You are on page 1of 30

1

2. Chemical Kinetics

2.1. Reaction Rates


Reaction rates are usually expressed as the concentration of reactant consumed or the
concentration of product formed per unit time. The units are thus moles per liter per unit
time, written as M/s, M/min, or M/h. To measure reaction rates, chemists initiate the
reaction, measure the concentration of the reactant or product at different times as the
reaction progresses, perhaps plot the concentration as a function of time on a graph, and
then calculate the change in the concentration per unit time.

Figure 2.1.1: The Progress of a Simple Reaction (A → B). The mixture initially contains only A molecules
(purple). Over time, the number of A molecules decreases and more B molecules (green) are formed (top).
The graph shows the change in the number of A and B molecules in the reaction as a function of time
over a 1 min period (bottom).

The progress of a simple reaction (A → B) is shown in Figure 2.1.1; the beakers are
snapshots of the composition of the solution at 10 s intervals. The number of molecules
of reactant (A) and product (B) are plotted as a function of time in the graph. Each point
in the graph corresponds to one beaker in Figure 2.1.1. The reaction rate is the change
in the concentration of either the reactant or the product over a period of time. The
concentration of A decreases with time, while the concentration of B increases with time.

Some content reproduced from:


1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
2

∆[B] ∆[A]
𝑟𝑎𝑡𝑒 = =− (2.1.1)
∆𝑡 ∆𝑡

Square brackets indicate molar concentrations, and the capital Greek delta (Δ) means
“change in.” Because chemists follow the convention of expressing all reaction rates as
positive numbers, however, a negative sign is inserted in front of Δ[A]/Δt to convert that
expression to a positive number. The reaction rate calculated for the reaction A → B
using Equation 2.1.1 is different for each interval (this is not true for every reaction, as
shown below). A greater change occurs in [A] and [B] during the first 10 s interval, for
example, than during the last, meaning that the reaction rate is greatest at first.

Determining the Reaction Rate of Hydrolysis of Aspirin


We can use Equation 2.2.1 to determine the reaction rate of hydrolysis of aspirin,
probably the most commonly used drug in the world (more than 25,000,000 kg are
produced annually worldwide). Aspirin (acetylsalicylic acid) reacts with water (such as
water in body fluids) to give salicylic acid and acetic acid, as shown in Figure 2.1.2.

Figure 2.1.2: Hydrolysis of Aspirin reaction.

Because salicylic acid is the actual substance that relieves pain and reduces fever and
inflammation, a great deal of research has focused on understanding this reaction and
the factors that affect its rate. Data for the hydrolysis of a sample of aspirin are in
Table 2.1 and are shown in the graph in Figure 2.1.3.

Table 2.1 Data for Aspirin Hydrolysis in Aqueous Solution at pH 7.0 and 37°C*
Time (h) [Aspirin] (M) [Salicylic Acid] (M)
0 5.55 × 10−3 0
2.0 5.51 × 10−3 0.040 × 10−3
5.0 5.45 × 10−3 0.10 × 10−3
10 5.35 × 10−3 0.20 × 10−3
20 5.15 × 10−3 0.40 × 10−3

Some content reproduced from:


1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
3

30 4.96 × 10−3 0.59 × 10−3


40 4.78 × 10−3 0.77 × 10−3
50 4.61 × 10−3 0.94 × 10−3
100 3.83 × 10−3 1.72 × 10−3
200 2.64 × 10−3 2.91 × 10−3
300 1.82 × 10−3 3.73 × 10−3
*The reaction at pH 7.0 is very slow. It is much faster under acidic conditions, such as
those found in the stomach.

The data in Table 2.1 were obtained by removing samples of the reaction mixture at the
indicated times and analyzing them for the concentrations of the reactant (aspirin) and
one of the products (salicylic acid).

Figure 2.1.3: The Hydrolysis of Aspirin. This graph shows the concentrations of aspirin and salicylic acid
as a function of time, based on the hydrolysis data in Table 1. The time dependence of the concentration
of the other product, acetate, is not shown, but based on the stoichiometry of the reaction, it is identical
to the data for salicylic acid.

The average reaction rate for a given time interval can be calculated from the
concentrations of either the reactant or one of the products at the beginning of the
interval (time = t0) and at the end of the interval (t1). Using salicylic acid, the reaction rate
for the interval between t = 0 h and t = 2.0 h (recall that change is always calculated as
final minus initial) is calculated as follows:
Some content reproduced from:
1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
4

[salicyclic acid]2 − [salicyclic acid]0


𝑟𝑎𝑡𝑒(𝑡=0−2.0 ℎ) =
2.0 h − 0 h

0.040 × 10−3 M − 0 M
𝑟𝑎𝑡𝑒(𝑡=0−2.0 ℎ) =
2.0 h

rate(t=0−2.0 h) = 2.0 × 10−5 M/h

The reaction rate can also be calculated from the concentrations of aspirin at the
beginning and the end of the same interval, remembering to insert a negative sign,
because its concentration decreases:
[aspirin]2 − [aspirin]0
𝑟𝑎𝑡𝑒(𝑡=0−2.0 ℎ) = −
2.0 h − 0 h

5.51 × 10−3 M − 5.55 × 10−3 M


𝑟𝑎𝑡𝑒(𝑡=0−2.0 ℎ) =
2.0 h

𝑟𝑎𝑡𝑒(𝑡=0−2.0 ℎ) = 2.0 × 10−5 M/h

If the reaction rate is calculated during the last interval given in Table 1 (the interval
between 200 h and 300 h after the start of the reaction), the reaction rate is significantly
slower than it was during the first interval (t = 0–2.0 h):

[salicyclic acid]2 − [salicyclic acid]0


𝑟𝑎𝑡𝑒(𝑡=200−300 ℎ) =
300 h − 200 h

3.73 × 10−3 M − 2.91 × 10−3 M


𝑟𝑎𝑡𝑒(𝑡=200−300 ℎ) =
100 h

𝑟𝑎𝑡𝑒(𝑡=200−300 ℎ) = 8.2 × 10−6 M/h

Calculating the Reaction Rate of Fermentation of Sucrose


In the preceding example, the stoichiometric coefficients in the balanced chemical
equation are the same for all reactants and products; that is, the reactants and products

Some content reproduced from:


1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
5

all have the coefficient 1. Consider a reaction in which the coefficients are not all the
same, the fermentation of sucrose to ethanol and carbon dioxide:

C12H22O11(aq) + H2O(l) → 4C2H5OH(aq)+4CO2(g)

The coefficients indicate that the reaction produces four molecules of ethanol and four
molecules of carbon dioxide for every one molecule of sucrose consumed. As before, the
reaction rate can be found from the change in the concentration of any reactant or
product. In this particular case, however, a chemist would probably use the concentration
of either sucrose or ethanol because gases are usually measured as volumes and the
volume of CO2 gas formed depends on the total volume of the solution being studied
and the solubility of the gas in the solution, not just the concentration of sucrose. The
coefficients in the balanced chemical equation tell us that the reaction rate at which
ethanol is formed is always four times faster than the reaction rate at which sucrose is
consumed:

Δ[C2 H5 OH] 4Δ[C12 H22 O11 ]


=− (2.1.2)
Δ𝑡 Δ𝑡

The concentration of the reactant—in this case sucrose—decreases with time, so the
value of Δ[sucrose] is negative. Consequently, a minus sign is inserted in front of
Δ[sucrose] in Equation 2.1.2 so the rate of change of the sucrose concentration is
expressed as a positive value. Conversely, the ethanol concentration increases with time,
so its rate of change is automatically expressed as a positive value.
Often the reaction rate is expressed in terms of the reactant or product with the smallest
coefficient in the balanced chemical equation. The smallest coefficient in the sucrose
fermentation reaction corresponds to sucrose, so the reaction rate is generally defined
as follows:

1 Δ[C2 H5 OH] Δ[C12 H22 O11 ]


𝑟𝑎𝑡𝑒 = =− (2.1.3)
4 Δ𝑡 Δ𝑡

In general, for the reaction

aA + bB ⟶ cC + dD
Then
1 ∆[A] 1 ∆[B] 1 ∆[C] 1 ∆[D]
𝑟𝑎𝑡𝑒 = − =− = = (2.1.4)
𝑎 ∆𝑡 𝑏 ∆𝑡 𝑐 ∆𝑡 𝑑 ∆𝑡
Some content reproduced from:
1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
6

Example 2.1.1: Expressions for Relative Reaction Rates


Consider the thermal decomposition of gaseous N2O5 to NO2 and O2 via the following
equation:

2N2 O5 (𝑔) → 4NO2 (𝑔) + O2 (𝑔)

Write expressions for the reaction rate in terms of the rates of change in the
concentrations of the reactant and each product with time.

Strategy:
A. Choose the species in the equation that has the smallest coefficient. Then write an
expression for the rate of change of that species with time.
B. For the remaining species in the equation, use molar ratios to obtain equivalent
expressions for the reaction rate.

Solution
A. Because O2 has the smallest coefficient in the balanced chemical equation for the
reaction, define the reaction rate as the rate of change in the concentration of O2 and
write that expression.
B. The balanced chemical equation shows that 2 mol of N2O5 must decompose for each
1 mol of O2 produced and that 4 mol of NO2 are produced for every 1 mol of
O2 produced. The molar ratios of O2 to N2O5 and to NO2 are thus 1:2 and 1:4,
respectively. This means that the rate of change of [N2O5] and [NO2] must be divided
by its stoichiometric coefficient to obtain equivalent expressions for the reaction rate.
For example, because NO2 is produced at four times the rate of O2, the rate of
production of NO2 is divided by 4. The reaction rate expressions are as follows:

∆[O2 ] 1 ∆[NO2 ] 1 ∆[N2 O5 ]


𝑟𝑎𝑡𝑒 = = =−
∆𝑡 4 ∆𝑡 2 ∆𝑡

Example 2.1.2: Rates and Reaction Stoichiometry


For the oxidation of ammonia:

4NH3(g)+ 3O2(g) ⟶ 2N2(g)+ 6H2O(g)

it was found that the rate of formation of N2 was 0.27 mol L–1 s–1.
a) At what rate was H2O being formed?
b) At what rate was ammonia being consumed?
c) At what rate was O2 disappearing?

Solution
The rate of the reaction is

1 ∆[NH3 ] 1 ∆[O2 ] 1 ∆[N2 ] 1 ∆[H2 O]


𝑟𝑎𝑡𝑒 = − =− = =
4 ∆𝑡 3 ∆𝑡 2 ∆𝑡 6 ∆𝑡

Some content reproduced from:


1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
7

a) From the equation stoichiometry,

1 ∆[N2 ] 1 ∆[H2 O]
=
2 ∆𝑡 6 ∆𝑡

And the rate of formation of H2O is

∆[H2 O] 6 ∆[N2 ]
= = 3 × 0.27 mol L−1 s−1 = 0.81 mol L−1 s−1
∆𝑡 2 ∆𝑡

b) 4 moles of NH3 are consumed for every 2 moles of N2 formed, so the rate of
disappearance of ammonia is

1 ∆[NH3 ] 1 ∆[N2 ]
− =
4 ∆𝑡 2 ∆𝑡
or

∆[NH3 ] 4 ∆[N2 ]
− = = 2 × 0.27 mol L−1 s−1 = 0.54 mol L−1 s−1
∆𝑡 2 ∆𝑡

or
∆[NH3 ]
= −0.54 mol L−1 s−1
∆𝑡

Instantaneous Rates of Reaction


The instantaneous rate of a reaction is the reaction rate at any given point in time. As the
period of time used to calculate an average rate of a reaction becomes shorter and
shorter, the average rate approaches the instantaneous rate. Comparing this to calculus,
the instantaneous rate of a reaction at a given time corresponds to the slope of a line
tangent to the concentration-versus-time curve at that point—that is, the derivative of
concentration with respect to time.

The instantaneous rate of a reaction at “time zero,” when the reaction commences, is
its initial rate. Consider the analogy of a car slowing down as it approaches a stop sign.
The vehicle’s initial rate—analogous to the beginning of a chemical reaction—would be
the speedometer reading at the moment the driver begins pressing the brakes (t0). A few
moments later, the instantaneous rate at a specific moment—call it t1—would be
somewhat slower, as indicated by the speedometer reading at that point in time. As time
passes, the instantaneous rate will continue to fall until it reaches zero, when the car (or
reaction) stops. Unlike instantaneous speed, the car’s average speed is not indicated by
the speedometer; but it can be calculated as the ratio of the distance travelled to the
time required to bring the vehicle to a complete stop (Δt). Like the decelerating car, the
average rate of a chemical reaction will fall somewhere between its initial and final rates.
Some content reproduced from:
1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
8

The instantaneous rate of a reaction may be determined one of two ways. If experimental
conditions permit the measurement of concentration changes over very short time
intervals, then average rates computed as described earlier provide reasonably good
approximations of instantaneous rates. Alternatively, a graphical procedure may be used
that, in effect, yields the results that would be obtained if short time interval
measurements were possible. In a plot of the concentration of hydrogen peroxide against
time, the instantaneous rate of decomposition of H2O2 at any time t is given by the slope
of a straight line that is tangent to the curve at that time (Figure 2.1.4).

Figure 2.1.4 This graph shows a plot of concentration versus time for a 1.000 M solution of H2O2. The rate
at any time is equal to the negative of the slope of a line tangent to the curve at that time. Tangents are
shown at t = 0 h (“initial rate”) and at t = 12 h (“instantaneous rate” at 12 h). Access for free at
https://openstax.org/books/chemistry-2e/pages/1-introduction

These tangent slopes are derivatives whose values can very at each point on the curve,
so that these instantaneous rates are really limiting rates defined as:

𝑑[H2 O2 ]
𝑟𝑎𝑡𝑒 = (2.1.5)
𝑑𝑡

Some content reproduced from:


1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
9

2.2. Concentration and Rates (Differential Rate Laws)

2.2.1. Rate Laws


Typically, reaction rates decrease with time because reactant concentrations decrease as
reactants are converted to products. Reaction rates generally increase when reactant
concentrations are increased. This section examines mathematical expressions called rate
laws, which describe the relationships between reactant rates and reactant
concentrations. Rate laws are mathematical descriptions of experimentally verifiable data.

Rate laws may be written from either of two different but related perspectives. A
differential rate law expresses the reaction rate in terms of changes in the concentration
of one or more reactants (Δ[R]) over a specific time interval (Δt). In contrast, an integrated
rate law describes the reaction rate in terms of the initial concentration ([R] 0) and the
measured concentration of one or more reactants ([R]) after a given amount of time (t);
integrated rate laws are discussed in more detail later. The integrated rate law is derived
by using calculus to integrate the differential rate law. Whether using a differential rate
law or integrated rate law, always make sure that the rate law gives the proper units for
the reaction rate, usually moles per liter per second (M/s).

2.2.2. Reaction Orders


For a reaction with the general equation:

aA + bB ⟶ cC + dD (2.2.1)

the experimentally determined rate law usually has the following form:

𝑟𝑎𝑡𝑒 = 𝑘[A]𝑚 [B]𝑛 (2.2.2)

The proportionality constant (k) is called the rate constant, and its value is characteristic
of the reaction and the reaction conditions. A given reaction has a particular rate constant
value under a given set of conditions, such as temperature, pressure, and solvent; varying
the temperature or the solvent usually changes the value of the rate constant. The
numerical value of k, however, does not change as the reaction progresses under a given
set of conditions.
Under a given set of conditions, the value of the rate constant does not change as the
reaction progresses.

Some content reproduced from:


1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
10

The reaction rate thus depends on the rate constant for the given set of reaction
conditions and the concentration of A and B raised to the powers m and n, respectively.
The values of m and n are derived from experimental measurements of the changes in
reactant concentrations over time and indicate the reaction order, the degree to which
the reaction rate depends on the concentration of each reactant; m and n need not be
integers. For example, Equation 2.2.2 tells us that Equation 2.2.1 is mth order in reactant
A and nth order in reactant B. It is important to remember that n and m are not related
to the stoichiometric coefficients a and b in the balanced chemical equation and must be
determined experimentally. The overall reaction order is the sum of all the exponents in
the rate law: m + n.
The orders of the reactions (e.g. n and m) are not related to the stoichiometric
coefficients in the balanced chemical (e.g., a and b).

Example 2.2.1: Writing Rate Laws from Reaction Orders


An experiment shows that the reaction of nitrogen dioxide with carbon monoxide:

NO2(g)+CO(g)⟶NO(g)+CO2(g)

is second order in NO2 and zero order in CO at 100 °C. What is the rate law for the
reaction?

Solution
The reaction will have the form:

𝑟𝑎𝑡𝑒 = 𝑘[NO2 ]𝑚 [CO]𝑛

The reaction is second order in NO2; thus m = 2. The reaction is zero order in CO;
thus n = 0. The rate law is:

𝑟𝑎𝑡𝑒 = 𝑘[NO2 ]2 [CO]0 = 𝑘[NO2 ]2

Remember that a number raised to the zero power is equal to 1, thus [CO] 0 = 1, which
is why we can simply drop the concentration of CO from the rate equation: the rate of
reaction is solely dependent on the concentration of NO2.

Example 2.2.2: Effect of Changing Concentration of Reactants on the Rate


The rate of oxidation of Br– ions by BrO3– ions in an acidic aqueous solution:

6H+(aq) + BrO3–(aq)+ 5Br–(aq)→ 3Br2(aq)+ 3H2O(l)

is found to follow the rate law:

rate = k[Br–][BrO3–][H+]2

Some content reproduced from:


1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
11

What happens to the rate if


a) [BrO3–] is doubled,
b) [H+] is decreased by a factor of 0.5?

Solution
a) Since the rate is first order in bromate, doubling its concentration will double the
reaction rate.

b) Since the reaction is second order in [H+], decreasing [H+] by a factor of 0.5 will
decrease the rate by a factor of 0.52 or by 1/4.

Example 2.2.3: Order of Reaction


How much will each of the following affect the rate of the reaction:
CO(g) + NO2(g) ⟶ CO2(g) + NO(g)
if the rate law for the reaction is rate = k[NO2]2?
a) Decreasing the pressure of NO2 from 0.50 atm to 0.250 atm.
b) Increasing the concentration of CO from 0.01 M to 0.03 M.
Solution
a) Because pressure is proportional to concentration for gaseous reactants, the rate law
can be written as
2
𝑟𝑎𝑡𝑒 = [NO2 ]2 = 𝑃NO 2

If rate at 0.50 atm is rate1 and rate at 0.25 atm is rate2 then

𝑟𝑎𝑡𝑒1 𝑘(0.50 atm)2 2 2


= =( ) =4
𝑟𝑎𝑡𝑒2 𝑘(0.25 atm)2 1
1
𝑟𝑎𝑡𝑒2 = 4𝑟𝑎𝑡𝑒1

Therefore, the process reduces the rate by a factor of 4.

b) Since CO does not appear in the rate law, the rate is not affected.

Zero order means that the rate is independent of the concentration of a particular
reactant.

2.2.3. Reaction Order and Rate Constant Units


In some of our examples, the reaction orders in the rate law happen to be the same as
the coefficients in the chemical equation for the reaction. This is merely a coincidence
and very often not the case.

Some content reproduced from:


1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
12

Rate laws may exhibit fractional orders for some reactants, and negative reaction orders
are sometimes observed when an increase in the concentration of one reactant causes a
decrease in reaction rate. A few examples illustrating these points are provided:

NO2 + CO ⟶ NO + CO2 rate = 𝑘[NO2 ]2


CH3 CHO ⟶ CO + CH4 rate = 𝑘[CH3 CHO]2
2N2 O5 ⟶ 2NO2 + O2 rate = 𝑘[2N2 O5 ]
2NO2 + F2 ⟶ 2NO2 F rate = 𝑘[NO2 ][F2 ]
2NO2 Cl ⟶ 2NO2 + Cl2 rate = 𝑘[NO2 Cl]

It is important to note that rate laws are determined by experiment only and are not
reliably predicted by reaction stoichiometry.
The units for a rate constant will vary as appropriate to accommodate the overall order
of the reaction. Dimensional analysis requires the rate constant unit for a reaction whose
overall order is n to be L(𝑛−1) ∙ mol(1−𝑛) ∙ s −1 . Table 2.2 summarizes the rate constant
units for common reaction orders.

Table 2.2 Rate Constant Units for Common Reaction Orders


Order Units of k
0 mol ∙ L−1 ∙ s−1
1 s−1
2 L ∙ mol−1 ∙ s −1
3 L2 ∙ mol−2 ∙ s−1
n (𝑛 ≠ 1) L(n−1) ∙ mol(1−n) ∙ s−1

2.2.4. Experimental Determination of Rate Laws


To determine the value of the exponent in a rate equation term, we need to see how the
rate varies with the concentration of the substance.
For a single-reactant decomposition reaction of the form
A → products
in which the rate is –d[A]/dt, we simply plot [A] as a function of time, draw tangents at
various intervals, and see how the slopes of these tangents (the instantaneous rates)
depend on [A]:
• If doubling the concentration of A doubles the rate, then the reaction is first order
in A.
Some content reproduced from:
1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
13

• If doubling the concentration results in a fourfold rate increase, the reaction is


second order in A.
A common experimental approach to the
determination of rate laws is the method of initial
rates. In the initial rate method, we measure only the rate
near the beginning of the reaction before the
concentrations have had time to change significantly. The
experiment is then repeated with a different starting
concentration of the reactant in question but keeping the
concentrations of any others the same. After the order
with respect to one component is found, another series of
trials is conducted in which the order of another
component is found. This approach is illustrated in the
next two example exercises.

Figure 2.2.1 The determination of reaction order for the reaction


2N2O5(g)→ 4NO2(g)+ O2(g). In this example, a series of runs
using five different initial concentrations of N2O5 has been made.
The slopes at t = 0 have been measured. These are the initial
rates.

Example 2.2.4: Determining a Rate Law from Initial Rates


Consider reaction:

F2(g) + 2ClO2(g) → 2FClO2(g)

Use initial rate data below to determine values of exponents x and y in the rate law:
Rate = k[F2]x[ClO2]y. Find value of rate constant k

Solution
Dividing rate from experiment 1 by rate from experiment 2, we get

𝑟𝑎𝑡𝑒1 1.2 × 10−3 𝑀/𝑠 1 𝑘(0.10𝑀)𝑥 (0.010𝑀)𝑦


= = =
𝑟𝑎𝑡𝑒2 4.8 × 10−3 𝑀/𝑠 4 𝑘(0.10𝑀)𝑥 (0.040𝑀)𝑦

Cancelling identical terms in the numerator and denominator gives

(0.010𝑀) 𝑦 1 𝑦 1
= ( ) =
(0.040𝑀) 𝑦 4 4

Therefore, y = 1. The reaction is first order in ClO2

Dividing rate from experiment 1 by rate from experiment 3, we get

𝑟𝑎𝑡𝑒1 1.2 × 10−3 𝑀/𝑠 1 𝑘(0.10𝑀)𝑥 (0.010𝑀)𝑦


= = =
𝑟𝑎𝑡𝑒3 2.4 × 10−3 𝑀/𝑠 2 𝑘(0.20𝑀)𝑥 (0.010𝑀)𝑦

Some content reproduced from:


1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
14

Cancelling identical terms in the numerator and denominator gives

(0.10𝑀)𝑥 1 𝑥 1
= ( ) =
(0.20𝑀)𝑥 2 2

Therefore, x = 1. The reaction is first order in F2

𝑟𝑎𝑡𝑒 = 𝑘[F2 ]1 [ClO2 ]1

We can use data from any of the experiments to calculate the value and units of k. Using
data from experiment 1 gives
rate 1.2 × 10−3 M/s
𝑘= = = 1.2 M −1 s−1
[F2 ]1 [ClO2 ]1 (0.10M)(0.010M)

Example 2.2.5: Determining a Rate Law from Initial Rates


A study of the gas-phase reduction of nitric oxide by hydrogen

2NO(g) + 2H2(g) → N2(g) + 2 H2O(g)

yielded the following initial-rate data (all pressures in torr):

Experiment P(NO) P(H2) Initial rate (torr s−1)


1 300 300 1.03
2 150 300 0.25
3 300 600 2.00

Find the order of the reaction with respect to each component.

Solution
In looking over this data, take note of the following:
All the data are expressed in pressures, rather than in concentrations. We can do this
because the reactants are gases, whose concentrations are directly proportional to their
partial pressures when T and V are held constant. Since we are only interested in
comparing the ratios of pressures and rates, the units cancel out and don't matter. It is
far easier experimentally to adjust and measure pressures than concentrations.

Experiments 1 and 2: Reduction of the initial partial pressure of NO by a factor of


(300/150) = 2, results in a reduction of the initial rate by a factor of (1.03/0.25) = about
4, so the reaction is second-order in nitric oxide.

Experiments 1 and 3: Increasing the initial partial pressure of hydrogen by a factor of 2


(600/300), causes a similar increase in the initial rate (2.00/1.03), so the reaction is first
order in hydrogen.

The rate law is thus rate = k[NO]2[H2].

Some content reproduced from:


1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
15

2.3. The Change of Concentration with Time (Integrated Rate Laws)


The differential rate law shows how the overall rate of reaction depends on concentration
of reactants. In order to determine the dependence of reactant concentration on reaction
time, the appropriate the differential rate law is integrated.

2.3.1. Zero Order Reaction


A zeroth-order reaction is one whose rate is independent of concentration. The
differential rate law for the zero order reaction A → products is

𝑟𝑎𝑡𝑒 = 𝑘 (2.3.1)

We refer to these reactions as zeroth order because we could also write their rate in a
form such that the exponent of the reactant in the rate law is 0:

𝑑[A]
rate = − = 𝑘[A]0 = 𝑘 (2.3.2)
𝑑𝑡

Because rate is independent of reactant concentration, a graph of the concentration of


any reactant as a function of time is a straight line with a slope of −k. The value of k is
negative because the concentration of the reactant decreases with time. Conversely, a
graph of the concentration of any product as a function of time is a straight line with a
slope of k, a positive value.

Figure 2.3.1 The graph of a zeroth-order reaction. The change in concentration of reactant and product
with time produces a straight line.

Some content reproduced from:


1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
16

The integrated rate law for a zeroth-order reaction also produces a straight line and has
the general form
[A]𝑡 = −𝑘𝑡 + [A]0 (2.3.3)

where [A]0 is the initial concentration of reactant A. Equation 2.3.3 has the form of the
algebraic equation for a straight line,

𝑦 = 𝑚𝑥 + 𝑏 (2.3.4)

with 𝑦 = [A]𝑡 , 𝑚𝑥 = −𝑘𝑡, and 𝑏 = [A]0 .

In a zeroth-order reaction, the rate constant must have the same units as the reaction
rate, typically moles per liter per second.

A zeroth-order reaction that takes place in the human liver is the oxidation of ethanol
(from alcoholic beverages) to acetaldehyde, catalyzed by the enzyme alcohol
dehydrogenase. At high ethanol concentrations, this reaction is also a zeroth-order
reaction. The overall reaction equation is

where NAD+ (nicotinamide adenine dinucleotide) and NADH (reduced nicotinamide


adenine dinucleotide) are the oxidized and reduced forms, respectively, of a species used
by all organisms to transport electrons. When an alcoholic beverage is consumed, the
ethanol is rapidly absorbed into the blood. Its concentration then decreases at a constant
rate until it reaches zero (part (a) in Figure 2.3.2). An average 70 kg person typically
takes about 2.5 h to oxidize the 15 mL of ethanol contained in a single 340 mL can of
beer, a 150 mL glass of wine, or a shot of distilled spirits (such as whiskey or brandy).
The actual rate, however, varies a great deal from person to person, depending on body
size and the amount of alcohol dehydrogenase in the liver. The reaction rate does not
increase if a greater quantity of alcohol is consumed over the same period of time
because the reaction rate is determined only by the amount of enzyme present in the
liver. Contrary to popular belief, the caffeine in coffee is ineffective at catalyzing the
oxidation of ethanol. When the ethanol has been completely oxidized and its
concentration drops to essentially zero, the rate of oxidation also drops rapidly (part (b)
in Figure 2.3.2).

Some content reproduced from:


1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
17

Figure 2.3.2 The Catalyzed Oxidation of Ethanol (a) The concentration of ethanol in human blood decreases
linearly with time, which is typical of a zeroth-order reaction. (b) The rate at which ethanol is oxidized is
constant until the ethanol concentration reaches essentially zero, at which point the reaction rate drops to
zero.

This example illustrates two important points:

• In a zeroth-order reaction, the reaction rate does not depend on the reactant
concentration.
• A linear change in concentration with time is a clear indication of a zeroth-order
reaction.

2.3.2. First-Order Reactions


In a first-order reaction, the reaction rate is directly proportional to the concentration of
one of the reactants. First-order reactions often have the general form A → products.
The differential rate for a first-order reaction is as follows:

𝑑[A]
rate = − = 𝑘[A] (2.3.5)
𝑑𝑡

If the concentration of A is doubled, the reaction rate doubles; if the concentration of A


is increased by a factor of 10, the reaction rate increases by a factor of 10, and so forth.
Because the units of the reaction rate are always moles per liter per second, the units of
a first-order rate constant are reciprocal seconds (s−1).

The integrated rate law for a first-order reaction can be written in two different ways: one
using exponents and one using logarithms. The exponential form is as follows:

Some content reproduced from:


1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
18

[A]t = [A]𝑜 𝑒 −𝑘𝑡 (2.3.6)

where [A]0 is the initial concentration of reactant A at t = 0; k is the rate constant;


and e is the base of the natural logarithms, which has the value 2.718 to three decimal
places. Recall that an integrated rate law gives the relationship between reactant
concentration and time. Equation 2.3.6 predicts that the concentration of A will decrease
in a smooth exponential curve over time. By taking the natural logarithm of each side
of Equation 2.3.6 and rearranging, we obtain an alternative logarithmic expression of the
relationship between the concentration of A and t:

ln[A]t − ln[A]o = −𝑘𝑡 (2.3.7)

Because Equation 2.3.7 has the form of the algebraic equation for a straight line,

𝑦 = 𝑚𝑥 + 𝑏 (2.3.8)

with 𝑦 = ln[A]t , and 𝑏 = ln[A]0 a plot of ln[A]t vs 𝑡 for a first-order reaction should give
a straight line with a slope of −𝑘 and an intercept of ln[A]o . Either the differential rate
law (Equation 2.3.5) or the integrated rate law (Equation 2.3.7) can be used to determine
whether a particular reaction is first order.

Figure 2.3.3 Graphs of a first-order reaction. The expected shapes of the curves for plots of reactant
concentration versus time (top) and the natural logarithm of reactant concentration versus time (bottom)
for a first-order reaction.

Some content reproduced from:


1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
19

2.3.3. Second-Order Reactions


The equations that relate the concentrations of reactants and the rate constant of second-
order reactions can be fairly complicated. To illustrate the point with minimal complexity,
only the simplest second-order reactions will be described here, namely, those whose
rates depend on the concentration of just one reactant. For these types of reactions, the
differential rate law is written as:

𝑑[A]
rate = − = 𝑘[A]2 (2.3.9)
𝑑𝑡

For these second-order reactions, the integrated rate law is:


1 1
= 𝑘𝑡 + (2.3.10)
[A]𝑡 [A]𝑜

where the terms in the equation have their usual meanings as defined earlier.
Because Equation 2.3.10 has the form of an algebraic equation for a straight line, 𝑦 =
𝑚𝑥 + 𝑏, with 𝑦 = 1/[A]𝑡 and 𝑏 = 1/[A]0 , a plot of 1/[A]𝑡 versus 𝑡 for a simple second-
order reaction is a straight line with a slope of 𝑘 and an intercept of 1/[A]0 .

2.3.4. The Half-Life of a Reaction


The half-life of a reaction (t1/2) is the time required for one-half of a given amount of
reactant to be consumed. In each succeeding half-life, half of the remaining concentration
of the reactant is consumed.

Figure 2.3.4 The rate of decomposition of H2O2 in an aqueous solution decreases as the concentration of
H2O2 decreases.

Some content reproduced from:


1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
20

Using the decomposition of hydrogen peroxide (Figure 2.3.4) as an example, we find that
during the first half-life (from 0.00 hours to 6.00 hours), the concentration of
H2O2 decreases from 1.000 M to 0.500 M. During the second half-life (from 6.00 hours
to 12.00 hours), it decreases from 0.500 M to 0.250 M; during the third half-life, it
decreases from 0.250 M to 0.125 M. The concentration of H2O2 decreases by half during
each successive period of 6.00 hours. The decomposition of hydrogen peroxide is a first-
order reaction, and, as can be shown, the half-life of a first-order reaction is independent
of the concentration of the reactant. However, half-lives of reactions with other orders
depend on the concentrations of the reactants.

2.3.4.1. First Order Reaction


An equation relating the half-life of a first-order reaction to its rate constant may be
derived from the integrated rate law as follows:

[A]𝑡 = [A]0 𝑒 −𝑘𝑡 (2.3.11)

Then since [A]𝑡 = 12[A]0 , we then have


1
2
[A]0
ln = −𝑘𝑡1⁄ (2.3.12)
[A]0 2

ln 12 = −𝑘𝑡1⁄ 𝑜𝑟 − ln 2 = −𝑘𝑡1⁄ (2.3.13)


2 2

ln 2 0.693
∴ 𝑡1⁄ = = (2.3.14)
2 𝑘 𝑘

This equation describes an expected inverse relation between the half-life of the reaction
and its rate constant, k. Faster reactions exhibit larger rate constants and correspondingly
shorter half-lives. Slower reactions exhibit smaller rate constants and longer half-lives.

2.3.4.2. Second Order Reaction


Following the same approach as used for first-order reactions, an equation relating the
half-life of a second-order reaction to its rate constant and initial concentration may be
derived from its integrated rate law:

1 1
− = 𝑘𝑡 (2.3.15)
[A]𝑡 [A]𝑜
Some content reproduced from:
1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
21

Then since [A]𝑡 = 12[A]0 , we then have


1 1
− = 𝑘𝑡1⁄ (2.3.16)
1
2
[A]0 [A]𝑜 2

1
∴ 𝑡1⁄ = (2.3.17)
2 𝑘[A]𝑜

For a second-order reaction, 𝑡1⁄ is inversely proportional to the concentration of the


2
reactant, and the half-life increases as the reaction proceeds because the concentration
of reactant decreases. Unlike with first-order reactions, the rate constant of a second-
order reaction cannot be calculated directly from the half-life unless the initial
concentration is known.

2.3.4.3. Zero-Order Reactions


As for other reaction orders, an equation for zero-order half-life may be derived from the
integrated rate law:

[A]𝑡 = −𝑘𝑡 + [A]0 (2.3.18)

Then since at 𝑡1⁄ , [A]𝑡 = 12[A]0 we then have


2

1
2
[A]0 = −𝑘𝑡1⁄ + [A]0 (2.3.19)
2

[A]𝑜
∴ 𝑡1⁄ = (2.3.20)
2 2𝑘

As for all reaction orders, the half-life for a zero-order reaction is inversely proportional
to its rate constant. However, the half-life of a zero-order reaction increases as the initial
concentration increases.

Some content reproduced from:


1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
22

Table 2.3.1 Summary of Rate Laws for Zero-, First-, and Second-Order Reactions
Zero Order First Oder Second Order

Rate law 𝑟𝑎𝑡𝑒 = 𝑘 𝑟𝑎𝑡𝑒 = 𝑘[𝐴] 𝑟𝑎𝑡𝑒 = 𝑘[𝐴]2


Units of rate constant 𝑚𝑜𝑙 ∙ 𝐿−1 ∙ 𝑠 −1 𝑠 −1 𝐿 ∙ 𝑚𝑜𝑙 −1 ∙ 𝑠 −1
Integrated rate law [A]𝑡 = −𝑘𝑡 + [A]0 ln[A]𝑡 = −𝑘𝑡 + ln[A]𝑜 1 1
= 𝑘𝑡 +
[A]𝑡 [A]𝑜
Plot needed for linear [A]𝑡 𝑣𝑠 𝑡 ln[A]𝑡 𝑣𝑠 𝑡 1
𝑣𝑠 𝑡
fit of rate data [A]𝑡
Slope −𝑘 −𝑘 +𝑘
Half Life [A]𝑜 0.693 1
𝑡1⁄ = 𝑡1⁄ = 𝑡1⁄ =
2 𝑘 2 𝑘 2 𝑘[A]𝑜

Example 2.3.1 First-order Half-Life and Rate Constant


SO2Cl2 decomposes by first order kinetics with 𝑘 = 2.81 × 10−3 min−1 at a certain
temperature.
a) Determine the half-life for the reaction.
b) Determine the time needed for the concentration of a SO2Cl2 sample to decrease to
10% of its initial concentration.

Solution
a) The half-life for a first reaction is given by

ln 2 0.693
𝑡1⁄ = = = 247 𝑚𝑖𝑛
2 𝑘 2.81 × 10−3 min−1

b) For a first order reaction

ln[A]t − ln[A]o = −𝑘𝑡

[A]t
ln ( ) = −𝑘𝑡
[A]0

When the SOCl2 sample has decreased to 10% of its initial concentration

10
[A]𝑡 = [A] = 0.1[A]0
100 0

Therefore
0.1[A]0
ln ( ) = −(2.81 × 10−3 min−1 ) × 𝑡
[A]0

ln (0.1)
𝑡= = 819 min
−2.81 × 10−3 𝑚𝑖𝑛−1
Some content reproduced from:
1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
23

Example 2.3.2 First Order Isomerization Reaction


There are two molecules with the formula C3H6. Propene, CH3CH=CH2, is the monomer of
the polymer polypropylene, which is used for indoor-outdoor carpets. Cyclopropane is
used as an anaesthetic. When heated to 499 °C, cyclopropane rearranges (isomerizes)
and forms propene with a rate constant of 5.95×10–4 s–1.

a) What is the half-life of this reaction?


b) What fraction of the cyclopropane remains after 0.75 h at 499 °C?

Solution
The provided rate constant’s unit is 5.95×10–4 s–1, indicating the reaction is first-order,
and so

ln 2 0.693
𝑡1⁄ = = −4 −1
= 1.16 × 103 s
2 𝑘 5.95 × 10 s

The fraction remaining after 0.75 h may be determined from the integrated rate law:
[A]t
ln ( ) = −𝑘𝑡
[A]0

Rearranging this equation to isolate the fraction remaining yields

[A]t
= 𝑒 −𝑘𝑡
[A]0

Converting the time to seconds and substituting values for k and t gives

[A]t −4 −1 )(0.75 h)(60 min)( 60 s )


= 𝑒 −𝑘𝑡 = e−(5.95×10 s 1h 1 min
[A]0

[A]t
= 𝑒 −1.6065
[A]0

[A]t
= 0.20
[A]0

And so, 20% of the reactant remains.

Some content reproduced from:


1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
24

2.4. Dependence of Reaction Rate on Temperature


It is possible to use kinetics studies of a chemical system, such as the effect of changes
in reactant concentrations, to deduce events that occur on a microscopic scale, such as
collisions between individual particles. Such studies have led to the collision model of
chemical kinetics, which is a useful tool for understanding the behavior of reacting
chemical species. According to the collision model, a chemical reaction can occur only
when the reactant molecules, atoms, or ions collide with more than a certain amount of
kinetic energy and in the proper orientation. The collision model explains why, for
example, most collisions between molecules do not result in a chemical reaction. Nitrogen
and oxygen molecules in a single liter of air at room temperature and 1 atm of pressure
collide about 1030 times per second. If every collision produced two molecules of NO, the
atmosphere would have been converted to NO and then NO2 a long time ago. Instead, in
most collisions, the molecules simply bounce off one another without reacting, much as
marbles bounce off each other when they collide. The collision model also explains why
such chemical reactions occur more rapidly at higher temperatures. For example, the
reaction rates of many reactions that occur at room temperature approximately double
with a temperature increase of only 10°C. In this section, we will use the collision model
to analyze this relationship between temperature and reaction rates.

2.4.1. Collision Theory


We should not be surprised that atoms, molecules, or ions must collide before they can
react with each other. Atoms must be close together to form chemical bonds. This simple
premise is the basis for a very powerful theory that explains many observations regarding
chemical kinetics, including factors affecting reaction rates. The collision theory is based
on the following postulates:

• The rate of a reaction is proportional to the rate of reactant collisions:

𝑛𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑐𝑜𝑙𝑙𝑖𝑠𝑖𝑜𝑛𝑠
𝑟𝑎𝑡𝑒 ∝ (2.4.1)
𝑡𝑖𝑚𝑒

• The reacting species must collide in an orientation that allows contact between
the atoms that will become bonded together in the product.

• The collision must occur with adequate energy to permit mutual penetration of
the reacting species’ valence shells so that the electrons can rearrange and form
new bonds (and new chemical species).

Some content reproduced from:


1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
25

We can see the importance of the two physical factors noted in the second and third
postulates, the orientation and energy of collisions, when we consider the reaction of
carbon monoxide with oxygen:

2CO(𝑔) + O2 (𝑔) → 2CO2 (𝑔) (2.4.2)

Carbon monoxide is a pollutant produced by the combustion of hydrocarbon fuels. To


reduce this pollutant, automobiles have catalytic converters that use a catalyst to carry
out this reaction. It is also a side reaction of the combustion of gunpowder that results
in muzzle flash for many firearms. If carbon monoxide and oxygen are present in sufficient
amounts, the reaction will occur at high temperature and pressure.

The first step in the gas-phase reaction between carbon monoxide and oxygen is a
collision between the two molecules:

CO(𝑔) + O2 (𝑔) → CO2 (𝑔) + O(𝑔) (2.4.3)

Although there are many different possible orientations the two molecules can have
relative to each other, consider the two presented in Figure 2.4.1. In the first case, the
oxygen side of the carbon monoxide molecule collides with the oxygen molecule. In the
second case, the carbon side of the carbon monoxide molecule collides with the oxygen
molecule. The second case is clearly more likely to result in the formation of carbon
dioxide, which has a central carbon atom bonded to two oxygen atoms (O=C=O). This is
a rather simple example of how important the orientation of the collision is in terms of
creating the desired product of the reaction.

Figure 2.4.1 Illustrated are two collisions that might take place between carbon monoxide and oxygen
molecules. The orientation of the colliding molecules partially determines whether a reaction between the
two molecules will occur.

Some content reproduced from:


1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
26

If the collision does take place with the correct orientation, there is still no guarantee that
the reaction will proceed to form carbon dioxide. In addition to a proper orientation, the
collision must also occur with sufficient energy to result in product formation. When
reactant species collide with both proper orientation and adequate energy, they combine
to form an unstable species called an activated complex or a transition state. These
species are very short lived and usually undetectable by most analytical instruments. In
some cases, sophisticated spectral measurements have been used to observe transition
states.
Collision theory explains why most reaction rates increase as concentrations increase.
With an increase in the concentration of any reacting substance, the chances for collisions
between molecules are increased because there are more molecules per unit of volume.
More collisions mean a faster reaction rate, assuming the energy of the collisions is
adequate.

2.4.2. Activation Energy and the Arrhenius Equation


The minimum energy necessary to form a product during a collision between reactants is
called the activation energy (Ea). How this energy compares to the kinetic energy provided
by colliding reactant molecules is a primary factor affecting the rate of a chemical reaction.
If the activation energy is much larger than the average kinetic energy of the molecules,
the reaction will occur slowly since only a few fast-moving molecules will have enough
energy to react. If the activation energy is much smaller than the average kinetic energy
of the molecules, a large fraction of molecules will be adequately energetic, and the
reaction will proceed rapidly.

Figure 2.4.2 shows how the energy of a chemical system changes as it undergoes a
reaction converting reactants to products according to the equation

A+B → C+D (2.4.4)

These reaction diagrams are widely used in chemical kinetics to illustrate various
properties of the reaction of interest. Viewing the diagram from left to right, the system
initially comprises reactants only, A + B. Reactant molecules with sufficient energy can
collide to form a high-energy activated complex or transition state. The unstable
transition state can then subsequently decay to yield stable products, C + D. The diagram
depicts the reaction's activation energy, Ea, as the energy difference between the
reactants and the transition state. Using a specific energy, the enthalpy (see chapter on
thermochemistry), the enthalpy change of the reaction, ΔH, is estimated as the energy
difference between the reactants and products. In this case, the reaction is exothermic
(ΔH < 0) since it yields a decrease in system enthalpy.

Some content reproduced from:


1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
27

Figure 2.4.2 Reaction diagram for the exothermic reaction A + B → C + D

The Arrhenius equation relates the activation energy and the rate constant, k, for many
chemical reactions:

𝐸𝑎⁄
𝑘 = 𝐴𝑒 − 𝑅𝑇 (2.4.5)

In this equation, R is the ideal gas constant, which has a value 8.314 J/mol/K, T is
temperature on the Kelvin scale, Ea is the activation energy in joules per mole, e is the
constant 2.7183, and A is a constant called the frequency factor, which is related to the
frequency of collisions and the orientation of the reacting molecules.

Postulates of collision theory are nicely accommodated by the Arrhenius equation. The
frequency factor, A, reflects how well the reaction conditions favour properly oriented
collisions between reactant molecules. An increased probability of effectively oriented
collisions results in larger values for A and faster reaction rates.

𝐸𝑎
The exponential term, 𝑒 − ⁄𝑅𝑇 , describes the effect of activation energy on reaction rate.
According to kinetic molecular theory (see chapter on gases), the temperature of matter
is a measure of the average kinetic energy of its constituent atoms or molecules. The
distribution of energies among the molecules composing a sample of matter at any given
temperature is described by the plot shown in Figure 2.4.3(a). Two shaded areas under
the curve represent the numbers of molecules possessing adequate energy (RT) to
overcome the activation barriers (Ea). A lower activation energy results in a greater
fraction of adequately energized molecules and a faster reaction.

Some content reproduced from:


1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
28

The exponential term also describes the effect of temperature on reaction rate. A higher
temperature represents a correspondingly greater fraction of molecules possessing
sufficient energy (RT) to overcome the activation barrier (Ea), as shown in Figure 2.4.3(b).
This yields a greater value for the rate constant and a correspondingly faster reaction
rate.

Figure 2.4.3 Molecular energy distributions showing numbers of molecules with energies exceeding (a)
two different activation energies at a given temperature, and (b) a given activation energy at two different
temperatures.

A convenient approach for determining Ea for a reaction involves the measurement of k at


two or more different temperatures and using an alternate version of the Arrhenius
equation that takes the form of a linear equation

𝐸𝑎
ln 𝑘 = − + ln 𝐴 (2.4.6)
𝑅𝑇
𝑦 = 𝑚𝑥 + 𝑏

When k is determined experimentally at several temperatures, 𝐸𝑎 can be calculated from


𝐸
the slope of a plot of ln 𝑘 vs. 1⁄𝑇. Slope = − 𝑅𝑎

At two temperatures where 𝑇2 > 𝑇1

𝑘2 𝐸𝑎 1 1
ln = ( − ) (2.4.7)
𝑘1 𝑅 𝑇1 𝑇2

𝑘2 𝐸𝑎 𝑇2 − 𝑇1
ln = ( ) (2.4.8)
𝑘1 𝑅 𝑇1 𝑇2
Some content reproduced from:
1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
29

Example 2.4.1 Arrhenius Equation


The rate constant of the reaction

O(g) + N2(g) → NO(g) + N(g)

which occurs in the stratosphere, is 9.7 1010 L mol−1 s−1 at 800 C. The activation
energy of the reaction is 315 kJ/mol. Determine the rate constant at 700 C.

Solution
𝑘2 𝐸𝑎 1 1
ln = ( − )
𝑘1 𝑅 𝑇1 𝑇2

And
𝑘2
𝑅 ln
𝑘1
𝐸𝑎 =
1 1
(𝑇 − 𝑇 )
1 2

But
𝑘1 = 0.0796 L mol−1 s−1 when 𝑇1 = 1010 K
𝑘2 = 0.0815 L mol−1 s−1 when 𝑇2 = 1220K

Then
0.0815 L mol−1 s −1
8.314 J K −1 mol−1 ln ( )
0.0796 L mol−1 s −1
𝐸𝑎 =
1 1
( − )
1010 K 1220 K

𝐸𝑎 = 1150 J mol−1

Example 2.4.2 Relative rates from the Arrhenius equation


The activation energy of a certain reaction is 65.7 kJ/mol. How many times faster will the
reaction occur at 50 C than at 0 C?

Solution
Let k2 be the rate constant at 50 C (323 K) and k1 the rate constant at C (273 K).

𝑟𝑎𝑡𝑒2 𝑘2
=
𝑟𝑎𝑡𝑒1 𝑘1

But
𝑘2 𝐸𝑎 1 1
ln = ( − )
𝑘1 𝑅 𝑇1 𝑇2

Some content reproduced from:


1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward
30

kJ 103 J
𝑘2 65.7 ×( ) 1 1
mol kJ
ln = ( − ) = 4.48
𝑘1 J 273 𝐾 323 𝐾
8.314
mol

𝑘2
= 𝑒 4.48 = 88
𝑘1

𝑘2 = 88𝑘1

The reaction will occur 88 times faster at 50 C than at 0 C.

2.5. Catalysis
Catalysts are substances that increase the reaction rate of a chemical reaction without
being consumed in the process. A catalyst, therefore, does not appear in the overall
stoichiometry of the reaction it catalyzes, but it must appear in at least one of the
elementary reactions in the mechanism for the catalyzed reaction. The catalyzed pathway
has a lower Ea, but the net change in energy that results from the reaction (the difference
between the energy of the reactants and the energy of the products) is not affected by
the presence of a catalyst (Figure 2.4.4). Nevertheless, because of its lower Ea, the
reaction rate of a catalyzed reaction is faster than the reaction rate of the uncatalyzed
reaction at the same temperature.

Figure 2.4.4 Reaction diagrams for an exothermic process in the absence and presence (blue curve) of a
catalyst.

Some content reproduced from:


1. OpenStax Chemistry 2e: Access for free at https://openstax.org/books/chemistry-2e/pages/1-introduction
2. A general chemistry Libretexts Textmap organized around the textbook Chemistry: The Central Science by Brown, LeMay, Busten,
Murphy, and Woodward

You might also like