Print

Chapter 9 – Reaction Energetics

Introduction

We use the term energetics to combine two very important fields of study: thermodynamics and kinetics. Thermodynamics is the study of energy and its transformations. Kinetics is the study of the rates and mechanisms of reactions. A thermodynamic study of a reaction examines energy differences between the reactants and products. These studies allow us to predict how much product can be obtained. Thermodynamics is not concerned with how long the reaction takes or how it occurs. These latter concerns are the domain of kinetics. In this chapter, we examine the energetics of some reactions to better understand how reactions occur, to learn how to predict the feasibility of a reaction, and to determine the effect of temperature on both the amount of product and the rate of reaction.

9.1 First Law of Thermodynamics

Introduction

Thermodynamics is the study of energy and its transformations. In chemistry, thermodynamics studies consider the energy difference between the reactants and products. These studies allow us to predict the amount of product that we can expect. We start our study of thermodynamics with the first law of thermodynamics—energy cannot be created or destroyed.

Objectives

9.1-1. Energy Change

The sign of ΔE indicates the direction of energy flow.
The energy change of a system is defined as the system's final energy minus its initial energy; that is,
ΔE = EfinalEinitial.

9.1-2. Sign of ΔE Exercise

Exercise 9.1:

Indicate whether
ΔE < 0,
ΔE > 0,
or
ΔE = 0
for each of the following. Assume that the system is the object mentioned in the problem.
    heating water 25 °C to 50 °C
  • ΔE < 0 The energy of an object always increases with temperature, so ΔE > 0.
  • ΔE > 0
  • ΔE = 0 The energy of an object always increases with temperature, so ΔE > 0.
    moving positive charge closer to negative charge
  • ΔE < 0
  • ΔE > 0 Opposite charges attract because their energy is reduced in doing so. Thus, ΔE < 0.
  • ΔE = 0 Opposite charges attract because their energy is reduced in doing so. Thus, ΔE < 0.
    evaporating 1 g of water at 25 °C
  • ΔE < 0 Recall from Chapter 8 that the energy of the gaseous state is much greater that of than the liquid state, so evaporation always involves an increase in energy, so ΔEvap > 0.
  • ΔE > 0
  • ΔE = 0 Recall from Chapter 8 that the energy of the gaseous state is much greater that of than the liquid state, so evaporation always involves an increase in energy, so ΔEvap > 0.
    moving a ball from a table top to the floor
  • ΔE < 0
  • ΔE > 0 The potential energy of the ball is decreased due to the gravitational field of the earth, so ΔE < 0.
  • ΔE = 0 The potential energy of the ball is decreased due to the gravitational field of the earth, so ΔE < 0.
    heating a piece of zinc from 25 °C to 75 °C and then cooling it back to
    25 °C
  • ΔE < 0 The initial and final temperatures are the same, so Efinal = Einitial. Therefore, ΔE = EfinalEinitial = 0.
  • ΔE > 0 The initial and final temperatures are the same, so Efinal = Einitial. Therefore, ΔE = EfinalEinitial = 0.
  • ΔE = 0

9.1-3. First Law

Energy can be moved from one place to another, but it cannot be created or destroyed.
The system in a thermodynamic problem is simply that portion of the universe that is being studied. It exchanges energy with its surroundings. The system and its surroundings constitute the thermodynamic universe.
ΔEuniv = ΔE + ΔEsur
Note that both the universe and surroundings quantities are subscripted, but the system quantity is not. The first law of thermodynamics, states that energy is neither created nor destroyed.
( 9.1 )
ΔEuniv = 0 for all processes
First Law of Thermodynamics
ΔEuniv = 0
implies the following.
( 9.2 )
ΔE = −ΔEsur
Energy Exchange
We conclude that the first law of thermodynamics implies that the energy of a system can be changed only by exchanging energy with its surroundings.

9.1-4. Determining ΔE Exericse

Exercise 9.2:

75 J of energy is transferred when a hot rod is cooled in water. Determine the value of ΔE for the process when each of the following are considered to be the system.
rod -75_0__ The rod cools, so it gives up energy, i.e., its energy decreases. ΔE = –75 J if the rod is the system. J
water 75_0__ The 75 J of energy that is lost by the rod must enter the water, so the energy of the water increases. Thus, ΔE = +75 J if the water is the system. J
water and rod 0_0__ If energy is only transferred between the water and the rod and none enters or leaves the system from its surroundings, then ΔE = 0. J
A reaction releases 175 kJ of energy. Answer the following if the reaction is the system.
ΔE = -175_0__ The system releases heat, so Efinal < Einitial, which means that ΔE < 0. Consequently, ΔE = –175 kJ. kJ
ΔEsur = 175_0__ ΔEsur = –ΔE = –(–175 kJ) = +175 kJ. The surroundings must absorb the heat released by the system. kJ
ΔEuniv = 0_0__ ΔEuniv = 0 for all processes. kJ

9.1-5. Heat and Work

When q and w are positive, energy flows into the system.
Heat (q) and work (w) are the two most common ways to move energy between a system and its surroundings. The energy of a system increases when it absorbs heat or when work is done on it, and its energy decreases when it gives off heat or does work. Thus, the following is another expression for the first law of thermodynamics.
( 9.3 )
ΔE = q + w
Heat and Work in Energy Flow
q is the heat absorbed by the system and w is the work done on the system.
Table 9.1: The Signs of Heat and Work and the Direction of Energy Flow

9.1-6. Heat and Work Exercise

Exercise 9.3:

A system absorbs 500 J of heat and does 300 J of work.
What is the value of q?
q = 500_0__ System absorbs 500 J of heat, which increases its energy increases by 500 J.
q = +500 J – an endothermic process
J
What is the value of w?
w = -300_0__ The system does 300 J of work, so its energy drops by 300 J.
w = –300 J
J
What is the energy change for the process?
ΔE = 200_0__ ΔE = q + w = +500 – 300 = +200 J

The energy of the system increases by 200 J.
J

9.2 Enthalpy

Introduction

Typically, reactions are carried out at constant pressure and the initial and final temperatures are the same. Gases that are produced by a reaction that is open to the atmosphere can carry away some of the energy of the reaction (they do work when they escape), and that energy is lost. Thus, the thermodynamic property of a reaction that is carried out at constant pressure that is most important is the heat that is exchanged with the surroundings. In this section, we define that thermodynamic property.

Objectives

9.2-1. Enthalpy of Reaction

The enthalpy of reaction is the heat absorbed at constant T and P. It is positive for endothermic reactions and negative for exothermic reactions.
The enthalpy of reaction, ΔH, is the heat absorbed by a reaction when it is carried out at constant temperature and pressure. Its sign indicates the direction of heat flow. ΔH is defined as the heat absorbed, so it is positive if the reaction absorbs heat from the surroundings and negative if the reaction gives off heat to the surroundings. The sign of the heat flow is often designated by indicating that the heat was absorbed (positive) or released (negative). Thus, +10 J is read 'ten joules are absorbed,' while –10 J is read 'ten joules are released.'
The reactants are at higher energy than the products in exothermic reactions.
exothermic reaction
Figure 9.1a: Exothermic Processes
The reactants are higher in enthalpy than the products, so energy must be given off to the surroundings when the reactants are converted to products. Thus,
ΔH < 0
for the reaction. The reaction is exothermic because it gives off heat to the surroundings causing the surroundings to warm.
The reactants are at lower energy than the products in endothermic reactions.
endothermic reaction
Figure 9.1b: Endothermic Processes
The reactants are lower in enthalpy than the products, so energy must be absorbed from the surroundings to convert the reactants to products. Thus,
ΔH > 0
for the reaction. The reaction is endothermic because it absorbs heat from the surroundings causing the surroundings to cool.

9.2-2. Standard States

Standard states are defined at 1 atm pressure, but there is no standard temperature.
Thermodynamic properties, such as ΔE and ΔH, depend upon the states of the substances. Consequently, tabulated values of thermodynamic properties are given for reactions in which the reactants and products are all in their standard states. The standard state of a substance also depends upon the temperature, but there is no standard temperature. Tabulated values of thermodynamic properties are most often given at a temperature of 298 K (25 °C), and 298 K should be assumed if no temperature is given.
Exercise 9.4:

Indicate the standard state of each of the following.
    oxygen at –20 °C
  • solid O2 is a gas at –20 °C. The standard state of gases is 1 atm pressure, so the standard state of O2 at –20 °C is 1 atm of gas.
  • liquid O2 is a gas at –20 °C. The standard state of gases is 1 atm pressure, so the standard state of O2 at –20 °C is 1 atm of gas.
  • 1 atm gas
  • 1 M concentration O2 is a gas at –20 °C. The standard state of gases is 1 atm pressure, so the standard state of O2 at –20 °C is 1 atm of gas.
    water at –20 °C
  • solid
  • liquid The stable form of H2O at –20 °C is the solid, so ice is the standard state of H2O at –20 °C.
  • 1 atm gas The stable form of H2O at –20 °C is the solid, so ice is the standard state of H2O at –20 °C.
  • 1 M concentration The stable form of H2O at –20 °C is the solid, so ice is the standard state of H2O at –20 °C.
    water at +20 °C
  • solid The stable form of H2O at +20 °C is the liquid, so liquid water is the standard state of H2O at +20 °C.
  • liquid
  • 1 atm gas The stable form of H2O at +20 °C is the liquid, so liquid water is the standard state of H2O at +20 °C.
  • 1 M concentration The stable form of H2O at +20 °C is the liquid, so liquid water is the standard state of H2O at +20 °C.
    table sugar (C12H22O11) at 20 °C
  • solid
  • liquid The stable form of C12H22O11 at 20 °C is the solid, so C12H22O11 crystals are the standard state of C12H22O11 at 20 °C.
  • 1 atm gas The stable form of C12H22O11 at 20 °C is the solid, so C12H22O11 crystals are the standard state of C12H22O11 at 20 °C.
  • 1 M concentration The stable form of C12H22O11 at 20 °C is the solid, so C12H22O11 crystals are the standard state of C12H22O11 at 20 °C.
    table sugar dissolved in water at
    20 °C
  • solid The standard state of a dissolved substance is a solution in which its concentration is 1 mol per liter. Therefore, the standard state of a solution of C12H22O11 is a solution in which there are 342 g (1 mole) of C12H22O11 per liter of solution.
  • liquid The standard state of a dissolved substance is a solution in which its concentration is 1 mol per liter. Therefore, the standard state of a solution of C12H22O11 is a solution in which there are 342 g (1 mole) of C12H22O11 per liter of solution.
  • 1 atm gas The standard state of a dissolved substance is a solution in which its concentration is 1 mol per liter. Therefore, the standard state of a solution of C12H22O11 is a solution in which there are 342 g (1 mole) of C12H22O11 per liter of solution.
  • 1 M concentration

9.2-3. Standard Enthalpy of Reaction

The enthalpy change for a reaction in which all reactants and products are in their standard states is called the standard enthalpy of reaction and given the symbol ΔH° (the superscript '°' implies standard state conditions). Most enthalpies of reaction used in this course are standard enthalpies. A thermochemical equation is a balanced chemical equation that includes a thermodynamic property such as
ΔH°. Enthalpy changes depend upon the states of the substances, so the state is indicated by the following in parentheses.
The amount of heat released during reaction depends upon the amount of material that is consumed or produced. The enthalpy given for a thermochemical equation is the amount of heat given off or absorbed when the specified number of moles of each substance reacts or is produced. As an example, consider the following thermochemical equation.
2 Na(s) + 2 H2O(l) 2 NaOH(aq) + H2(g)    ΔH° = –368 kJ
The above thermochemical equation tells us that 368 kJ of heat are liberated when

9.2-4. Stoichiometry and Thermodynamics

The thermochemical equation,
P4(s) + 6 Cl2(g) 4 PCl3(g)    ΔH° = –1152 kJ
indicates that 1152 kJ of heat is liberated (negative sign) when one mole of solid phosphorus reacts with six moles of chlorine gas to produce four moles phosphorus trichloride gas. Consequently, the heat of reaction can also be expressed as any of the following.
−1152 kJ
1 mol P4
    or    
−1152 kJ
6 mol Cl2
    or    
−1152 kJ
4 mol PCl3
Multiplication of any of the expressions above by the appropriate number of moles of product or reactant results in the ΔH° for the reaction in which that number of moles is formed or consumed. For example, the following determines the enthalpy change that would accompany the reaction of four moles of chlorine gas.
4 mol Cl2 ×
−1152 kJ
6 mol Cl2
= −768 kJ

9.2-5. Stoichiometry Exercise

Exercise 9.5:

Consider the thermochemical equation.
N2(g) + 3 H2(g) 2 NH3(g)    ΔH° = –92 kJ
What would be the value of ΔH° under the following circumstances.
one mole of ammonia is formed -46___ The thermochemical equation indicates that 92 kJ of heat are released (minus sign) when 2 moles of NH3 are formed. Thus,
1 mol NH3 ×
−92 kJ
2 mol NH3
= −46 kJ.
46 kJ would be released.
kJ
two moles of H2 react -61.3333___ The thermochemical equation indicates that 92 kJ of heat are released (minus sign) when 3 moles of H2 react. Thus,
2 mol H2 ×
−92 kJ
3 mol H2
= −61 kJ.
61 kJ would be released.
kJ
7.0 g of N2 react -23___ The thermochemical equation indicates that 92 kJ of heat are released (minus sign) when 1 mole of N2 reacts. Thus, ΔH° = (–92 kJ)/(1 mol N2), which can be multiplied by the number of moles of N2 to obtain the answer.
7.0 g N2 ×
1 mol N2
28 g N2
= 0.25 mol N2

0.25 mol N2 ×
−92 kJ
1 mol N2
= −23 kJ
23 kJ would be released.
kJ

9.2-6. Properties of Enthalpy

We have seen that ΔH is proportional to the number of moles of substance that react. This means the following.
If the coefficients of a chemical equation are multiplied by some number, then the enthalpy of the resulting reaction equals the enthalpy of the original reaction multiplied by the same number.
For example consider the following two processes.
H2O(s) H2O(l)    ΔH° = +6 kJ
2 H2O(s) 2 H2O(l)    ΔH° = +12 kJ
The second equation is obtained by multiplying the first by two, so the enthalpy change of the second reaction is twice the enthalpy change of the first. A second important property of enthalpy is the following.
The sign of ΔH is changed when the direction of the reaction is reversed.
For example consider the following two processes.
H2O(s) H2O(l)    ΔH° = +6 kJ
H2O(l) H2O(s)    ΔH° = –6 kJ
The first thermochemical equation indicates that 6 kJ are absorbed when 1 mole of ice melts. In the second equation, the direction of the process is reversed, so the sign of ΔH must be changed. We conclude that freezing
1 mole of liquid water releases 6 kJ of heat.
Exercise 9.6:

The following thermochemical equation for burning butane is given.
C4H10(l) + 6.5 O2(g) 4 CO2(g) + 5 H2O(l)    Δ H° = –2855 kJ
Determine the enthalpy of reaction for each of the following.
2 C4H10(l) + 13 O2(g) 8 CO2(g) + 10 H2O(l)
-5710___ The reaction involves twice as much butane, so the enthalpy change of the given reaction must be multiplied by two:
ΔH = 2(−2855) = −5710 kJ.
The resulting thermochemical equation is 2 C4H10(l) + 13 O2(g) 8 CO2(g) + 10 H2O(l)    ΔH° = –5710 kJ.
kJ
4 CO2(g) + 5 H2O(l) C4H10(l) + 6.5 O2(g)
2855___ The reaction is simply the reverse of the original reaction, so the sign of the enthalpy change of the given reaction must be changed:
ΔH = −(−2855) = +2855 kJ.
The resulting thermochemical equation is 4 CO2(g) + 5 H2O(l) C4H10(l) + 6.5 O2(g)    ΔH° = +2855 kJ.
kJ
8 CO2(g) + 10 H2O(l) 2 C4H10(l) + 13 O2(g)
5710___ The reaction involves twice the amounts in the reverse of the original reaction, so the given enthalpy change must be multiplied by two and the sign must be changed, i.e., the original value must be multiplied by –2:
ΔH = (−2)(−2855) = +5710 kJ.
The resulting thermochemical equation is 8 CO2(g) + 10 H2O(l) 2 C4H10(l) + 13 O2(g)    ΔH° = +5710 kJ.
kJ

9.3 Enthalpies of Combustion

Introduction

Most of the energy utilized by humans comes from combustion reactions (reactions with O2). The internal combustion engine that runs our cars gets its energy from the combustion of gasoline. Many homes are heated by the combustion of coal, oil, or natural gas. Many power plants generate electricity by burning (combusting) coal, and humans obtain the energy needed to sustain life by the combustion of carbohydrates. The carbon in the fuel is converted to CO2 and the hydrogen is converted to H2O by the reaction with O2. In this section, we examine enthalpies (heats) of combustion.

Objectives

9.3-1. Enthalpies of Combustion

Combustion is the reaction of a substance with oxygen, and it is the major source of harnessed energy. Combustion reactions are always exothermic, and it is the energy released in combustion reactions that fuels our vehicles, heats our homes, and powers our bodies. The products of combustion are usually, but not always, the oxides of the elements comprising the molecule being combusted.
The standard molar enthalpy of combustion of a substance is the heat absorbed when one mole of the substance reacts with O2 at standard conditions.
Consider the following thermochemical equations that deal with some common combustion reactions.
Table 9.2: Thermochemical Equations for Some Common Combustion Reactions

9.3-2. Combustion Exercise

Exercise 9.7:

The calorie content of carbohydrates, fats, and proteins are all derived from their heats of combustion because they are all burned in the body. In this example, we use the heat of combustion of sugar to determine the number of calories in a teaspoon of sugar. The thermochemical equation for the combustion of sucrose (table sugar) is the following.
C12H22O11(s) + 12 O2(g) 12 CO2(g) + 11 H2O(l)    ΔH° = –5650 kJ/mol
How many calories are in a heaping teaspoon of sugar (4.8 g)?
What is the number of moles of sugar present (Mm = 342 g/mol)?
0.014___
4.8 g C12H22O11 ×
1 mol C12H22O11
342 g C12H22O11
= 0.014 mol C12H22O11
mol
What is the amount of heat liberated during the combustion?
79.298___
0.014 mol ×
5650 kJ released
1 mol
= 79 kJ released
kJ
If there are 4.18 kJ in a dietary calorie, what is the number of dietary calories present in the sugar?
18.97___
79 kJ ×
1 Cal
4.18 kJ
= 19 Cal
Cal

9.4 Bond Energies

Introduction

Recall from Chapter 5 that bond formation involves a lowering of the energy of the system, i.e., bond formation is an exothermic process. Reversing the direction of the reaction, changes the sign of ΔH, so bond breaking is an endothermic process. The energy released or absorbed during a chemical reaction comes from or is stored in chemical bonds, and the enthalpy change of a reaction can be estimated from tabulated values of bond energies.

Prerequisites

Objectives

9.4-1. Bond Energies

The bond energy or dissociation energy (D) of a bond is the energy required to break one mole of bonds in the gas phase.
A table of bond energies can be found in the resource named Bond Energies. Bond energies are always positive because bonds require energy to be broken.
Table 9.3: Examples of Some Thermochemical Equations involving Bond Energies

9.4-2. Estimating Enthalpy

Reactions involve breaking and forming bonds, and the heat generated or absorbed during a reaction is the net result. If we knew the energies of all of the interactions involved, we could determine the enthalpy of reaction by determining how much energy must be supplied to break all of the interactions that had to be broken and subtracting the energy that is released when all of the new interactions form. Bond energies give us estimates of these energies, but tabulated bond energies are averages. For example, a C-Cl bond in CCl4 is different from a C-Cl bond in COCl2, but only one C-Cl bond energy is tabulated. In addition, bond energies are defined only for gas phase molecules, so other interactions can be ignored, but many reactions occur in solution where solvation and other interactions complicate the problem. However, bond energies can be used to obtain approximate enthalpies from the following.
The enthalpy of a reaction can be approximated by the sum of the bond energies of the bonds that must be broken less the sum of the bond energies of those that must be formed.
( 9.4 )
ΔH ~
Dbroken
Dformed
Enthalpy of Reaction from Bond Energies
Example:
For example, let us approximate the enthalpy of reaction for the following.
H2(g) + F2(g) 2 HF(g)
One H-H bond and one F-F bond must be broken and two H-F bonds must be formed. The needed bond energies are given in Table 9.3.
ΔH~
Dbroken
Dformed
~DH-H + DF-F − 2DH-F
= 436 + 155 − 2(565)
= −539 kJ
The above calculation works well for the HF example because there is only one kind of H-H, F-F, and H-F bonds and the reaction is a gas phase reaction.
Example:
Use tabulated bond energies to estimate the enthalpy of the following reaction.
C2H2 + H2 C2H4

9.4-3. Bond Energy Exercise

Exercise 9.8:

Use the Bond Energies resource to estimate the enthalpy changes for the following reactions.
exothermic reaction
ΔH = -91___ Break one N=N bond and one H-H bond and form one N-N bond and two N-H bonds.

ΔH ~ DN=N + DH-HDN-NDN-H = 418 + 436 – 163 – 2(391) = –91 kJ/mol
kJ/mol
extothermic reaction
ΔH = -1238___ Break two C-H bonds, one C≡C bond and 2.5 O=O bonds and form four C=O bonds and two O-H bonds.

ΔH ~ 2DC-H + DC≡C + 2.5DO=O – 4DC=O – 2DO-H = 2(413) + 820 + 2.5(495) – 4(799) – 2(463) = –1238 kJ/mol
kJ/mol

9.4-4. Bond Energy and Lewis Structure Exercise

Exercise 9.9:

Determine ΔH for the following reaction.
H2CO(g) + NH3(g) H2CNH(g) + H2O(g)
First determine the Lewis structures to get the bond orders of all bonds. Review Section 5.6 Determining Lewis Structures if you have forgotten how to determine Lewis structures.
Bond Orders: All bonds to hydrogen are single bonds, but the other bond orders need to be determined.
C-O bond order in H2CO
2___ There are 4 shared pairs in H2CO, 2 are used for C-H bonds, so 2 are in the C-O bond.
C-N bond order in H2CNH
2___ There are 5 shared pairs in H2CNH. Three are used for CH and NH bonds, so 2 are in the C-N bond.
Enthalpies: Use the Bond Energies resource to determine the following enthalpies. Break and form only those bonds that do not appear in both the reactant and product side.
energy required to break the reactant bonds
1581___ The two C-H bonds and one N-H bond appear on both sides of the equation, so they are not included in the calculation.

D(C=O) + 2D(N-H) = 799 + 2(391) = 1581 kJ
kJ
energy released when product bonds are formed
1541___ The two C-H bonds and one N-H bond appear on both sides of the equation, so they are not included in the calculation.

D(C=N) + 2D(O-H) = 615 + 2(463) = 1541 kJ
kJ
Finally, determine the enthalpy of reaction.
ΔH = 40___ Energy required to break reactant bonds – energy released when product bonds are formed = 1581 – 1541 = 40 kJ. kJ

9.5 Entropy

Introduction

A spontaneous process is one that takes place without intervention. A ball rolls down a hill spontaneously, but it does not roll uphill spontaneously. It can be made to roll uphill but only with intervention. We have used the concept that systems in nature strive to lower energy, so our first hypothesis might be that exothermic processes are spontaneous, while endothermic processes are not. However, there are spontaneous endothermic processes, such as evaporation. We saw in Chapter 8 that the energy of the molecules in the gas phase is much greater than those in the liquid phase, yet liquids do evaporate spontaneously, i.e., molecules go spontaneously from the lower energy liquid state into the higher energy gas phase. Clearly, there is another factor, and, in this section, we consider that factor.

Objectives

9.5-1. Probability and Constraints

Unconstrained systems are statistically favored over constrained ones.
There are
10! = 3,628,800
ways to arrange ten books on a shelf, but the books are aphabetized in only one. Thus, there are 3,628,799 ways to disburse the books in a random or disordered manner, but only one in which the books are constrained to be in alphabetical order. If the books were alphabetized originally but then removed and returned to the shelf without the intervention of a librarian, they would spontaneously move to a disordered state for strictly statistical reasons. If we assume that there are only two systems in our set of books, constrained or ordered and unconstrained or disordered, then the process constrained unconstrained is spontaneous because there are more ways in which the books can be disbursed in the unconstrained, disordered system. The process
unconstrained constrained is not spontaneous and occurs only with intervention (a librarian in this example).

9.5-2. Types of Motion

The kinetic energy of a molecule is a combination of translational, rotational, and vibrational contributions.
We saw above that systems that can be disbursed in a less constrained manner are favored over those whose disbursement is highly constrained, and it is the disbursement of energy that drives chemical processes. We now show how molecular systems disburse their energy. Recall from Section 7.1 Gases that thermal energy (RT) is a measure of the average kinetic energy of the molecules in a system. Kinetic energy is energy of motion, and the motion of a molecule is a combination of three types of motion called degrees of freedom. All atoms have three degrees of freedom because the motion of an atom can be described as a combination of its motion in the x-, y-, and z-directions. A molecule that contains n atoms has 3n degrees of freedom that are combinations of the three motions of its n atoms. The six degrees of freedom of a diatomic molecule are shown in Figure 9.2.
Figure 9.2: Degrees of Freedom of a Diatomic Molecule
Each atom of a diatomic molecule like CO has three degrees of freedom as they can each move in the x-, y-, or z-direction. Thus, the molecule has six degrees of freedom, which take the following forms.

9.5-3. Entropy

Entropy is a measure of the number of ways that a system can distribute its energy.
For purely statistical reasons, a system that has more ways in which it can distribute its energy is favored over one that has fewer ways. Indeed, the number of ways in which the energy of a system can be distributed is such an important property of the system that a thermodynamic property is used to measure it. The thermodynamic property is called the entropy (S) of the system, which has units of joules per kelvin (J/K). The entropy of a system increases as the freedom of motion of its molecules increases. Gas phase molecules have much more freedom of motion than do molecules in the liquid or solid states, so the entropy of a system of gas phase molecules is much greater than that of the same molecules in a condensed state. Molecules in the liquid state have slightly more freedom of motion than those in the solid, so the entropy of a system of molecules in the liquid state is slightly greater than in the solid state. We will use the following conclusion frequently.
Sgas >> Sliquid > Ssolid

9.5-4. Systems Strive to Increase Their Entropy

The two driving forces behind a process are the tendencies for systems to minimize their potential energy
(ΔH < 0)
and to maximize the number of ways in which they can distribute their energy
(ΔS > 0).
However, unless the entropy change is large, the tendency to minimize enthalpy is the dominant driving force. We have seen how to estimate enthalpy changes from bond energies, and we now examine how to predict whether the entropy change is large and positive, large and negative, or negligible.
The entropy change of a reaction is given as
ΔS = SproductsSreactants,
so
ΔS > 0
when the entropy of the products is greater than that of the reactants. At this point, we need only predict whether ΔS is large, and, if so, what is its sign. To do this, we recall that
Sgas >> Sliquid > Ssolid.
Indeed, we will assume that Sliquid ~ Ssolid ~ 0 to conclude the following.
Entropy changes are large only in reactions in which the number of moles of gas changes.
Thus, the sign and magnitude of the entropy change during a reaction can be estimated from the following. ΔS can be calculated, but the procedure is beyond the scope of this course. It is covered in Chapter 4 of Chemistry - A Quantitative Science.

9.5-5. Constraints and Disorder as Predictors of Entropy

Constrained systems are more ordered and have lower entropies than systems that are less constrained.
We have seen that adding constraints to the motion of particles reduces their freedom of motion, which decreases their entropy. In other words, constrained systems have lower entropies than do unconstrained systems. In addition, putting constraints on a system usually results in some ordering of the system. For example, motion is unconstrained in the gas phase, so the entropy of a gas is high. However, when intermolecular forces constrain the molecules, they align to increase their interactions. The effect of such alignments is to produce the short range order found in the liquid state. Increasing the constraints on a liquid forces the molecules to align so as to maximize their interactions and produce the long range order found in the solid state. Thus, the degree to which a system is ordered is a good indicator of its freedom of motion and, therefore, of its entropy. We conclude the following. Indeed, entropy is commonly defined as a measure of the order in a system. However, entropy is a measure of the number of ways in which energy can be distributed in a system; the order of a system is simply a good predictive tool because ordered systems cannot distribute energy in as many ways as disordered ones.

9.5-6. Entropy Exercise

Exercise 9.10:

Predict which system in each pair has the greater entropy.
  • (a)
  • (b) Four atoms can distribute energy in more ways than two atoms.
  • (a)
  • (b) The order of the 12 atoms in (a) indicates that they are being constrained more than those with the random distribution in (b).
  • (a) Each of the four molecules in (a) has 3(3) = 9 degrees of freedom, but each molecule in (b) has only 2(3) = 6 degrees. Therefore, the molecules in (a) can distribute their energy in more ways and have the higher entropy.
  • (b)

9.5-7. Effects of Heat and Temperature on Entropy

The number of ways a system can distribute its energy increases with the amount of energy it has to distribute, so the entropy of a system always increases with temperature. Similarly, adding heat to a system always increases the entropy of a system. However, 1 J of heat has much more impact when it is added to a system that has less energy (lower T) than it does when added to a system that has more energy (higher T). This relationship is expressed in the following equation, which defines entropy change.
( 9.5 )
ΔS =
heat added reversibly
temperature at which the heat is added
=
qrev
T
J/K
qrev is used to indicate that the heat must be added reversibly, which means that it must be added so slowly that the slightest change could reverse the direction of heat flow. Equation 9.5 can be rearranged to
TΔS = qrev,
so TΔS is an energy term.
Using order as a predictor of entropy, we can understand the above with a simple analogy: the effect on the order of a room resulting when a book is casually dropped in it. The probability is very high that the book will NOT drop into its allotted location, so the dropped book will increase the disorder (entropy) of the system. In a neat room (representing a more ordered, lower temperature system), the effect of the out-of-place book is dramatic. However, its effect on a messy room (representing a higher temperature, less ordered system) is negligible as the out-of-place book is hardly noticeable. Similarly, adding 1 J of heat to a solid at 5 K (very ordered, low entropy) has a much more dramatic effect on the entropy than adding 1 J to the vapor at 500 K (disordered, high entropy).

9.5-8. Predicting Entropy Change Example

Exercise 9.11:

Predict the sign of ΔS° for each of the following reactions.
    CaCO3(s) CaO(s) + CO2(g)
  • ΔS < 0
    The reaction produces gas, so ΔS° > 0. The actual value is 159 J/K.
  • ΔS > 0
  • ΔS ~ 0
    The reaction produces gas, so ΔS° > 0. The actual value is 159 J/K.
    3 H2(g) + N2(g) 2 NH3(g)
  • ΔS < 0
  • ΔS > 0
    The balanced equation shows four moles of gas on the right and only two on the left. Consequently, gas is consumed, so ΔS° < 0. The actual value is –199 J/K.
  • ΔS ~ 0
    The balanced equation shows four moles of gas on the right and only two on the left. Consequently, gas is consumed, so ΔS° < 0. The actual value is –199 J/K.
    Ag(s) + NaCl(s) AgCl(s) + Na(s)
  • ΔS < 0
    There are no gases involved in the reaction, so the entropy change is expected to be small, i.e., ΔS° ~ 0. The actual value is 33 J/K.
  • ΔS > 0
    There are no gases involved in the reaction, so the entropy change is expected to be small, i.e., ΔS° ~ 0. The actual value is 33 J/K.
  • ΔS ~ 0
    H2O(l) H2O(g)
  • ΔS < 0
    The reaction produces gas, so ΔS > 0. The actual value is 118 J/K.
  • ΔS > 0
  • ΔS ~ 0
    The reaction produces gas, so ΔS > 0. The actual value is 118 J/K.
    H2O(l) at 25 °C H2O(l) at 5 °C
  • ΔS < 0
  • ΔS > 0
    The entropy of a system always decreases when its temperature is lowered. Consequently, ΔS° < 0, but it is small because the temperature change is small. The actual value is –13 J/K.
  • ΔS ~ 0
    The entropy of a system always decreases when its temperature is lowered. Consequently, ΔS° < 0, but it is small because the temperature change is small. The actual value is –13 J/K.

9.5-9. Entropy Change Example

Exercise 9.12:

Indicate which process in each of the following pairs increases the entropy of the system more.
  • adding 10 J of heat to neon at 300 K Use ΔS = q/T. The heat is added at the same temperature, so the greater entropy change occurs when the greater amount of heat is added. Consequently, adding 50 J results in a greater entropy change than adding 10 J at 300 K.
  • adding 50 J of heat to neon at 300 K
  • adding 10 J of heat to neon at 300 K
  • adding 10 J of heat to neon at 800 K Use ΔS = q/T. The amount of heat added is the same, so the greater entropy change occurs when the heat is added at the lower temperature. Consequently, adding 10 J results in a greater entropy change at 300 K.
  • adding 50 J of heat to neon at 800 K Use ΔS = q/T. (50/800) > (10/300), so adding 50 J at 800 K causes a larger entropy change.
  • adding 10 J of heat to neon at 300 K Use ΔS = q/T. (50/800) > (10/300), so adding 50 J at 800 K causes a larger entropy change.
  • melting 1 g of neon at its melting point Use ΔS = SfinalSinitial.
    • melting: ΔS = SliquidSsolid
    • evaporating: ΔS = SgasSliquid
    Previously we concluded that Ssolid < Sliquid << Sgas, so (SgasSliquid) > (SliquidSsolid). Entropy of evaporation is greater than entropy of melting. As a very good rule, entropy increases for processes that create gases and decreases for processes that consume gases.
  • evaporating 1 g of neon at its boiling point

9.6 Second Law of Thermodynamics

Introduction

The fact that systems seek to increase their entropy is the basis of the Second Law of Thermodynamics, the topic of this chapter.

Objectives

9.6-1. The Second Law of Thermodynamics

Spontaneous processes must increase the entropy of the universe.
We have seen that systems with higher entropies are statistically favored over those with lower entropies, but processes impact the entropy of both the system and its surroundings. In order to determine the spontaneity of a process, we must consider the entropy change that it causes in the universe, not just the system. Indeed, the second law of thermodynamics defines spontaneity in terms of this change. In other words, spontaneous processes are those that increase the entropy of the universe. Processes that reduce the entropy of the system can be accomplished, but only if the entropy of the surroundings is increased more than the entropy of the system is decreased.

9.6-2. Entropy and Heat Flow

Heat flows spontaneously from hot to cold, never cold to hot. (It is forced to flow cold to hot in a heat pump, but energy is required, so it is not spontaneous.) We now show that this is a consequence of the second law. Consider to thermal resevoirs at T1 and T2 that are in thermal contact as shown below.
Figure 9.3
To determine the spontaneous direction of heat flow, we examine the entropy change associated with the flow of heat from T1 to T2. Applying the second law to the problem, we can write the following.
ΔSuniv = ΔS1 + ΔS2 > 0
ΔS1 and ΔS2 are the entropy changes that the heat flow causes in the two reservoirs, which can be obtained by using Equation 9.5
ΔS =
heat added reversibly
temperature at which the heat is added
=
qrev
T
J/K
.
ΔS1 =
q
T1
    ΔS2 =
+q
T2
The negative sign indicates that heat is leaving the reservoir at temperature T1, and the positive sign indicates that heat is entering the reservoir at temperature T2. Substitution of these ΔS values into the ΔSuniv expression yields the following.
ΔSuniv =
q
T1
+
+q
T2
Thus,
ΔSuniv > 0
and the process is spontaneous only if
T1 > T2.
In other words, only if the heat flows from the hotter sources to a cooler one.

9.7 Free Energy

Introduction

We now show that ΔSuniv can be written as a sum of system quantities, and, in doing so, derive a new thermodynamic property that includes the effects of both ΔH and ΔS on the spontaneity of any process carried out at constant temperature and pressure. The new property is called the free energy because it is the energy that is free to do work.

Objectives

9.7-1. Free Energy

A spontaneous process at constant T and P is one in which
ΔG < 0.
Entropy changes in the universe can be broken down into the changes in the system and its surroundings.
ΔSuniv = ΔS + ΔSsur
During the process, heat can be exchanged between the system and the surroundings, which changes the entropy of the surroundings. For a reaction carried out at constant T and P, the amount of heat that is transferred from the system to the surroundings is
qsur = − ΔH,
so
ΔSsur = −ΔH/T.
An exothermic reaction releases heat into the surroundings, which increases the entropy of the surroundings. The entropy change in the universe resulting from a reaction carried out at constant temperature and pressure is the following.
ΔSuniv = ΔSΔH/T
Multiplying both sides by –T, we obtain the following.
TΔSuniv = ΔHTΔS
The above expression for ΔSuniv is a function of system quantities only, so it too is a system quantity. This new thermodynamic function of the system is called the Gibbs free energy, ΔG or simply the free energy. Thus, the driving forces behind a process carried out at constant T and P are combined into the Gibbs free energy expression.
( 9.6 )
ΔG = ΔHTΔS
Gibbs Free Energy

9.7-2. Enthalpy and Free Energy

Equation 9.6 indicates that processes tend to minimize their potential energy
(ΔH < 0).
relationship of products and reactants in endothermic and exothermic process
Figure 9.4: Enthalpy and Free Energy

9.7-3. Entropy and Free Energy

Equation 9.6
ΔG = ΔHTΔS
indicates that processes tend maximize their entropy
(ΔS > 0).
relationship of entropy to the spontaneousness of a reaction
Figure 9.5: Entropy and Free Energy

9.7-4. Free Energy and Reaction Spontaneity

There are driving forces in both the forward and reverse directions of most reactions, so double arrows are often used in chemical equations. A equilibrium arrow B shows that there are driving forces for both A B and A B. ΔG is the difference between these forces. If
ΔG < 0,
the forward driving force is greater, so there is a net force in the forward direction and the reaction consumes A to produce B. When
ΔG > 0,
the reverse driving force is greater, so there is a net force in the reverse direction, and the reaction produces A by consuming B. If
ΔG = 0,
the two forces are equal, and there is no net driving force. At this point there is no longer a change in concentrations as the reaction has reached equilibrium. Indeed,
ΔG = 0
is the thermodynamic definition of equilibrium at constant temperature and pressure. As summarized in the Table 9.5, the following is true.
The sign of ΔG indicates the spontaneous direction of reaction.
The value of ΔG varies with the concentrations of the reactants and products, so it changes as the process continues. Consider the evaporation of water: H2O(l) equilibrium arrow H2O(g). Initially, only liquid is present, so there is no driving force in the reverse direction (no condensation). Consequently,
ΔG < 0
and the liquid begins to evaporate. However, as the pressure of the vapor increases, so does the driving force in the reverse direction, so ΔG gets less negative. Eventually, the pressure of the vapor is such that the driving force in the reverse direction equals that in the forward direction,
ΔG = 0,
and the process has reached equilibrium.
Table 9.4: The Sign of the Free Energy and the Direction of Reaction

9.8 Standard Free Energy and the Extent of Reaction

Introduction

The difference between the free energy of a reaction (ΔG) and the standard free energy of reaction (ΔG°) is an important one and one that is often confused by students. In this section, we examine the meaning of these two important thermodynamic properties.

Objectives

9.8-1. Standard Free Energy

As is the case with enthalpies, entropies and free energies of reaction are usually given for the reaction carried out under standard conditions. The standard free energy is defined in terms of the standard enthalpy and entropy of reaction as the following.
ΔG° = ΔH° − TΔS°
Recall that the superscript zero means that all reactants and products are in their standard states at a specified temperature, so ΔG° is a constant for a reaction at that temperature.
Exercise 9.13:

Predict the signs of ΔH° and ΔS°. Then use ΔG° = ΔH° – TΔS° to predict the temperature conditions under which ΔG° is likely to be negative.
(a)    H2(g) + Cl2(g) 2 HCl(g)    ΔH° = –180 kJ/mol
    What is the sign of ΔH°?
  • + ΔH° = –180 kJ/mol
    What is the sign of ΔS°?
  • + The number of moles of gas is the same on each side of the equation.
  • The number of moles of gas is the same on each side of the equation.
  • ~ 0
    ΔG° is most probably negative at which of the following?
  • all temperatures ΔS° ~ 0, so it has little effect. ΔH° << 0, so ΔG° << 0 and the reaction is extensive at all temperatures.
  • no temperatures ΔS° ~ 0, so it has little effect, and ΔH° << 0.
  • low temperatures only ΔS° ~ 0, so it has little effect, and ΔH° << 0.
  • high temperatures only ΔS° ~ 0, so it has little effect, and ΔH° << 0.
(b)    CCl4(l) CCl4(g)
    What is the sign of ΔH°?
  • +
  • Evaporation is always endothermic because the energy of the molecules in the gas is greater than that of those in the liquid.
    What is the sign of ΔS°?
  • +
  • The evaporation produces gas.
  • ~ 0 The evaporation produces gas.
    ΔG° is most probably negative at which of the following?
  • all temperatures ΔS° > 0, which is good. ΔH° > 0, which is bad.
  • no temperatures ΔS° > 0, which is good. ΔH° > 0, which is bad.
  • low temperatures only ΔS° > 0, which is good. ΔH° > 0, which is bad.
  • high temperatures only ΔS° > 0, which is good. ΔH° > 0, which is bad. Thus, the reaction is most extensive at high temperatures where the TΔS term can dominate and make ΔG° < 0, which makes the reaction extensive.
(c)    P4(s) + 6 Cl2(g) 4 PCl3(g)     ΔH < 0
    What is the sign of ΔH°?
  • + The problem states that ΔH° is negative.
    What is the sign of ΔS°?
  • + The reaction consumes gas (six moles go to four moles).
  • The reaction consumes gas (six moles go to four moles), so ΔS° < 0.
  • ~ 0 The reaction consumes gas (six moles go to four moles).
    ΔG° is most probably negative at which of the following?
  • all temperatures ΔS° < 0, which is bad. ΔH° < 0, which is good. Thus, the TΔS° term must be minimized.
  • no temperatures ΔS° < 0, which is bad. ΔH° < 0, which is good. Thus, the TΔS° term must be minimized.
  • low temperatures only ΔS° < 0, which is bad. ΔH° < 0, which is good. Thus, the TΔS° term must be minimized. At low temperature, the ΔH° term will dominate making ΔG° negative and the reaction extensive.
  • high temperatures only ΔS° < 0, which is bad. ΔH° < 0, which is good. Thus, the TΔS° term must be minimized.

9.8-2. Extent of Reaction

ΔG° is the value of ΔG when all reactants and products are in their standard states. Consider the process:
A(g) equilibrium arrow B(g). If ΔG° < 0, the reaction is spontaneous in the forward direction when both pressures are 1 atm, so equilibrium is attained by consuming A, which reduces PA to less than 1 atm, and producing B, which increases PB to more than 1 atm. If
PB > PA
at equilibrium, the reaction is said to be extensive because the amount of product at equilibrium is much greater than the amount of reactant. If ΔG° > 0, the process is spontaneous in the reverse direction at standard conditions. Equilibrium in this case is attained when
PA > PB,
so there is more reactant than product at equilibrium and the reaction A equilibrium B is not extensive. We conclude the following.
ΔG° indicates the extent of reaction.

9.8-3. Extent Exercise

Exercise 9.14:

One mole each of NH3 and N2H2 are allowed to react in identical 1-L containers at 298 K. Which flask would contain the greater number of moles of H2 when the reactions are complete?

You should read the replies to both answers to better understand the role of ΔG°.
  • 2 NH3 N2 + 3 H2    ΔG° = +34 kJ Stoichiometry predicts that 1.5 mol H2 forms for each mole of NH3 that reacts, but ΔG° > 0 for this reaction, so it is not extensive. Consequently, very little ammonia decomposes (less than 1%) at 298 K and very little H2 is produced by this reaction at this temperature.
  • N2H2 N2 + 2 H2    ΔG° = –160 kJ Stoichiometry predicts that 2 mol of H2 should form for each mole of N2H4 that reacts, but the amount that reacts is dictated by the extent of reaction. Since ΔG° is large and negative, we conclude that the reaction is extensive and that essentially all of the N2H4 does react. Consequently, this flask would contain close to 2 moles of H2 at the end of the reaction.

9.8-4. Extent of Reaction Versus Temperature

ΔG° = ΔH° – TΔS° can be re-arranged to give the common form of a straight line
(y = mx + b)
as follows.
y = m x + b
ΔG° = Δ T + ΔH°
Thus, a plot of ΔG° versus T is a straight line with a slope of –ΔS° and an intercept of ΔH°. The intercept is ΔH°, so ΔG° = ΔH° at T = 0. Indeed, we will assume that the sign of ΔG° is the same as the sign of ΔH° at low temperatures. However, the slope of the line is –ΔS°, so ΔS° can dictate the sign of ΔG° at very high temperatures. For example, consider the temperature dependence of ΔG° of reactions A - E as described in Table 9.5 and Figure 9.6.
Table 9.5: Standard Free Energy Versus Temperature
Figure 9.6: Standard Free Energy Versus Temperature

9.8-5. Summary

At very low temperatures, the TΔS° term is negligible, and ΔG° has the sign of ΔH°. At very high temperatures, the TΔS° term dominates (unless ΔS° ~ 0), and ΔG° has the sign of –ΔS°.
Table 9.6: Predicting the Extent of Reaction

9.9 Activation Energy

Introduction

Thermodynamics compares the initial and final states only. It does not consider how the transition from one state to another is accomplished; that is the domain of kinetics. Energy considerations are also an important part of kinetics, and we use the one-step displacement of iodide by hydroxide to examine reaction energetics in this section.

Objectives

9.9-1. The Reaction

In this section we examine both the thermodynamics and kinetics of the following one-step reaction in which iodide is displaced by a hydroxide ion.Thermodynamics is concerned only with differences between the initial and final states of the reaction, while kinetics is concerned with how the reaction proceeds.
reaction
Figure 9.7

9.9-2. Thermodynamic View

Thermodynamics is concerned only with the difference between the initial and final states.
Consider the following reaction.
The properties of interest are ΔH°, ΔS°, and ΔG°. Although all of these properties can be determined from tables, our goal is to estimate their sign and whether they are large or small.
potential energy of a reaction
Table 9.7

9.9-3. Kinetic View

Reaction Diagram and Transition State

  • Viewing the Video
  • View the video in this window by selecting the play button.
  • Use the video controls to view the video in full screen.
  • View the video in text format by scrolling down.
Kinetics deals with the transition from reactants to products.
Kinetics is the study of the rates and mechanisms of reactions. In other words, kinetics looks at the transition from the initial to the final state. To do this, we define a reaction coordinate, which is normally a complicated combination of intermolecular distances and molecular structural changes. In order to form the C-O bond, the OH1– ion must attack the carbon atom along the line through the center of the plane formed by the three H atoms. As the C-O distance decreases, the three H-C-H angles all increase from their original 109° as the H atoms separate, and the C-I bond elongates. Thus, the reaction coordinate involves changes in all of these parameters. As the C-O distance decreases, the H-C-H bond angles continue to increase, and the C-I bond continues to lengthen. The system reaches an energy maximum when the H-C-H bond angles are 120° and the CH3 group is planar. The species that is formed at this point is called the transition state, which is the species that the reactants must pass through in their transition to the products. It is at the energy maximum, so it immediately changes into one of the species at lower energy. If the C-O bond shortens and the C-I bond breaks, the transition state leads to the products, but if the C-O bond breaks and the C-I bond shortens, the transition state leads to the reactants. Either action is possible.
Reactants must go through a transition state to reach the products, and the energy required to reach the transition state is called the activation energy.
The energy required to reach the transition state is called the activation energy (Ea). Once the products form, they can also collide to form the same transition state. The reaction between the substances on the right side of the chemical equation is called the reverse reaction, while reaction of the substances on the left side is called the forward reaction. Both reactions have activation energies. Ea(f) is the activation energy for the forward reaction, while Ea(r) is the activation energy for the reverse reaction. Activation energies are always positive. Note that
ΔH = Ea(f) − Ea(r),
so
Ea(f) < Ea(r)
for exothermic reactions, but
Ea(f) > Ea(r)
for endothermic reactions.
reaction energy diagram
Figure 9.8: Reaction Diagram for CH3I + OH = CH3OH + I1–

9.10 Rates of Reaction and The Rate Law

Introduction

In this section, we examine the factors that dictate the rate of a reaction, and then pull all of the factors together into a rate law for the reaction.

Prerequisites

Objectives

9.10-1. Collision Frequency

The frequency of collisions between two particles is proportional to the product of their concentrations.
We now turn our attention to the rate of reaction, i.e., how fast the concentrations change with time. In order for two molecules to react, they must collide. Thus, the rate of reaction is proportional to the frequency at which the reactants collide. The collision frequency is the number of collisions per unit volume per unit time, which normally has units of (moles of collisions) · liter–1 · s–1. The frequency of collisions between particles is proportional to the product of the molar concentrations of the colliding particles. Recall that the molar concentration of particle A is the number of moles of A in a liter and given the symbol [A]. [A] = 1 M is read "the concentration of A is 1 molar or 1 mole per liter." In the following examples, C is a proportionality constant that differs for each reaction and increases with temperature.
Table 9.8

9.10-2. Orientation Effects

Only a fraction of colliding particles are oriented properly for reaction.
Only a fraction of the collisions between reactants results in the transition state. In fact, the reaction rate can be very small or even zero when the collision frequency between reactants is high if the fraction of collisions, X, that lead to the transition state is very small. The fraction depends upon both the energy and the orientation of the colliding molecules. We consider orientation effects here and energy in the following section. Consider the relative orientations of reactants shown in Figure 9.8. The orientations shown in (a) and (b) do not lead to the transition state, while orientation (c) does. There are many more ways in which the molecules can collide with an orientation that does not lead to the transitions, so the fraction that does lead to the transition state is fairly low in this reaction.
different orientations of reactants
Figure 9.8: Orientation Effect
(a) OH1– ion attacks the iodine atom not the carbon, so transition state cannot form; (b) OH1– ion attacks the correct position, but it is turned so that a C-O bond cannot form; (c) the only orientation that can lead to the transition state.

9.10-3. Energy Effect

Only a fraction of colliding particles have sufficient energy to react.
A collision must have enough energy to overcome the activation energy of the reaction. A collision between OH1–and CH3I must have enough energy to push the H atoms back and lengthen the C-I bond. The average energy of a collision depends upon the thermal energy (~RT). Consequently, the fraction of collisions leading to the transition state always increases with temperature.

9.10-4. Rate Law

Many reactions involve several steps to convert the reactants to the products. For example, the reaction between hydrogen and oxygen is a complicated reaction that involves many steps. However, some reactions, such as the displacement reaction between CH3I and OH1–, occur in a single collision (step). Consider the rate of the reaction A + B products. The rate of a reaction equals the fraction of collisions (X) that lead to the transition state times the collision frequency (F), so we may write the following.
rate = XF
The collision frequency is proportional to the concentrations of the reactants, or the following.
F = C[A][B]
Combining these two equations, we note that the rate of the reaction equals the product of two constants, which is also a constant k = XC, times the products of the reactant concentrations.
rate = XC[A][B] = k[A][B]
This expression is the rate law of the reaction, and k is the rate constant of the reaction. X and C are each temperature dependent, so k is also temperature dependent.
The rate law for a simple, one-step reaction is equal to the rate constant for the reaction times the product of the concentrations of the reactants each raised to an exponent equal to their coefficient in the balanced chemical equation.

9.10-5. Rate Law Exercise

Exercise 9.15:

What are the rate laws for the following reactions, which all occur in a single step?
O3 + O 2 O2
rate = o_k[O_3][O]_s rate = k[O3][O]
2 NOCl 2 NO + Cl2
rate = o_k[NOCl]^2_s rate =k[NOCl]2
2 CO + O2 2 CO2
rate = o_k[CO]^2[O_2]_s rate = k[CO]2[O2]
H2 2 H
rate = o_k[H_2]_s rate = k[H2]

9.10-6. Catalysts

A catalyst increases the speed of a reaction by lowering its activation energy.
The rate law for the simple reaction A + B C is rate = k[A][B]. Consequently, the rate of reaction can be changed by changing the concentrations: increasing the reactant concentrations increases the rate of reaction. Indeed, as a reaction proceeds, the rate of reaction slows because the reactants are consumed. The rate of reaction can also be changed by changing the rate constant. The rate constant can be increased in three ways. The last two methods of increasing the rate constant are accomplished with the use of catalysts. Catalysts speed reactions by assuring that the reacting molecules align correctly for reaction and/or by providing a pathway with a lower activation energy. Enzymes are proteins that bind to reacting molecules (substrates). The binding weakens the reactant bonds, which lowers the activation energy, and they can force close alignment of two reactants, which assures the correct orientation.

9.10-7. Ozone Depletion

As an example of catalytic behavior, we consider the problem of ozone depletion in the upper atmosphere. O3 in the upper atmosphere protects us from ultraviolet radiation from the sun. It is removed by reaction by atomic oxygen.
O + O3 2 O2     Ea = 17 kJ/mol
Fortunately, the activation energy is so high that the reaction is very slow at the temperature of the upper atmosphere. However, chlorine atoms can catalyze the reaction. The chlorine atoms are produced when high energy light from the sun is absorbed by chlorofluorocarbons (CFC's), which were common in refrigeration and spray cans.
CF2Cl2 + hv CF2Cl + Cl
The chlorine atoms catalyze the depletion of ozone in two steps. Their chemical equations and activation energies follow. Summing the two steps yields the same ozone depletion reaction that occurs in the absence of chlorine atoms
(O + O3 2 O2).
The chlorine atoms speed the reaction but are unchanged by it, so they are a catalyst for the reaction. They provide a different path with a lower activation energy (the activation energies for the two steps are 2 and 0.4 kJ/mol, respectively), so the reaction is much faster in the presence of chlorine.
reaction diagram for ozone depletion
Figure 9.9: Reaction Diagram for Ozone Depletion
Shown is the reaction diagram for the ozone depletion reaction in the absence (black curve) and presence (red curve) of chlorine atoms.

9.11 Equilibrium and the Equilibrium Constant

Introduction

Although it is common to say that a reaction has stopped, reactions don't really stop. Instead, they reach a dynamic equilibrium in which the molecules on the right side of the chemical equation continue to produce those on the left and vice versa. At equilibrium, however, the amounts of products and reactants are no longer changing because they are being consumed and produced at the same rate. In this section, we discuss the equilibrium process from the viewpoints of both kinetics and thermodynamics and introduce the equilibrium constant.

Objectives

9.11-1. Equilibrium and the Equilibrium Constant

A reaction reaches equilibrium when the rates of the forward and reverse reactions are equal.
As a reaction proceeds, the concentrations of the reactant molecules decrease while those of the product molecules increase. Consequently, the rate of the forward reaction decreases, while that of the reverse reaction increases. When the two rates are equal, reactants and products are formed at the same rate that they are consumed and the reaction reaches equilibrium. Chemical equilibria are dynamic because reaction continues, even though the concentrations of the reactants and products no longer change. Double or equilibrium arrows (equilibrium arrow) are used in reactions at equilibrium to emphasize that the reaction takes place in both the forward and reverse directions. Consider the reaction: CH3OH1– equilibrium arrow I1–CH3OH.
Reaction Rate
Forward Reaction CH3I + OH1– I1– + CH3OH Rf = kf[CH3I][OH1– ]
Reverse Reaction I1– + CH3OH CH3I + OH1– Rf = kf[CH3OH][I1– ]

At equilibrium,
Rf = Rr,
so we can write the following.
kf[CH3I][OH1−] = kr[CH3OH][I1−]
Rearranging the above to get the rate constants on one side and the concentrations on the other, we obtain the following.
kf
kr
=
[CH3OH][I1−]
[CH3I][OH1−]
The ratio of concentrations equals the ratio of rate constants, which is a constant at a given temperature. Thus, the above expression is a constant, which is called the equilibrium constant for the reaction and given the symbol K.
K =
[CH3OH][I1−]
[CH3I][OH1−]
The equilibrium constant expression equals the equilibrium concentrations of the substances on the right side of the chemical equation (the products) divided by the equilibrium concentrations of the substances on the left side of the chemical equation (the reactants). The concentration of each substance is raised to an exponent equal to the coefficient of the substance in the balanced chemical equation.

9.11-2. Some Examples of Equilibrium Constant Expressions

The concentration of a gas in an equilibrium constant expression equals its partial pressure in atmospheres, while the concentration of a solute is given by its molar concentration. Solids and liquids enter as unity (1).
The form of the 'concentration' of a substance in the equilibrium constant expression depends upon the state of the substance. For example, consider the following equilibrium.
aA(g) + bB(aq) equilibrium arrow xX(s) + yY(aq)

K =
1x[Y]y
PaA[B]b
=
[Y]y
PaA[B]b
The 'concentration' of X appears as 1 because X is a solid; the 'concentration' of A is its partial pressure because A is a gas; and the 'concentrations' of B and Y are their molar concentrations because they are solutes in an aqueous solution. The following examples should help.
Table 9.9: Some Examples of Equilibrium Constant Expressions
As shown in the following exercise, the equilibrium constant for a reaction can be determined from known equilibrium concentrations or it can be used with other known equilibrium concentrations to determine an unknown concentration.

9.11-3. K Calculation Exercise

Exercise 9.16:

Use the following chemical equation to answer the questions:
N2O4 equilibrium arrow 2 NO2.
Calculate the value of K at a temperature where the following equilibrium concentrations are found: PNO2 = 0.12 atm and PN2O4 = 0.075 atm.
K = 0.192___
K =
P2NO2
PN2O4
= 0.19
What is the equilibrium pressure of N2O4 if the equilibrium pressure of NO2 is 0.36 atm and K = 0.19?
0.6821___
K =
P2NO2
PN2O4
PN2O4 =
P2NO2
K
=
0.362
0.19
= 0.68 atm

9.11-4. Extensive Reactions

An extensive reaction is one in which K >> 1; that is, one in which the equilibrium concentrations of the products exceed those of the reactants.
Determining how much reactant should be used in a reaction to produce a given amount of product is a common calculation in chemistry. In such calculations, it is often assumed that one of the reactants, called the limiting reactant, disappears completely, but this is the case only when K is very large. When the equilibrium constant is not very large, there will be a substantial amount of at least some reactants at equilibrium, so the amounts can be determined only if the value of the equilibrium constant is known. A reaction with a very large equilibrium constant is said to be an extensive reaction. The value of K is large when at least one of the concentrations in the denominator is very small. Thus, almost no limiting reactant remains at equilibrium in an extensive reaction, and the approximation that it disappears completely is valid.
Exercise 9.17:

Indicate whether the products, reactants, or neither would dominate the equilibrium mixture formed by adding equal amounts of all reactants in each of the following reactions.
    H2(g) + I2(g) equilibrium arrow 2 HI(g)     K = 600
  • products
  • reactants K >> 1, so the products (numerator) dominate in the equilibrium mixture. That is, the equilibrium concentration of HI will be much greater than that of H2 or I2.
  • neither K >> 1, so the products (numerator) dominate in the equilibrium mixture. That is, the equilibrium concentration of HI will be much greater than that of H2 or I2.
    HF(aq) + H2O(l) equilibrium arrow F1−(aq) + H3O1+(aq)     K = 7 × 10−4
  • products K << 1, so the reactants (denominator) dominate in the equilibrium mixture. That is, the equilibrium concentration of HF will be much greater than that of F1– or H3O1+.
  • reactants
  • neither K << 1, so the reactants (denominator) dominate in the equilibrium mixture. That is, the equilibrium concentration of HF will be much greater than that of F1– or H3O1+.
    HCN(aq) + NH3(aq) equilibrium arrow CN1−(aq) + NH41+(aq)     K = 0.7
  • products K ~ 1, so neither side dominates. That is, the reactants and products will have similar concentrations in the equilibrium mixture.
  • reactants K ~ 1, so neither side dominates. That is, the reactants and products will have similar concentrations in the equilibrium mixture.
  • neither

9.11-5. Thermodynamic Definition of K

Thermodynamic Basis for Equilibrium

There is a driving force for the forward reaction that depends upon the concentrations of the substances on the left side of the chemical equation and one for the reverse direction that depends upon the concentrations of those on the right side. The difference between these two driving forces is ΔG. Thus, there are three possibilities. Thus,
ΔG = 0
is the criterion for equilibrium, but it is the value of ΔG° that indicates where the equilibrium will be established. Indeed, the equilibrium constant for a reaction is related to the standard free energy of the following reaction.
ΔG° = –RT ln K.

9.11-6. Predicting Extent of Reactions

The extent of reaction is determined by ΔG°, but ΔG° is related to the standard enthalpy and entropy changes by the following.
ΔG° = ΔH° – TΔS°
By analogy to the discussion in Section 9.7, we can draw the following conclusions about the extent of a reaction.
Table 9.10

9.11-7. Predicting Extent of Reactions Exercise

Exercise 9.18:

Use the chemical equation and the equilibrium composition to determine the sign of ΔG°. Use the change in moles of gas to determine the sign of ΔS°. Determine the sign of ΔH° from the other two.
2 HCl(g) equilibrium arrow H2(g) + Cl2(g);     Equilibrium: [H2] = [Cl2] << [HCl]
    ΔG°
  • +
    The concentrations of the products are much less than that of the reactants, so this reaction is not extensive at this temperature. Consequently, ΔG° is positive.
  • The concentrations of the products are much less than that of the reactants, so this reaction is not extensive at this temperature.
  • ~ 0 The concentrations of reactants and products are very different, so ΔG° cannot be close to zero.
    ΔS°
  • +
    The balanced equation does not indicate that the number of moles of gas on the product side is greater than on the reactant side.
  • The balanced equation does not indicate that the number of moles of gas on the product side is less than on the reactant side.
  • ~ 0 The balanced equation indicates that the number of moles of gas is the same on both sides of the equations, so ΔS° is small.
    ΔH°
  • +
    ΔG° is positive and ΔS° is very small, so ΔH° must also be positive.
  • ~ 0
HF(aq) + NO21–(aq) equilibrium arrow F1–(aq) + HNO2(aq);      Equilibrium: [F1–] ~ [HNO2] ~ [HF] ~ [NO21–]
    ΔG°
  • +
    The concentrations of the reactants and products are very similar, so ΔG° cannot be a large positive number.
  • The concentrations of the reactants and products are very similar, so this is not an extensive reaction.
  • ~ 0 All concentrations are similar, so K ~ 1 and ΔG° must be close to zero.
    ΔS°
  • +
    There are no gases involved, so entropy changes are not expected to be large.
  • There are no gases involved, so entropy changes are not expected to be large.
  • ~ 0 There are no gases involved, so entropy changes are not expected to be large. Thus, ΔS° ~ 0.
    ΔH°
  • +
    ΔG° ~ 0 and ΔS° ~ 0, so ΔH° must also be small.
  • ~ 0 ΔG° ~ 0 and ΔS ~ 0, so ΔH° must also be small.

N2(g) + 3 H2(g) equilibrium arrow 2 NH3(g);      Equilibrium: [N2] ~ [H2] << [NH3]
    ΔG°
  • +
    The concentration of the product is much greater than those of the reactants, so ΔG° cannot be large and positive.
  • The concentration of the product is much greater than those of the reactants, so ΔG° is negative for this reaction at this temperature.
  • ~ 0 The concentration of the product is much greater than those of the reactants, so ΔG° cannot be very small.
    ΔS°
  • +
    Gas is consumed by the reaction, so ΔS° cannot be positive.
  • Two moles of gas are formed for every four that react. Gas is consumed by the reaction, so ΔS° is negative.
  • ~ 0 Gas is consumed by the reaction, so ΔS° is not small.
    ΔH°
  • +
  • ΔG° < 0 and ΔS° < 0, so ΔH° must also be negative.
  • ~ 0

9.11-8. Effect of T on K

The value of an equilibrium constant varies with temperature. Changes in temperature change the thermal energy of the molecules, which changes the fraction of molecules at higher energy. The fraction of molecules at higher energy increases when the temperature is increased.

9.11-9. Example

Exercise 9.19:

Assume that all reactants and products in the following are gases to answer the questions.
line graph, energy vs reaction coordinate
    The equilibrium concentration of A-B is greater at which of the following?
  • high T
  • low T A-B is a product, which is at lower energy in exothermic reaction. Substances are low energy are favored at low T. Lowering the temperature increases the concentrations of the reactants and decreases the concentrations of the products, which has the effect of decreasing the value of the equilibrium constant. This reaction becomes less extensive as the temperature increases because it is exothermic.
    Which is the stronger bond?
  • A-A
  • A-B ΔH < 0, but ΔH ~ (A-A bond energy) - (A-B bond energy). Consequently, A-B bond energy is greater than the A-A bond energy.
    Is the reaction extensive?
  • Yes ΔH° < 0, and the number of gaseous molecules is unchanged during the reaction, so ΔS° ~ 0. Consequently, ΔG° < 0 and the reaction is extensive.
  • No
    Which reaction has the higher rate constant?
  • forward If the orientation factor is assumed to be similar for the two reactions, the one with the lower activation energy has the larger rate constant. The activation energy for the forward reaction (A-A + B A-B + A) is less than that for the reverse reaction (A-B + A A-A + B), so the rate constant for forward reaction is larger.
  • reverse

9.12 Le Châtelier's Principle

Introduction

An open door that is not moving is at equilibrium. However, the equilibrium position of the door can be changed by applying pressure to either side. Reactions at equilibrium react in a similar fashion when they are exposed to a chemical stress.

Objectives

9.12-1. Le Châtelier's Principle

Equilibrium mixtures respond to the addition or removal of material in a manner consistent with the following principle. Adding or removing a pure liquid or solid does not affect the equilibrium mixture because the concentration of the pure substance is unchanged. However, the concentrations of all solutes are changed when solvent is removed.
Exercise 9.20:

Consider the following equilibrium.
AgCl(s) + 2 NH3(aq) equilibrium arrow Ag(NH3)21+(aq) + Cl1−(aq)     ΔH < 0
Indicate whether each of the following actions would increase, decrease, or not affect the equilibrium concentration of Cl1–
    adding Ag(NH3)21+
  • increase
  • decrease Equilibrium must shift so as to remove some of the additional Ag(NH3)21+ by reacting it with Cl1–. Therefore, the concentration of Cl1– will decrease.
  • not affect
    removing Ag(NH3)21+
  • increase Equilibrium must shift so as to replace some of the lost Ag(NH3)21+ by reacting more AgCl and NH3, which increases the concentration of Cl1–.
  • decrease
  • not affect
    adding solid AgCl
  • increase
  • decrease
  • not affect Adding substances that appears as pure solids or liquids in the equilibrium reaction does not change their concentrations, so it has no effect on the other concentrations.
    adding NH3
  • increase The equilibrium removes some of the additional ammonia by reacting it with AgCl to produce more Ag(NH3)21+(aq) and Cl1–. Consequently, the equilibrium concentration of Cl1– will increase.
  • decrease
  • not affect
    increasing temperature
  • increase
  • decrease Changing the temperature is the only change that results in a change of K. Recall that an increase in temperature increases the concentrations of those substances at higher energy, which are the reactants (substances on the left of the chemical equation) in an exothermic reaction (ΔH < 0). Thus, the concentration of chloride ion, which is on the right side of the equation and at lower energy, decreases as the temperature increases.

    Alternatively, we can use LeChâtelier's principle by realizing that increasing the temperature is the same as adding heat, which causes the substances to react so as to remove some of the heat. Heat is a product of an exothermic reaction, so the concentrations of all products (substances on the right side of the equation) decrease, while the concentrations of the reactants all increase. Consequently, the concentration of chloride ion (a product) decreases.
  • not affect

9.13 Exercises and Solutions

Select the links to view either the end-of-chapter exercises or the solutions to the odd exercises.