide with each other, with sufficient energy and correct orientation, to produce a reaction. There are three ways termolecular reactions can react, and
all are third order.

Practice Problems
1. How are non-elementary steps and elementary steps related?
2. Choose the correct statements.
a. An elementary step has 0 intermediates.
b. An elementary step has 1 intermediate.
c. An elementary step has 2 intermediates.
d. An elementary step has 0 transition states.
e. An elementary step has 1 transition state.
f. An elementary step has 2 transition states.
3. Which of the following elementrary reactions is a termolecular reaction?

a.

A+2B+CD

b.
c. all of the above

A+B+BC

d.
e. b and d
f. none of the above
4. Which rate law corresponds to a bimolecular reaction?

a.

b.
c. all of the above

2A+2B+2C2D

rate=k[A][A]2

rate=k[A][B]

d.
rate=k[A]2
e. b and d
f. none of the above
5. Give an example of a reaction with a molecularity of 1/2.
6. True or False: Given species A and B inside a container, instruments detect that three (3) collisions occured before product was formed. That is, we know a reaction
occured after detecting three collisions in a box. We can conclude that the reaction is a termolecular reaction (as the reaction could have been produced from A+A+B or
A+B+B).

Solutions
1. Non-elementary steps, or complex reactions, are sets of elementary reactions. The addition of elementary steps produces complex, non-elementary reactions.
2. The correct statements are "a" and "e". By definition of elementary reactions they have 0 intermediates because they cannot be broken down. Again by definition of
an elementary reaction, a single-step reaction will have 1 transition state. There is no reaction with 0 transition states. Having 2 transition states implies having 1
intermediate, making the reaction non-elementary.
3. "e" is the answer.
"a" is not a termolecular reaction because it involves A + B + B + C, or 4 molecules
"b" is a termolecular reaction because it involves 3 particles: A + B + B
"c" is incorrect because "a" is incorrect

"d" is a termolecular reaction, simplifying to the reaction:


particles (A + B + C)
"e" is the correct answer because "b" and "d" are correct
4. "e" is the answer.
"a" is incorrect because the rate law describes a third-order reaction, which is true for termolecular reactions

A+B+CD, which involves 3

"b" is a possible rate law for the bimolecular reaction:


A+BProducts
"c" is incorrect because "a" is incorrect

"d" is a possible rate law for the bimolecular reaction:


A+AProducts
"e" is the correct answer because "b" and "d" are correct
5. Impossible. The molecularity of a reaction MUST be an integer because there cannot be a "half particle" producing a reaction.
6. False; nothing can be concluded. Although a termolecular reaction requires the collision of three particles, the reverse logic is not necessarily true. That is, having
three collisions is not sufficient for a termolecular reaction.
1. For example, particles A + A + B collide with each other at the same place and time. However, particle B was in the wrong orientation, so no reaction occurred.
Instead, the two A particles were in the correct orientation and produced a reaction, which is a bimolecular reaction.

2. Consider a second example: two collisions between particles A + A + B occurred, but there was not enough energy to produce a reaction. Instead, a third collision
between A and B had the sufficient energy and correct orientation to produce a reaction. Such a reaction is, again, only bimolecular.
3. A last example: particle A collides twice with a wall, and then once with B to produce a reaction. Such a reaction involving three collisions at different places and
different time is only a bimolecular reaction.

14.5 Reaction Mechanisms


A balanced equation for a chemical reaction indicates the substances present at the start of the reaction
and those produced as the reaction proceeds. However, it provides no information about how the
reaction occurs. The process by which a reaction occurs is called the reaction mechanism. At the
most sophisticated level, a reaction mechanism will describe in great detail the order in which bonds
are broken and formed and the changes in relative positions of the atoms in the course of the reaction.
In addition, the mechanism of a reaction can change as the temperature changes. We will begin with
more rudimentary descriptions of how reactions occur.

Elementary Steps
We have seen that reactions take place as a result of collisions between reacting molecules. For
example, the collisions between molecules of methyl isonitrile, CH 3NC, can provide the energy to
allow the CH3NC to rearrange:

Similarly, the reaction of NO and O3 to form NO2 and O2 appears to occur as a result of a single
collision involving suitably oriented and sufficiently energetic NO and O 3 molecules:

[14.23]
Both of these processes occur in a single event or step and are called elementary steps (or elementary
processes).
The number of molecules that participate as reactants in an elementary step defines
the molecularity of the step. If a single molecule is involved, the reaction is said to
be unimolecular. The rearrangement of methyl isonitrile is a unimolecular process. Elementary steps
involving the collision of two reactant molecules are said to be bimolecular. The reaction between NO
and O3 (Equation 14.23) is bimolecular. Elementary steps involving the simultaneous collision of three
molecules are said to be termolecular. Termolecular steps are far less probable than unimolecular or
bimolecular processes and are rarely encountered. The chance that four or more molecules will collide
simultaneously with any regularity is even more remote; consequently, such collisions are never
proposed as part of a reaction mechanism.

Multistep Mechanisms
The net change represented by a balanced chemical equation often occurs by a multistep
mechanism, which consists of a sequence of elementary steps. For example, consider the reaction of
NO2 and CO:

[14.24]
Below 225C this reaction appears to proceed in two elementary steps, each of which is bimolecular.
First, two NO2 molecules collide, and an oxygen atom is transferred from one to the other. The
resultant NO3 then transfers an oxygen atom to CO during a collision between these molecules:

The elementary steps in a multistep mechanism must always add to give the chemical equation of the
overall process. In the present example the sum of the elementary steps is

Simplifying this equation by eliminating substances that appear on both sides of the arrow gives the
net equation for the process, Equation 14.24. Because NO 3 is neither a reactant nor a product in the
overall reactionit is formed in one elementary step and consumed in the nextit is called
an intermediate. Multistep mechanisms involve one or more intermediates.

SAMPLE EXERCISE 14.9


It has been proposed that the conversion of ozone into O 2 proceeds via two elementary steps:

(a) Describe the molecularity of each step in this mechanism. (b) Write the equation for the overall
reaction. (c) Identify the intermediate, if any.
SOLUTION (a) The first elementary step involves a single reactant and is consequently unimolecular.
The second step, which involves two reactant molecules, is bimolecular.
(b) Adding the two elementary steps gives

Because O(g) appears in equal amounts on both sides of the equation, it can be eliminated to give the
net equation for the chemical process:

(c) The intermediate is O(g). It is neither an original reactant nor a final product but is formed in the
first step and consumed in the second.
PRACTICE EXERCISE
For the following reaction of Mo(CO)6,

the proposed mechanism is

(a) Is the proposed mechanism consistent with the equation for the overall reaction? (b) Identify the
intermediate or intermediates. Answers: (a) yes, the two equations add to yield the equation for the
reaction; (b) Mo(CO)5

Rate Laws of Elementary Steps


In Section 14.2 we stressed that rate laws must be determined experimentally; they cannot be predicted
from the coefficients of balanced chemical equations. We are now in a position to understand why this
is so: Every reaction is made up of a series of one or more elementary steps, and the rate laws and
relative speeds of these steps will dictate the overall rate law. Indeed, the rate law for a reaction can be
determined from its mechanism, as we will see shortly. Thus, our next challenge in kinetics is to arrive
at reaction mechanisms that lead to rate laws that are consistent with those observed experimentally.
We will start by examining the rate laws of elementary steps.
Elementary steps are significant in a very important way: If we know that a reaction is an elementary
step, then we know its rate law. The rate law of any elementary step is based directly on its
molecularity. For example, consider the general unimolecular process

As the number of A molecules increases, the number that decompose in a given interval of time will
increase proportionally. Thus, the rate of a unimolecular process will be first order:

In the case of bimolecular elementary steps, the rate law is second order, as in the following example:

The second-order rate law follows directly from the collision theory. If we double the concentration of
A, the number of collisions between molecules of A and B will double; likewise, if we double [B], the
number of collisions will double. Therefore, the rate law will be first order in both [A] and [B], and
second order overall.
The rate laws for all feasible elementary steps are given in Table 14.4. Notice how the rate law for each
kind of elementary step follows directly from the molecularity of that step. It is important to remember,
however, that we cannot tell by merely looking at a balanced chemical equation whether the reaction
involves one or several elementary steps.

SAMPLE EXERCISE 14.10


If the following reaction occurs in a single elementary step, predict the rate law:

SOLUTION Because we are assuming that the reaction occurs as a single elementary step, we are
able to write the rate law. (Remember, we are making an assumption. We have no way of knowing
whether the reaction occurs via a single elementary step.) Under this assumption the elementary step is
bimolecular, involving the collision of one molecule of H 2 with one molecule of Br2. The rate law is
therefore first order in each reactant and second order overall:

Experimental studies of this reaction show that it follows a very different rate law:

This experimental rate law is different from the one obtained by assuming that the reaction occurs by a
single elementary step. Thus, a more complicated reaction mechanism, involving more than one
elementary step, must be used to explain the experimentally observed rate law.
PRACTICE EXERCISE
Consider the following reaction: 2NO(g) + Br2(g)
2NOBr(g). (a) Write the rate law for the
reaction, assuming it involves a single elementary step. (b) Is a single-step mechanism likely for this
reaction?Answers: (a) rate = k[NO]2[Br2]; (b) no, because termolecular reactions are very rare

Rate Laws of Multistep Mechanisms


As with the reaction in Sample Exercise 14.10, most chemical reactions occur by mechanisms that
involve more than one elementary step. Often one of the steps is much slower than the others. The
overall rate of a reaction cannot exceed the rate of the slowest elementary step of its mechanism.
Because the slow step limits the overall reaction rate, it is called the rate-determining step.
To understand the concept of a rate-determining step, it is helpful to think of a toll road with two toll
plazas (Figure 14.19).

Figure 14.19 The flow of traffic on a toll road is limited by the flow of traffic through the slowest toll
plaza. As cars pass from point 1 to point 3, they pass through plazas A and B. In (a) the rate at which
cars can reach point 3 is limited by how quickly they can get through plaza A; getting from point 1 to
point 2 is the rate-determining step. In (b) getting from point 2 to point 3 is the rate-determining step.
We will measure the rate at which cars exit the toll road. Cars enter the toll road at point 1 and pass
through toll plaza A. They then pass an intermediate point 2 before passing through toll plaza B. Upon
exiting, they pass point 3. We can therefore envision this trip along the toll road as occurring in two
elementary steps:

Now suppose that several of the gates at toll plaza A are malfunctioning, so that traffic backs up behind
it [Figure 14.19(a)]. The rate at which cars can get to point 3 is limited by the rate at which they can
get through the traffic jam at plaza A. Thus, step 1 is the rate-determining step of the journey along the
toll road. If, however, traffic flows quickly through plaza A but gets backed up at plaza B [Figure
14.19(b)], there will be a buildup of cars in the intermediate region between the plazas. In this case
step 2 is the rate-determining step: The rate at which cars can travel the toll road is limited by the rate
at which they can pass through plaza B.
In the same way, the slowest step in a multistep reaction determines the overall rate. By analogy to
Figure 14.19(a), the rate of a faster step following the rate-determining step does not affect the overall
rate. If the slow step is not the first one, as in Figure 14.19(b), the faster preceding steps produce
intermediate products that accumulate before being consumed in the slow step. In either case the ratedetermining step governs the rate law for the overall reaction.

As an example of a slow first step determining the rate law of a reaction, consider the reaction of
NO2 and CO to produce NO and CO2 (Equation 14.24). Below 225C it is found experimentally that
the rate law is second order in NO2 and zero order in CO: Rate = k[NO2]2. Can we propose a reaction
mechanism that is consistent with this rate law? Consider the following two-step mechanism: (The
subscript on the rate constant identifies the elementary step involved. Thus, k1 is the rate constant for
step 1, k2 is the rate constant for step 2, and so forth. A negative subscript refers to the rate constant for
the reverse of an elementary step. For example, k1 is the rate constant for the reverse of the first step.)

Step 2 is much faster than step 1; that is, k2


and is immediately consumed in step 2.

k1. The intermediate NO3(g) is slowly produced in step 1

Because step 1 is slow and step 2 is fast, step 1 is the rate-determining step. Thus, the rate of the
overall reaction equals the rate of step 1, and the rate law of the overall reaction equals the rate law of
step 1. Step 1 is a bimolecular process that has the rate law

Thus, the rate law predicted by this mechanism agrees with the one observed experimentally.
Could we propose a one-step mechanism for the preceding reaction? We might suppose that the overall
reaction is a single bimolecular elementary process that involves the collision of a molecule of
NO2with one of CO. However, the rate law predicted by this mechanism would be

Because this mechanism predicts a rate law different from that observed experimentally, we can rule it
out.

Mechanisms with an Initial Fast Step


It is not easy to derive the rate law for a mechanism in which an intermediate is a reactant in the ratedetermining step. This situation arises in multistep mechanisms when the first step is not ratedetermining. Let's consider one example, the gas-phase reaction of nitric oxide, NO, with bromine,
Br2:

[14.25]
The experimentally determined rate law for this reaction is second order in NO and first order in Br 2:

[14.26]
We seek a reaction mechanism that is consistent with this rate law. One possibility is that the reaction
occurs in a single termolecular step:

[14.27]
As noted in Practice Exercise 14.10, this does not seem likely because termolecular processes are so
rare. Let's consider an alternative mechanism that does not invoke termolecular steps:

We see that in this mechanism step 1 actually involves two processes: a forward reaction and its
reverse.
Because step 2 is the slow, rate-determining step, the rate of the overall reaction is governed by the rate
law for that step:

[14.28]
However, NOBr2 is an intermediate generated in step 1. Intermediates are usually unstable molecules
that have a low, unknown concentration. Thus, we have a problem in that our rate law depends on the
unknown concentration of an intermediate. Fortunately, with the aid of some assumptions, we can
express the concentration of NOBr2 in terms of the concentrations of NO and Br 2. We first assume that
NOBr2 is intrinsically unstable, and that it does not accumulate to a significant extent in the reaction
mixture. There are two ways for NOBr2 to be consumed once it is formed: It can either react with NO
to form NOBr or fall back apart into NO and Br 2. The first of these possibilities is step 2, a slow
process. The second is the reverse of step 1, a unimolecular process:

[14.29]
Because step 2 is slow, we assume that most of the NOBr 2 falls apart according to Equation 14.29.
Thus, we have both the forward and reverse reactions of step 1 occurring much faster than step 2.
Because they occur rapidly with respect to the reaction in step 2, the forward and reverse processes of
step 1 establish an equilibrium. We have seen examples of dynamic equilibrium before, in the
equilibrium between a liquid and its vapor (Section 11.5) and between a solid solute and its solution
(Section 13.3). As in any dynamic equilibrium, the rates of the forward and reverse reactions are equal.
Thus we can equate the rate expression for the forward reaction in step 1 with the rate expression for
the reverse reaction:

Solving for [NOBr2], we have

Substituting this relationship into the rate law for the rate-determining step (Equation 14.28), we have

This is consistent with the experimental rate law (Equation 14.26). The experimental rate constant k is
equal to k2k1 / k1. This mechanism, which involves only unimolecular and bimolecular processes, is far
more probable than the single termolecular step (Equation 14.27).
In general, whenever a fast step precedes a slow one, we can solve for the concentration of an
intermediate by assuming that an equilibrium is established in the fast step.

SAMPLE EXERCISE 14.11


Show that the following mechanism for Equation 14.25 also produces a rate law consistent with the
experimentally observed one:

SOLUTION The second step is rate determining, and so the overall rate is

We solve for the concentration of the intermediate N 2O2 by assuming that an equilibrium is established
in step 1; thus, the rates of the forward and reverse reactions in step 1 are equal:

Substituting this expression into the rate expression gives

Thus, this mechanism also yields a rate law consistent with the experimental one.

PRACTICE EXERCISE
The first step of a mechanism involving the reaction of bromine is

What is the expression relating the concentration of Br(g) to that of Br2(g)?


Answer:

Some reactions are slow, such as rusting, and some are fast, like burning. The rate of reaction depends on the temperature and
concentration of the reactants, and the surface area of any solid reactants.
The rate of reaction can be found by measuring the amount of reactant used up, or the amount of product formed, in a given time.
Catalysts increase the rate of a reaction without being changed themselves by the end of the reaction.

Collision theory
Different reactions can happen at different rates. Reactions that occur slowly have a low rate of reaction. Reactions that happen quickly have a
high rate of reaction. For example, rusting is a slow reaction: it has a low rate of reaction. Burning and explosions are very fast reactions: they
have a high rate of reaction.

Collisions
For a chemical reaction to occur, the reactant particles must collide. But collisions with too little energy do not produce a reaction.

The particles must have enough energy for the collision to be successful in producing a reaction.

The rate of reaction depends on the rate of successful collisions between reactant particles. The more successful collisions there are, the faster the
rate of reaction.

Measuring rates of reaction


There are two ways to find the rate of a reaction:

measure the rate at which a reactant is used up


measure the rate at which a product is formed
The method chosen depends on the reaction being studied. Sometimes it is easier to measure the change in the amount of a reactant that has
been used up; sometimes it is easier to measure the change in the amount of a product that has been produced.

Things to measure
The measurement itself depends on the nature of the reactant or product:

the mass of a substance - solid, liquid or gas - is measured with a balance


the volume of a gas is usually measured with a gas syringe, or sometimes an upside-down measuring cylinder or burette

Measuring the production of a gas using a gas syringe

Effect of temperature and concentration


The rate of a chemical reaction can be increased by raising the temperature. It can also be increased by increasing the concentration of a reactant
in solution, or the pressure of a reactant gas.

Changing the temperature


If the temperature is increased:

the reactant particles move more quickly


they have more energy
the particles collide more often, and more of the collisions result in a reaction
the rate of reaction increases

Changing the concentration or pressure


If the concentration of a dissolved reactant is increased, or the pressure of a reacting gas is increased:

the reactant particles become more crowded


there is a greater chance of the particles colliding
the rate of reaction increases

Effect of surface area


The rate of a chemical reaction can be raised by increasing the surface area of a solid reactant. This is done by cutting the substance into small
pieces, or grinding it into a powder.

Lumps react more slowly than powders

Changing the surface area


If a solid reactant is broken into small pieces or ground into a powder:

its surface area increases


more particles are exposed to the other reactant
there are more collisions
the rate of reaction increases

Explosions
Explosions are very fast reactions in which a lot of gaseous product is released very quickly. For example:

burning hydrogen
explosions from TNT or dynamite
Fine powders can also cause explosions - for example, custard powder, flour or sulfur. This is a particular hazard in factories. Great care is needed
to avoid naked flames or other sources of ignition.

Effect of catalysts
A catalyst is a substance that can increase the rate of a reaction. The catalyst itself remains unchanged at the end of the reaction it catalyses.
Only a very small amount of catalyst is needed to increase the rate of reaction between large amounts of reactants.
Different catalysts catalyse different reactions. The table summarises some common catalysts used in industry and the reactions they catalyse:

Some common catalysts used in industry and the reactions they catalyse
catalyst

reaction catalysed

iron

making ammonia from nitrogen and hydrogen

platinum

making nitric acid from ammonia

vanadium(V) oxide

making sulphuric acid

Now try a Test Bite - foundation

Calculating rates - higher


The faster the rate, the more reactant is used, or product is made, in a given time. The faster the reaction, the steeper the line on a graph
showing total product against time will be.
A reaction finishes when one of the reactants is all used up. No more product is made, so the line on a graph of total product against time will
become horizontal.
The rate can be calculated using this equation:
rate of reaction = total amount of reactant used or product made time taken

You might also like