Additional Reading Materials

Chapter 10: Reaction Mechanisms

A balanced equation for a chemical reaction indicates what is reacting and what is produced, but it reveals nothing about how the reaction actually takes place. The reaction mechanism (or reaction path) is the process, or pathway, by which a reaction occurs.

Ch10.1 Elementary Reaction

A chemical reaction usually occurs in steps, although it may not always be obvious to an observer. The decomposition of ozone, for example, appears to follow a mechanism with two steps:

[latex]\begin{array}{r @{{}\longrightarrow{}} l} \text{O}_3(g) & \text{O}_2(g)\;+\;\text{O} \\[0.5em] \text{O}\;+\;\text{O}_3(g) & 2\text{O}_2(g) \end{array}[/latex]

We call each step in a reaction mechanism an elementary reaction. Elementary reactions occur exactly as they are written and cannot be broken down into simpler steps. Elementary reactions add up to the overall reaction, which, for the decomposition, is:

[latex]2\text{O}_3(g)\;{\longrightarrow}\;3\text{O}_2(g)[/latex]

Notice that the oxygen atom produced in the first step of this mechanism is consumed in the second step and therefore does not appear as a product in the overall reaction. Species that are produced in one step and consumed in a subsequent step are called intermediates.

While the overall reaction equation for the decomposition of ozone indicates that two molecules of ozone react to give three molecules of oxygen, the mechanism of the reaction does not involve the collision and reaction of two ozone molecules. Rather, it involves a molecule of ozone decomposing to an oxygen molecule and an intermediate oxygen atom; the oxygen atom then reacts with a second ozone molecule to give two oxygen molecules. These two elementary reactions occur exactly as they are shown in the reaction mechanism.

Ch10.2 Unimolecular Elementary Reactions

The molecularity of an elementary reaction is the number of reactant species (atoms, molecules, or ions). For example, a unimolecular reaction involves the rearrangement of a single reactant species to produce one or more molecules of product:

A → products

The rate equation for a unimolecular reaction is:

rate = k[A]

A unimolecular reaction may be one of several elementary reactions in a complex mechanism. For example, the reaction:

O3 → O2 + O

illustrates a unimolecular elementary reaction that occurs as one step of a two-step reaction mechanism. An elementary reaction can also be an overall reaction in some cases. For example, the gas-phase decomposition of cyclobutane, C4H8, to ethylene, C2H4, occurs via a unimolecular, single-step mechanism:
In this figure, structural formulas are used to illustrate a chemical reaction. On the left, a structural formula for cyclobutane is shown. This structure is composed of 4 C atoms connected with single bonds in a square shape. Each C atom is bonded to two other C atoms in the structure, leaving two bonds for H atoms pointing outward above, below, left, and right. An arrow points right to two identical ethane molecules with a plus symbol between them. Each of these molecules contains two C atoms connected with a double bond oriented vertically between them. The C atom at the top of these molecules has H atoms bonded above to the right and left. Similarly, the lower C atom has two H atoms bonded below to the right and left.

For these unimolecular reactions to occur, all that is required is the separation of parts of single reactant molecules into products.

Chemical bonds do not simply fall apart during chemical reactions. Energy is required to break chemical bonds. The activation energy for the decomposition of C4H8, for example, is 261 kJ/mol. This means that it requires 261 kJ to distort one mole of these molecules into activated complexes that then decompose into products:
In this figure, structural formulas are used to illustrate a chemical reaction, including an intermediate step. On the left, a structural formula for cyclobutane is shown. This structure is composed of 4 C atoms connected with single bonds in a square shape. Each C atom is bonded to two other C atoms in the structure, leaving two bonds for H atoms pointing outward above, below, left, and right. This structure is labeled, “Cyclohexane.” An arrow points right to a similar structure which has the upper and lower bonds replaced by rows of 4 dots. Similarly, columns of 3 dots appear just inside the line segments indicating the vertically oriented single bonds in the structure. The label “Activated complex” appears beneath this structure. A second arrow points right to two identical ethane molecules with a plus symbol between them. Each of these molecules contains two C atoms connected with a double bond oriented vertically between them. The C atom at the top of these molecules has H atoms bonded above to the right and left. Similarly, the lower C atom has two H atoms bonded below to the right and left. Below these two molecules appears the label “Ethylene.”

In a sample of C4H8, a rapidly moving C4H8 molecule can collide with other rapidly moving molecules and pick up additional energies. When this C4H8 molecule gains enough energy, it can transform into an activated complex that leads to the formation of ethylene molecules. In effect, energetic collisions can knock a C4H8 molecule into the geometry of the activated complex. However, depending on the temperature, only some fraction of gas molecules travel at sufficiently high speeds with large enough kinetic energies to accomplish this. Hence, at any given moment, only a few molecules pick up enough energy from collisions to react.

The rate of decomposition of C4H8 is directly proportional to its concentration. Doubling the concentration of C4H8 in a sample gives twice as many molecules per liter. Although the fraction of molecules with enough energy to react remains the same, the total number of such molecules is twice as great, and the reaction rate is twice as fast:

[latex]\text{rate} = -\frac{{\Delta}[\text{C}_4\text{H}_8]}{{\Delta}t} = k[\text{C}_4\text{H}_8][/latex]

A similar relationship applies to any unimolecular elementary reaction; the reaction rate is directly proportional to the concentration of the reactant, and the reaction exhibits first-order behavior. The proportionality constant is the rate constant for the particular unimolecular reaction.

Ch10.3 Bimolecular Elementary Reactions

The collision and combination of two molecules or atoms to form an activated complex in an elementary reaction is called a bimolecular reaction. There are two types of bimolecular elementary reactions:

A + B → products

where the two reactant molecules are different, and the rate law is first-order in A and first order in B:

rate = k[A][B]

and

2A → products

in which two identical molecules collide and react, the rate law is second order in A:

rate = k[A]2

Some chemical reactions have mechanisms that consist of a single bimolecular elementary reaction. One example is the reaction of nitrogen dioxide with carbon monoxide:

NO2(g) + CO(g) → NO(g) + CO2(g)

Another is the decomposition of two hydrogen iodide molecules to produce hydrogen, H2, and iodine, I2, Figure 1:

2HI(g) → H2(g) + I2(g)

This figure provides an illustration of a reaction between two H I molecules using space filling models. H atoms are shown as white spheres, and I atoms are shown as purple spheres. On the left, two H I molecules are shownwith a small white sphere bonded to a much larger purple sphere. The label, “Two H I molecules,” appears below. An arrow points right to a similar structure in which the two molecules appear pushed together, so that the purple spheres of the two molecules are touching. Below appears the label, “Transition state.” Following another arrow, two white spheres are shown vertically oriented and bonded together with the label, “H subscript 2” above. The H subscript 2 molecule is followed by a plus sign and two purple spheres bonded together with the label, “I subscript 2” above. Below these structures is the label, “Hydrogen iodide molecules decompose to produce hydrogen H subscript 2 and iodine I subscript 2.”
Figure 1. The probable mechanism for the reaction of two HI molecules to produce one molecule of H2 and one molecule of I2.

Bimolecular elementary reactions may also be involved as steps in a multistep reaction mechanism. For example, the reaction of atomic oxygen with ozone is one of the steps in the multistep ozone decomposition reaction:

O(g) + O3(g) → 2O2(g)

Ch10.4 Termolecular Elementary Reactions

An elementary termolecular reaction involves the simultaneous collision of three atoms or molecules. Termolecular elementary reactions rarely occur because the probability of three particles colliding simultaneously is less than 0.1% of the probability of two particles colliding. There are, however, a few established termolecular elementary reactions. The reaction mechanism of nitric oxide with oxygen appears to involve a termolecular step:

2NO(g) + O2(g) → 2NO2(g)

rate = k[NO]2[O2]

Likewise, the reaction of nitric oxide with chlorine appears to involve a termolecular step:

2NO(g) + Cl2(g) → 2NOCl(g)

rate = k[NO]2[Cl2]

Ch10.5 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.

Collision theory is based on the following postulates:

  1. The rate of a reaction is proportional to the rate of reactant collisions:
    [latex]\text{reaction\;rate}\;{\propto}\;\frac{\#\;\text{collisions}}{\text{time}}[/latex]
  2. The reacting species must collide in an orientation that allows contact between the atoms that will become bonded together in the product.
  3. The collision must occur with adequate energy so that the electrons can rearrange and form new bonds (and new chemical species).

We can see the importance of postulates 2 when we consider the reaction of carbon monoxide with oxygen:

2CO(g) + O2(g) → 2CO2(g)

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 quantity, the reaction is spontaneous at high temperature and pressure.

The first step in this reaction is a collision between two molecules:

CO(g) + O2(g) → CO2(g) + O(g)

Although there are many different possible orientations the two molecules can have relative to each other, consider the two presented in Figure 2: In case (1), the oxygen side of the CO molecule collides with the O2 molecule; in case (2), the carbon side of the CO molecule collides with O2. Case (2) is clearly more likely to result in the formation of a CO2 molecule, which has a central carbon atom bonded to two oxygen atoms (O=C=O). This is a simple example of how important molecular orientations are in terms of creating the desired product.

A diagram is shown that illustrates two possible collisions between C O and O subscript 2. In the diagram, oxygen atoms are represented as red spheres and carbon atoms are represented as black spheres. The diagram is divided into upper and lower halves by a horizontal dashed line. At the top left, a C O molecule is shown striking an O subscript 2 molecule such that the O atom from the C O molecule is at the point of collision. Surrounding this collision are a mix of molecules of C O, and O subscript 2 of varying sizes. At the top middle region of the figure, two separated O atoms are shown as red spheres with the label, “Oxygen to oxygen,” beneath them. To the upper right, “No reaction” is written. Similarly in the lower left of the diagram, a C O molecule is shown striking an O subscript 2 molecule such that the C atom from the C O molecule is at the point of collision. Surrounding this collision are a mix of molecules of C O, and O subscript 2 of varying sizes. At the lower middle region of the figure, a black sphere and a red spheres are shown with the label, “Carbon to oxygen,” beneath them. To the lower right, “More C O subscript 2 formation” is written and three models of C O subscript 2 composed each of a single central black sphere and two red spheres in a linear arrangement are shown.
Figure 2. 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.

Collision theory also 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.

Ch10.6 Transition State

We used to example of the first step in the gas-phase reaction between CO and O2 (2CO(g) + O2(g) → 2CO2(g)) to illustrate the importance of the orientation of the collision in terms of creating the desired product.

CO(g) + O2(g) → CO2(g) + O(g)

However, even if the collision does take place with the proper orientation, there is still no guarantee that this bimolecular reaction step will proceed to form CO2. Every reaction requires a certain amount of activation energy for it to proceed in the forward direction, yielding an appropriate activated complex along the way. As Figure 3 demonstrates, even a collision with the correct orientation can fail to form the reaction product. In the study of reaction mechanisms, each of these three arrangements of atoms is called a proposed activated complex or transition state.

This figure shows three rows of structures. In the first row, an O atom on the left is connected to a C atom on its right with a double bond indicated by a pair of short parallel line segments. To the right of the C atom are three dots in a horizontal row followed by an O atom double bonded to another O atom on its right. In the second row, an O atom is followed by three dots in a horizontal row, which are followed by a C atom and a second grouping of three dots. To the right is an O atom double bonded to another O atom. In the third row, an O atom on the left is connected to a C atom on its right with a double bond indicated by a pair of short parallel line segments. To the right of the C atom are three dots in a horizontal row followed by an O atom followed by another grouping of three dots to another O atom on its right.
Figure 3. Possible transition states (activated complexes) for CO reacting with O2. Solid lines represent covalent bonds, while dotted lines represent unstable orbital overlaps that may, or may not, become covalent bonds as the reaction proceeds. In the first two examples, the O=O double bond is not impacted; therefore, carbon dioxide cannot form. The third proposed transition state will result in the formation of carbon dioxide if the third “extra” oxygen atom separates from the rest of the molecule.

In most circumstances, it is impossible to isolate or identify a transition state or activated complex. For the above example, activated complexes have only been observed spectroscopically in systems that utilize a heterogeneous catalyst. The gas-phase reaction occurs too rapidly to isolate any such chemical compound.

Ch10.7 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). The kinetic energy of reactant molecules plays an important role in a reaction because the energy necessary to form a product is provided by a collision of a reactant molecule with another reactant molecule. (In single-reactant reactions, activation energy may be provided by a collision of the reactant molecule with the wall of the reaction vessel or with inert molecules.) If the activation energy is much larger than the average kinetic energy of the molecules, the reaction will occur slowly—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, the fraction of molecules possessing the necessary kinetic energy will be large—most collisions between molecules will result in reaction, and the reaction will occur rapidly.

Figure 4 shows the energy relationships for the general reaction:

A + B → C + D

The energy of the transition state is higher than that of the reactants A and B by an amount equal to Ea, the activation energy. Thus, the sum of the kinetic energies of an A molecule and a B molecule must be equal to or greater than Ea to reach the transition state. After the transition state has been reached, and as C and D begin to form, the system releases energy until its total energy is lower than that of the initial mixture. This lost energy is transferred to other molecules, giving them enough energy to reach the transition state. In Figure 4, ΔH represents the difference in enthalpy between the reactants (A and B) and the products (C and D). Hence, the sum of Ea and ΔH represents the activation energy for the reverse reaction:

C + D → A + B

A graph is shown with the label, “Extent of reaction,” bon the x-axis and the label, “Potential energy,” on the y-axis. Above the x-axis, a portion of a dashed curve is labeled “A plus B.” From the right end of this region, the concave down curve continues upward to reach a maximum near the height of the y-axis. The peak of this curve is labeled, “Transition state.” A double sided arrow extends from a dashed horizontal line that originates at the y-axis at a common endpoint with the curve to the peak of the curve. This arrow is labeled “E subscript a.” A second horizontal dashed line segment is drawn from the right end of the black curve left to the vertical axis at a level significantly lower than the initial “A plus B” labeled end of the curve. The end of the curve that is shared with this segment is labeled, “C plus D.” The curve, which was initially dashed, continues as a solid curve from the maximum to its endpoint at the right side of the diagram. A second double sided arrow is shown. This arrow extends between the two dashed horizontal lines and is labeled, “capital delta H.”
Figure 4. This graph shows the potential energy relationships for the reaction A + BC + D. The dashed portion of the curve represents the energy of the system with a molecule of A and a molecule of B, and the solid portion the energy of the system with a molecule of C and a molecule of D. The activation energy for the forward reaction is represented by Ea. The activation energy for the reverse reaction is greater than that for the forward reaction by an amount equal to ΔH. The curve’s maximum represents the transition state.

We can use the Arrhenius equation to relate the activation energy and the rate constant, k, of a given reaction:

[latex]k = Ae^{-E_a/RT}[/latex]

In this equation, R is the ideal gas constant, which has a value of 8.314 [latex]\frac{J}{mol*K}[/latex], T is temperature on the Kelvin scale, and Ea is the activation energy in J/mol. The constant A is the frequency factor, or prefactor, which is related to the rate at which collisions having the correct orientation occur. The overall exponential term, [latex]e^{-E_a/RT}[/latex], is related to the fraction of collisions providing adequate energy to overcome the activation barrier of the reaction.

At one extreme, the system does not contain enough energy for collisions to overcome the activation barrier. In such cases, no reaction occurs. At the other extreme, the system has so much energy that every collision with the correct orientation can overcome the activation barrier, causing the reaction to proceed nearly instantaneously.

The Arrhenius equation describes quantitatively much of what we have already discussed about reaction rates. For two reactions at the same temperature, the reaction with the higher activation energy has the lower rate constant and hence the slower rate. (A larger Ea results in a smaller value for [latex]e^{-E_a/RT}[/latex], reflecting the smaller fraction of molecules with enough energy to react, which leads to a smaller rate constant.) The rate constant is also directly proportional to the frequency factor, A. Hence a change in conditions or reactants that increases the number of collisions with a favorable orientation for reaction results in an increase in A and, consequently, an increase in k.

Ch10.8 Effect of Temperature

In a gas sample, individual molecules have widely varying speeds; however, because of the vast number of molecules and collisions involved, the molecular speed distribution and average speed are constant. This molecular speed distribution is known as a Maxwell-Boltzmann distribution, and it depicts the relative numbers of molecules in a bulk sample of gas that possesses a given speed (Figure 5).

A graph is shown. The horizontal axis is labeled, “Velocity v ( m divided by s ).” This axis is marked by increments of 20 beginning at 0 and extending up to 120. The vertical axis is labeled, “Fraction of molecules.” A positively or right-skewed curve is shown in red which begins at the origin and approaches the horizontal axis around 120 m per s. At the peak of the curve, a point is indicated with a black dot and is labeled, “v subscript p.” A vertical dashed line extends from this point to the horizontal axis at which point the intersection is labeled, “v subscript p.” Slightly to the right of the peak a second black dot is placed on the curve. This point is labeled, “v subscript r m s.” A vertical dashed line extends from this point to the horizontal axis at which point the intersection is labeled, “v subscript r m s.” The label, “O subscript 2 at T equals 300 K” appears in the open space to the right of the curve.
Figure 5. The molecular speed distribution for oxygen gas at 300 K is shown here. Very few molecules move at either very low or very high speeds. The number of molecules with intermediate speeds increases rapidly up to a maximum, which is the most probable speed, then drops off rapidly. Note that the most probable speed, νp, is a little less than 400 m/s, while the root mean square speed, urms, is closer to 500 m/s.

The average kinetic energy, KEavg, is then equal to:

[latex]\text{KE}_{\text{avg}} = \frac{1}{2}mu^2_{\text{rms}}[/latex]

The KEavg of a collection of gas molecules is also directly proportional to the temperature of the gas and may be described by the equation:

[latex]\text{KE}_{\text{avg}} = \frac{3}{2}RT[/latex]

where R is the gas constant and T is the kelvin temperature. When used in this equation, the appropriate form of the gas constant is 8.314 J/K (8.314 kg m2s–2K–1).

If the temperature of a gas increases, its KEavg increases, more molecules have higher speeds and fewer molecules have lower speeds, and the distribution shifts toward higher speeds overall, that is, to the right. If temperature decreases, KEavg decreases, more molecules have lower speeds and fewer molecules have higher speeds, and the distribution shifts toward lower speeds overall, that is, to the left. This behavior is illustrated for nitrogen gas in Figure 6.

A graph with four positively or right-skewed curves of varying heights is shown. The horizontal axis is labeled, “Velocity v ( m divided by s ).” This axis is marked by increments of 500 beginning at 0 and extending up to 1500. The vertical axis is labeled, “Fraction of molecules.” The label, “N subscript 2,” appears in the open space in the upper right area of the graph. The tallest and narrowest of these curves is labeled, “100 K.” Its right end appears to touch the horizontal axis around 700 m per s. It is followed by a slightly wider curve which is labeled, “200 K,” that is about three quarters of the height of the initial curve. Its right end appears to touch the horizontal axis around 850 m per s. The third curve is significantly wider and only about half the height of the initial curve. It is labeled, “500 K.” Its right end appears to touch the horizontal axis around 1450 m per s. The final curve is only about one third the height of the initial curve. It is much wider than the others, so much so that its right end has not yet reached the horizontal axis. This curve is labeled, “1000 K.”
Figure 6. The molecular speed distribution for nitrogen gas (N2) shifts to the right and flattens as the temperature increases; it shifts to the left and heightens as the temperature decreases.

At a given temperature, all gases have the same KEavg for their molecules. Gases composed of lighter molecules have more high-speed particles and a higher urms, with a speed distribution that peaks at relatively higher velocities. Gases consisting of heavier molecules have more low-speed particles, a lower urms, and a speed distribution that peaks at relatively lower velocities. This trend is demonstrated by the data for a series of noble gases shown in Figure 7.

A graph is shown with four positively or right-skewed curves of varying heights. The horizontal axis is labeled, “Velocity v ( m divided by s ).” This axis is marked by increments of 500 beginning at 0 and extending up to 3000. The vertical axis is labeled, “Fraction of molecules.” The tallest and narrowest of these curves is labeled, “X e.” Its right end appears to touch the horizontal axis around 600 m per s. It is followed by a slightly wider curve which is labeled, “A r,” that is about half the height of the initial curve. Its right end appears to touch the horizontal axis around 900 m per s. The third curve is significantly wider and just over a third of the height of the initial curve. It is labeled, “N e.” Its right end appears to touch the horizontal axis around 1200 m per s. The final curve is only about one fourth the height of the initial curve. It is much wider than the others, so much so that its right reaches the horizontal axis around 2500 m per s. This curve is labeled, “H e.”
Figure 7. Molecular velocity is directly related to molecular mass. At a given temperature, lighter molecules move faster on average than heavier molecules.

An increase in temperature has a similar effect as a decrease in activation energy, as shown in Figure 8. At a given temperature, a sample of molecules have a distribution of kinetic energies described by the Maxwell-Boltzmann distribution, and some fraction of the molecules would have kinetic energies higher than the reaction’s Ea. Therefore, at a higher temperature, a larger fraction of molecules would have the necessary energy to react, as indicated by an increase in the value of [latex]e^{-E_a/RT}[/latex].

Two graphs are shown each with an x-axis label of “Kinetic energy” and a y-axis label of “Fraction of molecules.” Each contains a positively skewed curve indicated in red that begins at the origin and approaches the x-axis at the right side of the graph. In a, a small area under the far right end of the curve is shaded orange. An arrow points down from above the curve to the left end of this region where the shading begins. This arrow is labeled, “Higher activation energy, E subscript a.” In b, the same red curve appears, and a second curve is drawn in black. It is also positively skewed, but reaches a lower maximum value and takes on a broadened appearance as compared to the curve in red. In this graph, the red curve is labeled, “T subscript 1” and the black curve is labeled, “T subscript 2.” In the open space at the upper right on the graph is the label, “T subscript 1 less than T subscript 2.” As with the first graph, the region under the curves at the far right is shaded orange and a downward arrow labeled “E subscript a” points to the left end of this shaded region.
Figure 8. (a) As the activation energy of a reaction decreases, the number of molecules with at least this much energy increases, as shown by the shaded areas. (b) At a higher temperature, T2, more molecules have kinetic energies greater than Ea, as shown by the yellow shaded area.

Ch10.9 Arrhenius Plot

A convenient approach to determining Ea for a reaction involves the measurement of k at different temperatures and using an alternate version of the Arrhenius equation that takes the form of linear equation:

[latex]\begin{array}{r @{{}={}} l} \text{ln}(k) & (\frac{-E_{\text{a}}}{R})(\frac{1}{T})\;+\;\text{ln}(A) \\[0.5em] y & mx\;+\;b \end{array}[/latex]

Thus, a plot of ln(k) versus [latex]\frac{1}{T}[/latex] gives a straight line where [latex]slope=\frac{-E_a}{R}[/latex], from which Ea may be determined. The intercept gives the value of ln(A).

Example 1

Determination of Ea
The variation of the rate constant with temperature for the decomposition of HI:

2HI(g) → H2(g) + I2(g)

is given here. What is the activation energy for the reaction?

T (K) k (M-1s-1)
555 3.52 × 10−7
575 1.22 × 10−6
645 8.59 × 10−5
700 1.16 × 10−3
781 3.95 × 10−2

Solution
Values of [latex]\frac{1}{T}[/latex] and ln(k) are:

[latex]\frac{1}{T}\;(K^{-1})[/latex] ln k
1.80 × 10−3 −14.860
1.74 × 10−3 −13.617
1.55 × 10−3 −9.362
1.43 × 10−3 −6.759
1.28 × 10−3 −3.231

Figure 9 is a graph of ln(k) versus [latex]\frac{1}{T}[/latex]. A linear fit of the data gives:

slope = -2.23×104 K.

A graph is shown with the label “1 divided by T ( K superscript negative 1 )” on the x-axis and “l n k” on the y-axis. The horizontal axis has markings at 1.4 times 10 superscript 3, 1.6 times 10 superscript 3, and 1.8 times 10 superscript 3. The y-axis shows markings at intervals of 2 from negative 14 through negative 2. A decreasing linear trend line is drawn through five points at the coordinates: (1.28 times 10 superscript negative 3, negative 3.231), (1.43 times 10 superscript negative 3, negative 6.759), (1.55 times 10 superscript negative 3, negative 9.362), (1.74 times 10 superscript negative 3, negative 13.617), and (1.80 times 10 superscript negative 3, negative 14.860). A vertical dashed line is drawn from a point just left of the data point nearest the y-axis. Similarly, a horizontal dashed line is draw from a point just above the data point closest to the x-axis. These dashed lines intersect to form a right triangle with a vertical leg label of “capital delta l n k” and a horizontal leg label of “capital delta 1 divided by T.”
Figure 9. This graph shows the linear relationship between ln(k) and 1/T for the reaction 2HI⟶H2 + I2.

To determine the activation energy:

[latex]E_{\text{a}} = -\text{slope}\;\times\;R = -(-2.23\;\times\;10^4\;\text{K}\;\times\;8.314\;\frac{\text{J}}{\text{mol*K}})[/latex]
[latex]E_{\text{a}} = 1.85\;\times\;10^5\;\frac{\text{J}}{\text{mol}}[/latex]

 

It is possible to obtain a reasonable estimate of the activation energy without going through the entire process of constructing the Arrhenius plot. The slope of the plot is

[latex]\text{Slope} = \frac{{\Delta}(\text{ln}\;k)}{{\Delta}(\frac{1}{T})} = -\frac{E_{\text{a}}}{R}[/latex]

The Arrhenius equation:

[latex]\text{ln}(k) = (\frac{-E_{\text{a}}}{R})(\frac{1}{T})\;+\;\text{ln}(A)[/latex]

can therefore take the form (by subtracting the Arrhenius equation with k2 and T2 from the Arrhenius equation with k1 and T1):

[latex]\frac{\text{ln}(k_1)-\text{ln}(k_2)}{(\frac{1}{T_1}-\frac{1}{T_2})} = -\frac{E_{\text{a}}}{R}[/latex]

or

[latex]\text{ln}\;(\frac{k_1}{k_2}) = -\frac{E_{\text{a}}}{R}\;(\frac{1}{T_1}\;-\;\frac{1}{T_2})[/latex]

or

[latex]E_{\text{a}} = -R\;(\frac{\text{ln}\;(\frac{k_1}{k_2})}{(\frac{1}{T_1})\;-\;(\frac{1}{T_2})})[/latex]

Using the experimental data presented here, we can simply select two data entries, for example, the first entry and the last entry:

T (K) k (M-1s-1)
555 3.52 × 10−7
781 3.95 × 10−2

and substitute into the equation:

[latex]E_{\text{a}} = -8.314\;\frac{\text{J}}{\text{mol*K}}\;(\frac{\text{ln}\;(\frac{3.52\;\times\;10^{-7}\;M^{-1}s^{-1}}{3.95\;\times\;10^{-2}\;M^{-1}s^{-1}})}{(\frac{1}{555\;K})\;-\;(\frac{1}{781\;K})})=\;1.86\;\times\;10^{5}\;\frac{J}{mol}[/latex]

This method works well especially when a limited number of temperature-dependent rate constants are available for the reaction of interest.

Check Your Learning
The rate constant for the rate of decomposition of N2O5 to NO and O2 in the gas phase:

2N2O5(g) → 4NO(g) + 3O2(g)

is 1.66 M-1s-1 at 650. K and 7.39 M-1s-1 at 700. K. Assuming the kinetics of this reaction are consistent with the Arrhenius equation, calculate the activation energy for this decomposition.

Answer:

1.13×105 J/mol

Ch10.10 Relating Reaction Mechanisms to Rate Laws

It’s often the case that one step in a multistep reaction mechanism is significantly slower than the others. Because a reaction cannot proceed faster than its slowest step, this step will limit the rate at which the overall reaction occurs. The slowest step is therefore called the rate-determining step (or rate-limiting step) of the reaction Figure 10.

A photo is shown of cattle passing through a narrow chute into a holding pen. A person directs them through the gate with a long white and red pole.
Figure 10. A cattle chute is a nonchemical example of a rate-determining step. Cattle can only be moved from one holding pen to another as quickly as one animal can make its way through the chute. (credit: Loren Kerns)

For elementary reactions, rate laws (and reaction order) can be derived directly from the stoichiometry of the chemical equations. This is not the case, however, for multistep reactions where the balanced equations represent the overall change for the chemical system. In such a case, we must determine the overall rate law from experimental data and deduce the mechanism from the rate law (and sometimes from other data).

Ch10.11 Rate-Determining Step is the First Step in a Multistep Reaction

The reaction of NO2 and CO provides an illustrative example:

NO2(g) + CO(g) → CO2(g) + NO(g)

For temperatures above 225 °C, the rate law has been found to be:

rate = k[NO2][CO]

where the reaction is first order with respect to NO2 and first-order with respect to CO. This is consistent with a single-step bimolecular mechanism and it is possible that this is the mechanism for this reaction at high temperatures.

At temperatures below 225 °C, the reaction is described by a rate law that is second order with respect to NO2:

rate = k[NO2]2

This is consistent with a mechanism that involves the following two elementary reaction steps:

1) NO2(g) + NO2(g) → NO3(g) + NO(g) (slow)

2) NO3(g) + CO(g) → CO2(g) + NO2(g) (fast)

___

NO2(g) + CO(g) → CO2(g) + NO(g) (overall)

The first step is slower and is therefore the rate-determining step, and gives a rate law showing second-order dependence on the NO2 concentration. The sum of the two elementary reaction steps gives the net overall reaction—the NO3 that is formed in step (1) is consumed in step (2), and one of the NO2 consumed in step (1) is reformed in step (2).

When the rate-determining (slower) step is the first step in a mechanism, as is the case for this example, the rate law for the overall reaction is the same as the rate law for this first step.

Ch10.12 Rate-Determining Step is Not the First Step in a Multistep Reaction

When the rate-determining step is preceded by a step involving an equilibrium reaction, the rate law for the overall reaction cannot be simply represented by the rate law of the rate-determining step.

An elementary reaction is at equilibrium when it proceeds in both the forward and reverse directions at equal rates. Consider the dimerization of NO to N2O2:

[latex]\text{NO}\;+\;\text{NO}\;{\rightleftharpoons}\;\text{N}_2\text{O}_2[/latex]

At equilibrium, the rate of the forward reaction is equal to the rate of the reverse reaction:

rateforward = ratereverse

k1[NO]2 = k1[N2O2]

where k1 represents the rate constant of the forward reaction and k-1 represents the rate constant of the reverse reaction.

If N2O2 was an intermediate in a mechanism, this expression can be rearranged to represent the concentration of N2O2 in the overall rate law expression using algebraic manipulation:

[latex](\frac{k_1}{k_{-1}})[\text{NO}]^2 = [\text{N}_2\text{O}_2][/latex]

Intermediates cannot be listed as part of the overall rate law expression, partly because it is extremely difficult to obtain their concentrations. However, they can be included in an individual elementary reaction of a mechanism. Example 1 will illustrate how to derive overall rate laws from mechanisms involving equilibrium steps preceding the rate-determining step.

Example 2

Deriving the Overall Rate Law Expression for a Multistep Reaction Mechanism
Nitryl chloride (NO2Cl) decomposes to nitrogen dioxide (NO2) and chlorine gas (Cl2) according to the following mechanism:

1) [latex]2\text{NO}_2\text{Cl}(g)\;{\rightleftharpoons}\;\text{ClO}_2(g)\;+\;\text{N}_2\text{O}(g)\;+\;\text{ClO}(g)[/latex] (fast)
2) [latex]\text{N}_2\text{O}(g)\;+\;\text{ClO}_2(g)\;{\rightleftharpoons}\;\text{NO}_2(g)\;+\;\text{NOCl}(g)[/latex] (fast)
3) [latex]\text{NOCl}\;+\;\text{ClO}\;{\longrightarrow}\;\text{NO}_2\;+\;\text{Cl}_2[/latex] (slow)

Determine the overall reaction, write the rate law expression for each elementary reaction step, identify any intermediates, and determine the overall rate law expression.

Solution
For determining the overall reaction: sum the three steps, cancel species that appear on both sides, and combine same molecules on same side:

2NO2Cl(g) → 2NO2(g) + Cl2(g)

To write the rate law for each elementary reaction step, remember that for elementary reactions, the rate law expression can be derived directly from the stoichiometry (We will use: for step (1) k1 for rate constant for forward reaction and k−1 for rate constant for the reverse reaction; for step (2) k2 for forward reaction and k−2 for reverse reaction; for step (3) k3: forward reaction.)

rate1 = k1[NO2Cl]2 = k-1[ClO2][N2O][ClO]

rate2 = k2[N2O][ClO2] = k-2[NO2][NOCl]

rate3 = k3[NOCl][ClO]

The last step, which is the slow step, is the rate-determining step. Therefore, the overall rate law expression could be written as Rateoverall = k3[NOCl][ClO]. However, both NOCl and ClO are intermediates, with difficult to determine concentrations. Algebraic expressions must be used to represent [NOCl] and [ClO] in terms of measurable reactant and product concentrations.

Using step (1), [latex][\text{ClO}] = \frac{k_1[\text{NO}_2\text{Cl}]^2}{k_{-1}[\text{ClO}_2][\text{N}_2\text{O}]}[/latex].

Using step (2), [latex][\text{NOCl}] = \frac{k_2[\text{N}_2\text{O}][\text{ClO}_2]}{k_{-2}[\text{NO}_2]}[/latex].

Substitute into the overall rate law:

[latex]\begin{array}{r @{{}={}} l} \text{rate}_\text{overall} & k_3(\frac{k_2[\text{N}_2\text{O}][\text{ClO}_2]}{k_{-2}[\text{NO}_2]})(\frac{k_1[\text{NO}_2\text{Cl}]^2}{k_{-1}[\text{ClO}_2][\text{N}_2\text{O}]}) \\[1em] & (\frac{k_3k_2k_1}{k_{-2}k_{-1}})\frac{[\text{NO}_2\text{Cl}]^2}{[\text{NO}_2]} \\[1em] & k_\text{overall}\frac{[\text{NO}_2\text{Cl}]^2}{[\text{NO}_2]}\end{array}[/latex]

where, [latex]k_\text{overall}=\frac{k_3k_2k_1}{k_{-2}k_{-1}}[/latex].

Notice that this rate law shows an inverse dependence on the concentration of one of the product species, which is an effect of the equilibria present in this particular multistep reaction.

Check Your Learning
The reaction of methanol (CH3OH) and HBr is proposed to have two elementary reaction steps:

1) [latex]\text{CH}_3\text{OH}\;+\;\text{H}^+\;{\rightleftharpoons}\;\text{CH}_3\text{OH}_2^+[/latex] (fast)

2) [latex]\text{CH}_3\text{OH}_2^+\;+\;\text{Br}^-\;{\longrightarrow}\;\text{CH}_3\text{Br}\;+\;\text{H}_2\text{O}[/latex] (slow)

Determine the overall reaction, write the rate law expression for each elementary reaction, and determine the overall rate law expression.

Answer:

overall reaction: CH3OH + H+ + Br → CH3Br + H2

rate1 = k1[CH3OH][H+] = k1[CH3OH2+]

rate2 = k2[CH3OH2+][Br]

[latex]\text{rate}_\text{overall} = (\frac{k_2k_1}{k_{-1}})[\text{CH}_3\text{OH}][\text{H}^+][\text{Br}^-][/latex]

Ch10.13 Catalysis

The rate of a reaction can be accelerated by a catalyst, which speeds up the rate by lowering the activation energy. The catalyst itself is not consumed by the reaction, as it is regenerated in the process.

Catalysts function by providing an alternate reaction mechanism that has a lower activation energy than would be found in the absence of the catalyst. Often, the catalyzed mechanism include additional steps, as shown by the example reaction diagrams in Figure 11. The lower activation energy involved in the catalyzed pathway results in an increase in rate as described by the Arrhenius equation. Note that a catalyst decreases the activation energy for both the forward and the reverse reactions and hence accelerates both the forward and the reverse reactions. Consequently, the presence of a catalyst will permit a system to reach equilibrium more quickly, but it has no effect on the relative energies of the reactants and products.

A graph is shown with the label, “Extent of reaction,” appearing in a right pointing arrow below the x-axis and the label, “Energy,” in an upward pointing arrow just left of the y-axis. Approximately one-fifth of the way up the y-axis, a very short, somewhat flattened portion of both a red and a blue curve are shown. This region is labeled “Reactants.” A red concave down curve extends upward to reach a maximum near the height of the y-axis. This curve is labeled, “Uncatalyzed pathway.” From the peak, the curve continues downward to a second horizontally flattened region at a height of about one-third the height of the y-axis. This flattened region is labeled, “Products.” A second curve is drawn in blue with the same flattened regions at the start and end of the curve. The height of this curve is about two-thirds the height of the first curve and just right of its maximum, the curve dips low, then rises back and continues a downward trend at a lower height, but similar to that of the red curve. This blue curve is labeled, “Catalyzed pathway.”
Figure 11. This potential energy diagram shows the effect of a catalyst on the activation energy. The catalyst provides an alternate reaction pathway with a lower activation energy. As shown, the catalyzed pathway involves a two-step mechanism (note the presence of two transition states) and an intermediate species (represented by the valley between the two transitions states).

Example 3

Using Reaction Diagrams to Compare Catalyzed and Uncatalyzed Reactions
The two reaction diagrams here represent the same reaction: one without a catalyst and one with a catalyst. Identify which diagram suggests the presence of a catalyst, and determine the activation energy for the catalyzed reaction:
In this figure, two graphs are shown. The x-axes are labeled, “Extent of reaction,” and the y-axes are labeled, “Energy ( k J ).” The y-axes are marked off from 0 to 50 in intervals of five. In a, a blue curve is shown. It begins with a horizontal segment at about 6. The curve then rises sharply near the middle to reach a maximum of about 32 and similarly falls to another horizontal segment at about 10. In b, the curve begins and ends similarly, but the maximum reached near the center of the graph is only 20.

Solution
A catalyst does not affect the energy of reactants or products, so those aspects of the diagrams are the same. There is, however, a noticeable difference in the transition state, which is distinctly lower in diagram (b) than it is in (a). This indicates the use of a catalyst in (b). The activation energy is the difference between the energy of the starting reagents and the transition state—a maximum on the reaction coordinate diagram. The reagents are at 6 kJ/mol and the transition state is at 20 kJ/mol, so the activation energy can be calculated as follows:

Ea = 20 kJ/mol – 6 kJ/mol = 14 kJ/mol

Check Your Learning
Determine which of the two diagrams here (both for the same reaction) involves a catalyst, and identify the activation energy for the catalyzed reaction:
In this figure, two graphs are shown. The x-axes are labeled, “Extent of reaction,” and the y-axes are labeledc “Energy (k J).” The y-axes are marked off from 0 to 100 at intervals of 10. In a, a blue curve is shown. It begins with a horizontal segment at about 10. The curve then rises sharply near the middle to reach a maximum of about 91, then sharply falls to about 52, again rises sharply to about 73 and falls to another horizontal segment at about 5. In b, the curve begins and ends similarly, but the first peak reaches about 81, drops to about 55, then rises to about 77 before falling to the horizontal region at about 5.

Answer:

Diagram (b) is a catalyzed reaction with an activation energy of about 70 kJ/mol.

Ch10.14 Homogeneous Catalysts

A homogeneous catalyst is present in the same phase as the reactants. It interacts with a reactant to form an intermediate substance, which then decomposes or reacts with another reactant in one or more steps to regenerate the original catalyst and form product.

As an important illustration of homogeneous catalysis, consider the earth’s ozone layer. Ozone in the upper atmosphere, which protects the earth from ultraviolet radiation, is formed when oxygen molecules absorb UV photons and undergo the reaction:

[latex]3\text{O}_2(g)\;{\xrightarrow{hv}}\;2\text{O}_3(g)[/latex]

Ozone is a relatively unstable molecule that decomposes to yield O2 with the following mechanism:

[latex]\begin{array}{r @{{}\longrightarrow{}} l} \text{O}_3 & \text{O}_2\;+\;\text{O} \\[0.5em] \text{O}\;+\;\text{O}_3 & 2\text{O}_2 \end{array}[/latex]

The presence of nitric oxide, NO, influences the rate of decomposition of ozone. Nitric oxide acts as a catalyst in the following mechanism:

[latex]\begin{array}{r @{{}\longrightarrow{}} l} \text{NO}(g)\;+\;\text{O}_3(g) & \text{NO}_2(g)\;+\;\text{O}_2(g) \\[0.5em] \text{O}_3(g) & \text{O}_2(g)\;+\;\text{O}(g) \\[0.5em] \text{NO}_2(g)\;+\;\text{O}(g) & \text{NO}(g)\;+\;\text{O}_2(g) \end{array}[/latex]

Nitric oxide reacts and is regenerated in these reactions. It is not permanently used up; thus, it acts as a catalyst. The overall chemical change for the catalyzed mechanism is the same as the uncatalyzed reaction:

[latex]2\text{O}_3(g)\;{\longrightarrow}\;3\text{O}_2(g)[/latex]

The rate of decomposition of ozone is greater in the presence of nitric oxide because of the catalytic activity of NO. Certain compounds that contain chlorine also catalyze the decomposition of ozone.

Mario J. Molina

The 1995 Nobel Prize in Chemistry was shared by Paul J. Crutzen, Mario J. Molina (Figure 12), and F. Sherwood Rowland “for their work in atmospheric chemistry, particularly concerning the formation and decomposition of ozone.”

A photograph is shown of Mario Molina. To the right of the photo, an image of Earth’s southern hemisphere is shown with a central circular region in purple with a radius of about half that of the entire hemisphere. Just outside this region is a narrow royal blue band, followed by an outer thin turquoise blue band. The majority of the outermost region is green. Two small bands of yellow are present in the lower regions of the image.
Figure 12. (a) Mexican chemist Mario Molina (1943 –) shared the Nobel Prize in Chemistry in 1995 for his research on (b) the Antarctic ozone hole. (credit a: courtesy of Mario Molina; credit b: modification of work by NASA)

In 1974, Molina and Rowland published a paper in the journal Nature detailing the threat of chlorofluorocarbon gases to the stability of the ozone layer in earth’s upper atmosphere. The ozone layer protects earth from solar radiation by absorbing ultraviolet light. As chemical reactions deplete the amount of ozone in the upper atmosphere, a measurable “hole” forms above Antarctica, and an increase in the amount of solar ultraviolet radiation— strongly linked to the prevalence of skin cancers—reaches earth’s surface. The work of Molina and Rowland was instrumental in the adoption of the Montreal Protocol, an international treaty signed in 1987 that successfully began phasing out production of chemicals linked to ozone destruction.

Molina and Rowland demonstrated that chlorine atoms from human-made chemicals can catalyze ozone destruction in a process similar to that by which NO accelerates the depletion of ozone. Chlorine atoms are generated when chlorocarbons or chlorofluorocarbons—once widely used as refrigerants and propellants—are photochemically decomposed by ultraviolet light or react with hydroxyl radicals. For example:

[latex]\text{CH}_3\text{Cl}\;+\;\text{OH}\;{\longrightarrow}\;\text{Cl}\;+\;\text{other\;products}[/latex]

Chlorine radicals break down ozone and are regenerated by the following catalytic cycle:

[latex]\begin{array}{r @{{}\longrightarrow{}} l} 1)\;\;\;\text{Cl}\;+\;\text{O}_3 & \text{ClO}\;+\;\text{O}_2 \\[0.5em] 2)\;\;\;\text{ClO}\;+\;\text{O} & \text{Cl}\;+\;\text{O}_2 \\[0.5em] \text{overall\;Reaction:\;O}_3\;+\;\text{O} & 2\text{O}_2 \end{array}[/latex]

A single chlorine atom can break down thousands of ozone molecules. Luckily, the majority of atmospheric chlorine exists as the catalytically inactive forms Cl2 and ClONO2.

Ch10.15 Heterogeneous Catalysts

A heterogeneous catalyst is a catalyst that is present in a different phase (usually a solid) than the reactants. Such catalysts generally function by furnishing an active surface upon which a reaction can occur. Gas and liquid phase reactions catalyzed by heterogeneous catalysts occur on the surface of the catalyst rather than within the gas or liquid phase.

Heterogeneous catalysis has at least four steps:

  1. Adsorption of the reactant onto the surface of the catalyst
  2. Activation of the adsorbed reactant
  3. Reaction of the adsorbed reactant
  4. Diffusion of the product from the surface into the gas or liquid phase (desorption).

Any one of these steps may be slow and thus may serve as the rate determining step. In general, however, in the presence of the catalyst, the overall rate of the reaction is faster than it would be without the catalyst.

Figure 13 illustrates the reaction of alkenes with hydrogen on a nickel catalyst. Nickel is the catalyst used in the hydrogenation of polyunsaturated fats and oils to produce saturated fats and oils.

In this figure, four diagrams labeled a through d are shown. In each, a green square surface is shown in perspective to provide a three-dimensional appearance. In a, the label “N i surface” is placed above with a line segment extending to the green square. At the lower left and upper right, pairs of white spheres bonded tougher together appear as well as white spheres on the green surface. Black arrows are drawn from each of the white spheres above the surface to the white sphere on the green surface. In b, the white spheres are still present on the green surface. Near the center of this surface is a molecule with two central black spheres with a double bond indicated by two horizontal black rods between them. Above and below to the left and right, a total of four white spheres are connected to the black spheres with white rods. A line segment extends from this structure to the label, “Ethylene absorbed on surface breaking pi bonds.” Just above this is a nearly identical structure greyed out with three downward pointing arrows to the black and white structure to indicate downward motion. The label “Ethylene” at the top of the diagram is connected to the greyed out structure with a line segment. In c, the diagram is very similar to b except that the greyed out structure and labels are gone and one of the white spheres near the black and white structure in each pair on the green surface is greyed out. Arrows point from the greyed out white spheres to the double bond between the two black spheres. In d, only a single white sphere remains from each pair in the green surface. A curved arrow points from the middle of the green surface to a model above with two central black spheres with a single black rod indicating a single bond between them. Each of the black rods has three small white spheres bonded as indicated by white rods between the black spheres and the small white spheres. The four bonds around each black sphere are evenly distributed about the black spheres.
Figure 13. There are multiple steps in the catalytic hydrogenation of C2H4+H2⟶C2H6 by nickel. (a) Hydrogen is adsorbed on the surface, breaking the H–H bonds and forming Ni–H bonds. (b) Ethylene is adsorbed on the surface, breaking the π-bond and forming Ni–C bonds. (c) Atoms diffuse across the surface and form new C–H bonds when they collide. (d) C2H6 molecules escape from the nickel surface, since they are not strongly attracted to nickel.

Other significant industrial processes that involve the use of heterogeneous catalysts include the preparation of sulfuric acid, the preparation of ammonia, the oxidation of ammonia to nitric acid, and the synthesis of methanol. Heterogeneous catalysts are also used in the catalytic converters found on most gasoline-powered automobiles.

Automobile Catalytic Converters

Scientists developed catalytic converters to reduce the amount of toxic emissions produced by burning gasoline in internal combustion engines. Catalytic converters take advantage of all five factors that affect the speed of chemical reactions to ensure that exhaust emissions are as safe as possible.

By utilizing a carefully selected blend of catalytically active metals, it is possible to ensure all carbon-containing compounds are converted to carbon dioxide while also reducing the output of nitrogen oxides. This is particularly impressive when we consider that one step involves adding more oxygen to the molecule and the other involves removing the oxygen (Figure 14).

An image is shown of a catalytic converter. At the upper left, a blue arrow pointing into a pipe that enters a larger, widened chamber is labeled, “Dirty emissions.” A small black arrow that points to the lower right is positioned along the upper left side of the widened region. This arrow is labeled, “Additional oxygen from air pump.” The image shows the converter with the upper surface removed, exposing a red-brown interior. The portion of the converter closest to the dirty emissions inlet shows small, round components in an interior layer. This layer is labeled “Three-way reduction catalyst.” The middle region shows closely packed small brown rods that are aligned parallel to the dirty emissions inlet pipe. The final quarter of the interior of the catalytic converter again shows a layer of closely packed small red brown circles. Two large light grey arrows extend from this layer to the open region at the lower right of the image to the label “Clean emissions.”
Figure 14. A catalytic converter converts toxic gases and pollutents created by the internal combustion engine via usage of heterogeneous catalysts.

Most modern, three-way catalytic converters possess a surface impregnated with a platinum-rhodium catalyst, which catalyzes the conversion nitric oxide into N2 and O2:

2NO2(g) → N2(g) + 2O2(g)

as well as the conversion of carbon monoxide into carbon dioxide

2CO(g) + O2(g) → 2CO2(g)

and hydrocarbons such as octane into carbon dioxide and water vapor:

2C8H18(g) + 25O2(g) → 16CO2(g) + 18H2O(g)

In order to be as efficient as possible, most catalytic converters are preheated by an electric heater. This ensures that the metals in the catalyst are fully active even before the automobile exhaust is hot enough to maintain appropriate reaction temperatures.

Ch10.16 Enzymes

The study of enzymes is an important interconnection between biology and chemistry. Enzymes are usually proteins (polypeptides) that help to control the rate of chemical reactions between biologically important compounds, particularly those that are involved in cellular metabolism. Different classes of enzymes perform a variety of functions, a few examples are shown in Table 1.

Class Function
oxidoreductases redox reactions
transferases transfer of functional groups
hydrolases hydrolysis reactions
lyases group elimination to form double bonds
isomerases isomerization
ligases bond formation with ATP hydrolysis
Table 1. Classes of Enzymes and Their Functions

Enzyme molecules possess an active site, a part of the molecule with a shape that allows it to bond to a specific substrate (a reactant molecule), forming an enzyme-substrate complex as a reaction intermediate. There are two models that attempt to explain how this active site works. The most simplistic model is referred to as the lock-and-key hypothesis, which suggests that the molecular shapes of the active site and substrate are complementary, fitting together like a key in a lock. The induced fit hypothesis, on the other hand, suggests that the enzyme molecule is flexible and changes shape to accommodate a bond with the substrate. This is not to suggest that an enzyme’s active site is completely malleable, however. Both the lock-and-key model and the induced fit model account for the fact that enzymes can only bind with specific substrates, since in general a particular enzyme only catalyzes a particular reaction (Figure 15).

A diagram is shown of two possible interactions of an enzyme and a substrate. In a, which is labeled “Lock-and-key,” two diagrams are shown. The first shows a green wedge-like shape with two small depressions in the upper surface of similar size, but the depression on the left has a curved shape, and the depression on the right has a pointed shape. This green shape is labeled “Enzyme.” Just above this shape are two smaller, irregular, lavender shapes each with a projection from its lower surface. The lavender shape on the left has a curved projection which matches the shape of the depression on the left in the green shape below. This projection is shaded orange and has a curved arrow extending from in to the matching depression in the green shape below. Similarly, the lavender shape on the right has a projection with a pointed tip which matches the shape of the depression on the right in the green shape below. This projection is shaded orange and has a curved arrow extending from in to the matching depression in the green shape below. Two line segments extend from the depressions in the green shape to form an inverted V shape above the depressions. Above this and between the lavender shapes is the label, “Active site is proper shape.” The label “Substrates” is at the very top of the diagram with line segments extending to the two lavender shapes. To the right of this diagram is a second diagram showing the lavender shapes positioned next to each other, fit snugly into the depressions in the green shape, which is labeled “Enzyme.” Above this diagram is the label, “Substrate complex formed.” In b, which is labeled “Induced fit,” two diagrams are shown. The first shows a green wedge-like shape with two small depressions in the upper surface of similar size, but irregular shape. This green shape is labeled “Enzyme.” Just above this shape are two smaller irregular lavender shapes each with a projection from its lower surface. The lavender shape on the left has a curved projection. This projection is shaded orange and has a curved arrow extending from it to the irregular depression just below it in the green shape below. Similarly, the lavender shape on the right has a projection with a pointed tip. This projection is shaded orange and has a curved arrow extending from it to the irregular depression just below it in the green shape below. Two line segments extend from the depressions in the green shape to form an inverted V shape above the depressions. Above this and between the lavender shapes is the label, “Active site changes to fit.” The label, “Substrates” is at the very top of the diagram with line segments extending to the two lavender shapes. To the right of this diagram is a second diagram showing the purple shapes positioned next to each other, fit snugly into the depressions in the green shape, which is labeled “Enzyme.” Above this diagram is the label “Substrate complex formed.” The projections from the lavender shapes match the depression shapes in the green shape, resulting in a proper fit.
Figure 15. (a) According to the lock-and-key model, the shape of an enzyme’s active site is a perfect fit for the substrate. (b) According to the induced fit model, the active site is somewhat flexible, and can change shape in order to bond with the substrate.

Glucose-6-Phosphate Dehydrogenase Deficiency

Enzymes in the human body act as catalysts for important chemical reactions in cellular metabolism. As such, a deficiency of a particular enzyme can translate to a life-threatening disease. G6PD (glucose-6-phosphate dehydrogenase) deficiency, a genetic condition that results in a shortage of the enzyme glucose-6-phosphate dehydrogenase, is the most common enzyme deficiency in humans. This enzyme, shown in Figure 16, is the rate-limiting enzyme for the metabolic pathway that supplies NADPH to cells (Figure 17).

A colorful model of the Glucose-6-phosphate dehydrogenase structure is shown. The molecule has two distinct lobes which are filled with spiraled ribbon-like regions of yellow, lavender, blue, silver, green, and pink.
Figure 16. Glucose-6-phosphate dehydrogenase is a rate-limiting enzyme for the metabolic pathway that supplies NADPH to cells.

A disruption in this pathway can lead to reduced glutathione in red blood cells; once all glutathione is consumed, enzymes and other proteins such as hemoglobin are susceptible to damage. For example, hemoglobin can be metabolized to bilirubin, which leads to jaundice, a condition that can become severe. People who suffer from G6PD deficiency must avoid certain foods and medicines containing chemicals that can trigger damage their glutathione-deficient red blood cells.

A reaction mechanism is diagrammed in this figure. At the left, the name Glucose is followed by a horizontal, right pointing arrow, labeled, “Hexokinase.” Below this arrow and to the left is a yellow star shape labeled, “A T P.” A curved arrow extends from this shape to the right pointing arrow, and down to the right to a small brown oval labeled, “A D P.” To the right of the horizontal arrow is the name Glucose 6 phosphate, which is followed by another horizontal, right pointing arrow which is labeled, “G 6 P D.” A small orange rectangle below and left of this arrow is labeled “N A D P superscript plus.” A curved arrow extends from this shape to the right pointing arrow, and down to the right to a small salmon-colored rectangle labeled “N A P D H.” A curved arrow extends from this shape below and to the left, back to the orange rectangle labeled, “N A D P superscript plus.” Another curved arrow extends from a green oval labeled “G S S G” below the orange rectangle, up to the arrow curving back to the orange rectangle. This last curved arrow continues on to the lower right to a second green oval labeled, “G S H.” The end of this curved arrow is labeled, “Glutathione reductase.” To the right of the rightmost horizontal arrow appears the name 6 phosphogluconate.
Figure 17. In the mechanism for the pentose phosphate pathway, G6PD catalyzes the reaction that regulates NAPDH, a co-enzyme that regulates glutathione, an antioxidant that protects red blood cells and other cells from oxidative damage.

Ch10.17 Enzyme Kinetics

The Michaelis-Menten mechanism is one which many enzyme catalyzed reactions follow. The basic mechanism involves an enzyme (E) and a substrate (S) which must connect to form an enzyme-substrate complex (ES) in order for the substrate to be reacted to form a product (P). The overall reaction is:

S → P

And the simple two-step mechanism is:

$E+S \xrightleftharpoons[k_{-1}]{k_1} ES $
$ES \xrightarrow{k_2} P+E$

The kinetics of this reaction can be derived using the steady state approximation, which applies to reactions where the concentration of a highly reactive (i.e. short-lived) intermediate is approximately constant over the course of the reaction. Hence, for such an intermediate (I),

$\dfrac{d[I]}{dt}=0$

In the enzyme catalyzed reaction, we can apply steady state approximation to the concentration of the ES intermediate complex:

$\dfrac{d[ES]}{dt}=0=k_1[E][S]-k_{-1}[ES]-k_2[ES]$

Solving for [ES] gives:

$[ES]=\dfrac{k_1[E][S]}{k_{-1}+k_2}$

or

$[ES]=\dfrac{[E][S]}{K_M}$

where KM is the Michaelis constant,

$K_M=\dfrac{k_{-1}+K_2}{k_1}$

which is affected by a number of factors, including pH, temperature, and the nature of the substrate itself.

Using a mass conservation relationship on the enzyme—the enzyme must be either in its bare form (E) or complexed with a substrate (ES), we have:

[E]total = [E] + [ES]

Or:

[E] = [E]total – [ES]

Substituting this into the above equation:

$[ES]=\dfrac{([E]_{\text{total}}-[ES])[S]}{K_M}$

Rearranging the variables:

$[ES]=\dfrac{[E]_{\text{total}}[S]}{K_M}-\dfrac{[ES][S]}{K_M}$
$\left(\dfrac{[S]}{K_M}+1\right)[ES]=\dfrac{[E]_{\text{total}}[S]}{K_M}$
$[ES]=\left(\dfrac{[E]_{\text{total}}[S]}{K_M}\right)\left(\dfrac{K_M}{[S]+K_M}\right)=\dfrac{[E]_{\text{total}}[S]}{[S]+K_M}$

Substituting this into the expression for the rate of production of the product, or the observed reaction rate:

$rate=\dfrac{d[P]}{dt}=k_2[ES]$

yields

$rate=\dfrac{d[P]}{dt}=\dfrac{k_2[E]_{\text{total}}[S]}{[S]+K_M}$

or

$rate=\dfrac{V_{\text{max}}[S]}{[S]+K_M}$

where Vmax = k2[E]total is the maximum rate the reaction can achieve.

In the limit that [S] is very large, where [S] ≫ KM and [S] + KM ≈ [S], the reaction reaches its maximum rate (rate = Vmax) and becomes independent of any concentrations.

In the limit that [S]≪KM, hence [S] + KM ≈ KM, the reaction is first order in [S]:

$rate=\left(\dfrac{V_{\text{max}}}{K_M}\right)[S]$
Figure 18. Reaction rate as a function of substrate concentration. At low substrate concentrations, the reaction rate is first order in [S]. As substrate concentration increases, the reaction rate approaches Vmax.

The Michaelis constant KM is the substrate concentration at which the reaction rate is at half-maximum, and is an inverse measure of the substrate’s affinity for the enzyme—as a small KM indicates high affinity, meaning that the rate will approach Vmax more quickly. The value of KM is dependent on both the enzyme and the substrate, as well as conditions such as temperature and pH.

Denaturation

Enzymes are generally globular proteins, acting alone or in larger complexes. Like all proteins, enzymes are linear chains of amino acids that fold to produce a three-dimensional structure. The sequence of the amino acids specifies the structure which in turn determines the catalytic activity of the enzyme. Enzyme structures unfold (denature) when heated or exposed to chemical denaturants and this disruption to the structure typically causes a loss of activity.

Protein folding is key to whether a globular protein or a membrane protein can do its job correctly. It must be folded into the right shape to function. But hydrogen bonds, which play a big part in folding, are rather weak, and it does not take much heat, acidity, or other stress to break some and form others, denaturing the protein. This is one reason why tight homeostasis is physiologically necessary in many life forms.

Denaturation is a process in which proteins or nucleic acids lose the quaternary structure, tertiary structure and secondary structure which is present in their native state, by application of some external stress or compound such as a strong acid or base, a concentrated inorganic salt, an organic solvent (e.g., alcohol or chloroform), radiation or heat. If proteins in a living cell are denatured, this results in disruption of cell activity and possibly cell death. Denatured proteins can exhibit a wide range of characteristics, from conformational change and loss of solubility to aggregation due to the exposure of hydrophobic groups.

Figure 19. Enzyme activity initially increases with temperature until the enzyme’s structure unfolds (denaturation), leading to an optimal rate of reaction at an intermediate temperature. Image used with permission (CC-BY-SA-4.0; Thomas Shafee)

Enzyme denaturation is normally linked to temperatures above a species normal level; as a result, enzymes from bacteria living in volcanic environments such as hot springs are prized by industrial users for their ability to function at high temperatures, allowing enzyme-catalyzed reactions to be operated at a very high rate.

Inhibition

When an inhibitor interacts with an enzyme it decreases the enzyme’s catalytic efficiency. An irreversible inhibitor covalently binds to the enzyme’s active site, producing a permanent loss in catalytic efficiency even if we decrease the inhibitor’s concentration. A reversible inhibitor forms a noncovalent complex with the enzyme, resulting in a temporary decrease in catalytic efficiency. If we remove the inhibitor, the enzyme’s catalytic efficiency returns to its normal level.

There are several pathways for the reversible binding of an inhibitor to an enzyme, as shown in Figure 20. In competitive inhibition the substrate and the inhibitor compete for the same active site on the enzyme. Because the substrate cannot bind to an enzyme–inhibitor complex, EI, the enzyme’s catalytic efficiency for the substrate decreases. With noncompetitive inhibition the substrate and the inhibitor bind to different active sites on the enzyme, forming an enzyme–substrate–inhibitor, or ESI complex. The formation of an ESI complex decreases catalytic efficiency because only the enzyme–substrate complex reacts to form the product. Finally, in uncompetitive inhibition the inhibitor binds to the enzyme–substrate complex, forming an inactive ESI complex.

Figure 20. Mechanisms for the reversible inhibition of enzyme catalysis. E: enzyme, S: substrate, P: product, I: inhibitor, ES: enzyme–substrate complex, EI: enzyme–inhibitor complex, ESI: enzyme–substrate–inhibitor complex.

License

Icon for the Creative Commons Attribution-NonCommercial-ShareAlike 4.0 International License

Chemistry 109 Copyright © by John Moore; Jia Zhou; and Etienne Garand is licensed under a Creative Commons Attribution-NonCommercial-ShareAlike 4.0 International License, except where otherwise noted.