Print

Chapter 11 – Electron Transfer Reactions and Electrochemistry

Introduction

Redox, or electron transfer, reactions constitute one of the broadest and most important classes of reactions in chemistry. All reactions that involve molecular oxygen, such as combustion and corrosion, are electron transfer reactions. Biological processes, such as respiration, photosynthesis, and the breakdown of food molecules, consist of sequences of electron transfer reactions that serve to transport and utilize energy from the sun. Batteries are devices that allow us to utilize the free energy of electron transfer reactions.

11.1 Electron Transfer or Redox Reactions

Introduction

We begin our study of electron transfer reactions by introducing some terms and definitions and examining the electron transfer process.

Prerequisites

Objectives

11.1-1. Electron Transfer Introduction Video

  • 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.

11.1-2. Electron Transfer

Electrons move from one species to another in electron transfer reactions.
The reaction that occurs when iron (steel wool) is placed in a solution of CuSO4 is shown in the table below.

image of steel wool and solution of CuSO4 before reaction image of steel wool and solution of CuSO4 after reaction
Steel wool consists primarily of Fe atoms. Cu2+ ions gives a CuSO4 solution its blue color. The steel wool is coated with metallic Cu where it was immersed in the CuSO4 solution. The solution loses color because the Cu2+ have been displaced by colorless Fe2+ ions.
Table 11.1: An Electron Transfer Reaction
We can make the following observations: and draw the following conclusions: In the reaction, the oxidation state of copper changes from +2 in the Cu2+ ions in solution to 0 in the atoms comprising metallic copper. Each Cu2+ ion must gain two electrons to become a Cu atom. Similarly, the oxidation state of iron changes from 0 in the atoms comprising the steel wool to +2 in the Fe2+ ions in solution. Each Fe atom must lose two electrons to be converted to an Fe2+ ion. Thus, each Fe atom gives up two electrons, while each Cu2+ gains two electrons, i.e., two electrons are transferred from iron atoms to Cu2+ ions in solution. This is an example of an electron transfer reaction. The reaction is written as
Cu2+ + Fe Cu + Fe2+

11.1-3. Oxidation and Reduction

Reduction is a gain of electrons, oxidation is a loss of electrons, and electron transfer reactions are also called redox reactions.
Reduction is a gain of electrons. The added electrons "reduce" the oxidation state of the substance. Cu2+ ions gain two electrons, so they are reduced to Cu atoms. Note that the two-electron reduction lowers the oxidation state of copper from +2 in the ion to 0 in the atom.
Figure 11.1
Oxidation is a loss of electrons. The loss of negative charge causes an increase in oxidation state of the substance. Fe atoms lose two electrons, so they are oxidized to Fe2+ ions. Note that the two-electron oxidation raises the oxidation state of iron from 0 in the atom to +2 in the ion. Redox reactions are those that involve oxidation and reduction. Electron transfer reactions always involve both an oxidation and a reduction because electrons cannot be gained if none are lost.

11.1-4. Oxidizing and Reducing Agents

Oxidizing reagents (oxidants) are reduced when reducing agents are oxidized.
Electron transfer results from a combination of oxidation and reduction. A species cannot be oxidized unless another species accepts the electrons and is reduced. That is, oxidation causes reduction and vice versa. Consequently, the species that is oxidized by the reaction is referred to as the reducing agent or reductant, and the species that is reduced by the reaction is called the oxidizing agent or oxidant. The reducing agent contains the electrons that are transferred during the reaction, so it is in its reduced form, which we will designate Red1. Transferring the electrons converts it to its oxidized form, which we will call Ox1. Similarly, the oxidizing agent has unfilled orbitals that can accept the transferred electrons, so it is in its oxidized form, Ox2. Accepting the electrons converts it to its reduced form, Red2. A typical redox reaction can be expressed as follows.
Red1 + Ox2 Ox1 + Red2
Thus, the reducing agent, Red1, can be identified as the reduced form (form in lower oxidation state) of species 1, while the oxidizing agent, Ox2 is the oxidized form (form in higher oxidation state) of species 2.

11.1-5. Requirements for Reducing and Oxidizing Agents

The oxidation state of an atom should be high if it is to be an oxidizing agent and low if it is to be a reducing agent.
In order to function as a reducing agent, a substance must be able to give up electrons and attain a higher oxidation state, so reducing agents must contain atoms that can be oxidized. Similarly, oxidizing agents must be able to accept electrons to attain a lower oxidation state, so oxidizing agents must contain atoms that can be reduced. For example, the nitrogen atom in NH3 has an oxidation state of –3, which is the lowest oxidation state that nitrogen can have. Therefore, NH3 can be oxidized but not reduced, so it can be a reducing agent. The nitrogen atom in NO31– is in the +5 oxidation state, the highest it can have, so nitrate ion cannot be oxidized or serve as a reducing agent, but it can be reduced and function as an oxidizing agent.
Exercise 11.1:

Indicate whether each of the following could function only as a reducing agent, only as an oxidizing agent, or as both.
    Na
  • reducing Sodium cannot have a negative oxidation state, so it can only be oxidized and function as a reducing agent. Indeed, Na is a strong reducing agent.
  • oxidizing Metals do not achieve negative oxidation states, so sodium cannot be reduced and cannot function as an oxidizing agent.
  • both Metals do not achieve negative oxidation states, so sodium cannot be reduced and cannot function as an oxidizing agent.
    Br2
  • reducing Br2 can be oxidized to several oxidation states, as in the BrO21– ion, so it can function as a reducing agent. However, bromine atoms are in Group 7A, so they can attain any oxidation state between –1 and +7.
  • oxidizing Br2 can be reduced to Br1– ion, so it can function as an oxidizing agent. However, bromine atoms are in Group 7A, so they can attain any oxidation state between –1 and +7.
  • both Br2 can be reduced to Br1– ion, or it can be oxidized to one of several oxidation states, as in the BrO21– ion. Consequently, it can function as both an oxidizing and reducing agent.
    MnO41–
  • reducing Mn is in the +7 oxidation state, which is the highest oxidation state it can achieve. Therefore, MnO41– cannot be oxidized or function as a reducing agent.
  • oxidizing Mn is in the +7 oxidation state, which is the highest oxidation state it can achieve. Therefore, MnO41– can be reduced (typically to Mn2+ or MnO2). The MnO41– ion is a strong oxidizing agent.
  • both Mn is in the +7 oxidation state, which is the highest oxidation state it can achieve. Therefore, MnO41– cannot be oxidized or function as a reducing agent.

11.1-6. Oxidizing and Reducing Agents in Reactions

To determine the oxidizing and reducing agents in a redox reaction, you must first identify the atoms whose oxidation states are changing. Oxidation state changes are usually assigned to a single atom of a molecule or polyatomic ion involved in a redox reaction, but the molecule or ion is the oxidizing or reducing agent. The following points should help you identify the oxidizing and reducing agents: Identify the oxidizing agent, the reducing agent, and the number of electrons transferred in the following redox reaction.
3 Cu + 2 NO31− + 8 H1+ 3 Cu2+ + 2 NO + 4 H2O
We identify elemental copper (Point 1) as a redox reactant. Point 2 can be used to ignore O and H. Copper is oxidized from 0 to +2, so Cu is the reducing agent.
reducing agent in a chemical reaction
We identify nitrate ion (Point 3) as a redox reactant. Point 2 can be used to ignore O and H. The oxidation state of nitrogen in the nitrate ion is determined using the rules developed in Section 4.4 as
charge on ion  = OXN + 3 OXO
−1 = OXN + 3(−2)
OXN = −1 + 6
= +5
Similarly, the oxidation state of nitrogen in NO is +2, so nitrogen is reduced from +5 to +2. However, the oxidizing agent is the molecule or ion, not the atom, so nitrate ion is the oxidizing agent. Actually, the oxidizing agent is nitric acid due to the presence of the H1+ ions. The simplest definition of an acid is a substance that produces H1+ ions in water. Acids are discussed in detail in Chapter 12.
oxidizing agent in a chemical reaction

11.1-7. Electrons Transferred

The number of electrons transferred in a reaction, i.e., the number gained by the oxidizing agent or lost by the reducing agent, is an important characteristic of a balanced redox equation. Consider the following redox equation:
Cr + 3 Ag1+ Cr3+ + 3 Ag
Each chromium atom loses three electrons as its oxidation state changes from 0 to +3. There is only one chromium atom in the equation, so the number of electrons lost by the chromium atoms in the above equation is
n = 1 Cr atom ×
3 electrons
1 Cr atom
= 3 electrons
Each silver ion gains one electron as its oxidation state changes from +1 to 0. However, there are three silver ions, so the number of electrons lost is
n = 3 Ag1+ ions ×
1 electron
1 Ag1+ ion
= 3 electrons
The balanced equation involves a three-electron transfer, i.e., n = 3.

11.1-8. Electrons Transferred Example

The following demonstrates the method for determining the number of electrons transferred in both the oxidation and reduction processes. Redox Reaction:
3 Cu + 2 NO31− + 8 H1+ 3 Cu2+ + 2 NO + 4 H2O
Oxidation:
electrons lost during reduction in a chemical reaction
Copper's oxidation state changes from 0 to +2, which is a two-electron loss by each copper atom. There are three copper atoms in the equation, so
n = 3 Cu atoms ×
2 electrons
1 Cu atom
= 6 electrons
This is a six-electron oxidation. Reduction:
The nitrogen atom in the nitrate ion is in a +5 oxidation state, but the N atom is +2 in NO. Thus, each nitrogen atoms gains three electrons. There are two nitrogen atoms on each side of the equation, so
n = 2 N atoms ×
3 electrons
1 N atom
= 6 electrons
This is a six-electron reduction. As required, the number of electrons gained in the reduction equals the number lost in the oxidation. We conclude that the reaction is a six-electron transfer reaction.

11.1-9. Practice

Exercise 11.2:

Indicate the oxidizing agent (Ox), the reducing agent (Red), and the number of electrons transferred (n) in the following redox reactions. Indicate any subscripted characters with an underscore (_) and any superscripted characters with a carat (^). For example, NH_4^1+ for NH41+.
2 Ag1+ + Cu 2 Ag + Cu2+
Ox = o_Ag^1+_s Ag1+ is reduced to Ag by accepting electrons from Cu, so Ag1+ oxidizes copper and is the oxidizing agent.
Red = o_Cu_s Cu is oxidized from 0 to +2 by donating electrons to Ag1+, so Cu reduces Ag1+ and is the reducing agent.
n = 2_0__ Each silver ion is reduced from +1 to 0, a one electron reduction. However, two ions are reduced in the reaction, so the two silver ions undergo a two electron reduction. Alternatively, copper is oxidized from 0 to +2, a two electron oxidation. Either way, 2 electrons are transferred.
2 Cr + 6 H1+ 2 Cr3+ + 3 H2
Ox = o_H^1+_s H is reduced from to +1 to 0 as H1+ accepts electrons from Cr. Thus, H1+ oxidizes Cr and serves as the oxidizing agent.
Red = o_Cr_s Cr is oxidized from 0 to +3 by donating electrons to H1+, so Cr reduces H1+ and is the reducing agent.
n = 6_0__ Each H1+ ion is reduced from +1 to 0, a one electron reduction. However, six ions are reduced in the reaction, so the six protons undergo a six electron reduction. Alternatively, each chromium is oxidized from 0 to +3, so two chromium atoms under go a (2 atom)(3 electrons/atom) = six electron oxidation. Either way, 6 electrons are transferred.
Fe2O3 + 3 CO 2 Fe + 3 CO2
Ox = o_Fe_2O_3_s The iron atoms are reduced by accepting electrons from CO, so Fe2O3 oxidizes CO and is the oxidizing agent.
Red = o_CO_s The carbon atoms are oxidized from +2 to +4 by donating electrons to Fe2O3, so CO reduces Fe2O3 and is the reducing agent.
n = 6_0__ Each iron atom is reduced from +3 to 0, which is three electrons per atom. However, two atoms are reduced in the reaction, so the iron atoms undergo a (2 atoms)(3 electrons/atom) = 6 electron reduction. Alternatively, the each carbon atom is oxidized from +2 to +4, which is two electrons per atom. Thus, the carbons atoms undergo a (3 atoms)(2 electrons/atom) = 6 electron oxidation. Either way, 6 electrons are transferred.
5 Au + 3 MnO41− + 24 H1+ 5 Au3+ + 3 Mn2+ + 12 H2O
Ox = o_MnO_4^1-_s The Mn atom is reduced from +7 in MnO41– to +2 in Mn2+, so MnO41– is the oxidizing agent.
Red = o_Au_s Au is oxidized from 0 to +3, so Au is the reducing agent.
n = 15_0__ Each gold atom undergoes a three-electron oxidation (0 3), and there are five gold atoms involved in the balanced equation. Thus, it is a 15 electron oxidation. Alternatively, each manganese atom undergoes a five-electron reduction (+7 +2), and there are three manganese atoms involved in the balanced equation. Thus, the equation represents a 15-electron reduction. Either way, 15 electrons are transferred.

11.1-10. Donor/Acceptor

  • 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.
Reducing agents are also called electron donors, while oxidizing agents are also called electron acceptors.
In a redox reaction, electrons transfer from a set of orbitals on the electron donor called the donor orbitals into a set of orbitals on the acceptor called the acceptor orbitals. The redox electrons are in the donor orbitals, so the donor must be in a reduced form of the substance, which is designated Red1. The acceptor orbitals can accept the redox electrons, so the acceptor must be in an oxidized form, which is designated Ox2. Electron transfer removes electrons from the donor orbitals of the reducing agent, so they are converted into acceptor orbitals, and the substance into an oxidizing agent. Similarly, the transfer places electrons into the acceptor orbitals of the oxidizing agent, so they are converted into donor orbitals and the substance into a reducing agent. The spontaneous process Red1 + Ox2 Ox1 + Red2 is considered below. Reactants Reactant 1 is in its reduced form (Red1), so it is a donor. Reactant 2 is in the oxidized form (Ox2), so it is an acceptor. As we shall see in the next section, the fact that the reaction is spontaneous usually implies that the donor orbitals are higher in energy than the acceptor orbitals.
electron transfer - reactants
Figure 11.2: Reactant Energies
Products Electron transfer empties the redox orbital of Reactant 1, so Product 1 is in its oxidized form (Ox1). Electron transfer fills the redox orbital on Reactant 2, so it is in its reduced form (Red2). Ox1 has an empty orbital, so it is now an acceptor (oxidizing agent), and Red2 has a filled orbital, so it is now a donor (reducing agent). Thus, electron transfer converts the oxidizing agent into a reducing agent and the reducing agent into an oxidizing agent.
electron transfer - reactants
Figure 11.3: Product Energies

11.1-11. Lower Energy

Spontaneous electron transfer occurs between oxidizing and reducing agents if the products are weaker reducing and oxidizing agents.
Recall from Section 9.7 that a reaction proceeds spontaneously at constant temperature and pressure so as to minimize its free energy. Furthermore, a reaction is extensive if its standard free energy change is negative. Thus, electrons transfer spontaneously from a donor to an acceptor if the transfer reduces their free energy, and they transfer extensively if the transfer reduces the standard free energy. We shall soon see that there are several factors that contribute to the standard free energy change, but the most important is often the orbital energy difference. Thus, the driving force behind many redox reactions arises because the acceptor orbitals are much lower in energy than the donor orbitals. In an extensive electron transfer, the donor orbitals are at higher energy than the acceptor orbitals in the reactants, but the acceptor orbitals are at higher energy in the products. Consequently, the produced oxidizing agent is weaker than the reacting oxidizing agent (the acceptor orbital is lower in the reactants). In addition, the produced reducing agent is weaker than the reactant reducing agent (the donor orbital is higher in the reactants). We conclude that spontaneous electron transfer occurs between stronger oxidizing and reducing agents to produce weaker reducing and oxidizing agents. Weak reactants do not react to produce strong ones.
electron transfer - extensive
Figure 11.4
Electron transfer between Fe and Cu2+ is extensive because the occupied orbitals of Fe are much higher in energy than the unfilled orbitals of Cu2+. Electron transfer from Cu back to Fe2+ is not extensive (Cu does not reduce Fe2+) because the electrons of Cu are much lower in energy than the unfilled orbitals of Fe2+. Electrons do not transfer extensively to higher energy.

11.1-12. Oxidizing and Reducing Agent Strengths

Strong reducing agents have high-energy electrons, and strong oxidizing agents have unfilled orbitals at low energy.
Strong reducing agents have high-energy electrons, and strong oxidizing agents have unfilled orbitals at low energy. When a strong oxidizing agent comes into contact with a strong reducing agent, the high-energy electrons in the donor orbitals flow spontaneously into the unfilled orbitals at much lower energy on the acceptor. As we shall soon see, the free energy that is released can be used to do work in batteries. If the donor orbitals are much lower than the acceptor orbitals, the reducing and oxidizing agents are too weak, and the electrons are not transferred.
Exercise 11.3:

Use the following energy diagram of the valence orbitals of A, B, and C to answer the question. Add two electrons to the unfilled orbital on A to form A2– and remove the electrons from the occupied orbitals on A, B, and C to form A2+, B2+ and C2+. Assume that orbital energy differences are the dominant term in the free energy to answer the following.
Free Energy
    best oxidizing agent
  • A2+ A2+ has the lowest energy unfilled orbitals, so it is the best oxidizing agent.
  • A A has unfilled valence orbitals to accept electrons, so it can function as an oxidizing agent, but it does not have the lowest energy unfilled orbitals.
  • A2– A2– has no unfilled valence orbitals to accept electrons, so it cannot be an oxidizing agent.
  • B B has no unfilled valence orbitals to accept electrons, so it cannot be an oxidizing agent.
  • B2+ B2+ does have unfilled valence orbitals to accept electrons, so it can function as an oxidizing agent.
  • C C has no unfilled valence orbitals to accept electrons, so it cannot be an oxidizing agent.
  • C2+ C2+ does have unfilled valence orbitals to accept electrons, so it can function as an oxidizing agent. However, the orbitals are high in energy, so it is a weak oxidizing agent.
    best reducing agent
  • A2+ A2+ has no valence electrons to donate, so it cannot be a reducing agent.
  • A A does have valence electrons to donate, but they are at very low energy, so A is a weak reducing agent.
  • A2– The valence electrons in A2– are relatively high in energy, so it is a good reducing agent. However, it is not the best in this group.
  • B B has valence electrons to donate, but they are at low energy, so B is not the strongest reducing agent.
  • B2+ B2+ has not valence electrons to donate, so it cannot be a reducing agent.
  • C C has valence electrons at the highest energy, so it is the best reducing agent.
  • C2+ C2+ has no valence electrons to donate, so it cannot be a reducing agent.
    reduces A
  • A2+ A2+ has no valence electrons to donate, so it cannot be a reducing agent.
  • A2– A2– can reduce A, but A2– is a product. Thus, there is no net change.
  • B B has valence electrons to donate, but they are lower in energy than the unfilled orbitals on A. Thus, the reducing power of B is not strong enough to reduce A.
  • B2+ B2+ has no valence electrons to donate, so it cannot be a reducing agent.
  • C C is the only substance with valence electrons that are at higher energy than the unfilled orbitals on A. Thus, the reducing power of C is great enough to reduce A to A2–.
  • C2+ C2+ has no valence electrons to donate, so it cannot be a reducing agent.
    oxidizes A2– but not A
  • A A can oxidize A2– but A2– is also a product, so no net change takes place.
  • A2– A2– has no unfilled valence orbitals to accept electrons, so it cannot be an oxidizing agent.
  • B B has no unfilled valence orbitals to accept electrons, so it cannot be an oxidizing agent.
  • B2+ The unfilled orbital on B2+ is below the donor orbital on A2–, so it will oxidize A2–. However, the unfilled orbital is above the donor orbital on A, so it will not oxidize A.
  • C C has no unfilled valence orbitals to accept electrons, so it cannot be an oxidizing agent.
  • C2+ The unfilled orbital on C2+ is too high to oxidize either species.
    reduces B2+ but not C2+
  • A2+ A2+ has no electrons to donate, so it cannot be a reducing agent.
  • A The electrons in A are too low in energy to reduce either B2+ or C2+.
  • A2– The higher energy electrons in A2– are higher in energy than the unfilled orbital on B2+ but lower than the unfilled orbital on C2+.
  • B Electron transfer from B to B2+ does not affect the concentration of either reactant.
  • B2+ B2+ has no electrons to donate, so it cannot be a reducing agent.
  • C2+ C2+ has no electrons to donate, so it cannot be a reducing agent.
    oxidizing and reducing agent
  • A2+ A2+ has no electrons to donate, so it cannot be a reducing agent.
  • A A can be oxidized to A2+ or reduced to A2–.
  • A2– A2– has no unfilled valence orbitals, so it cannot be an oxidizing agent.
  • B B has no unfilled valence orbitals, so it cannot be an oxidizing agent.
  • B2+ B2+ has no electrons to donate, so it cannot be a reducing agent.
  • C C has no unfilled valence orbitals, so it cannot be an oxidizing agent.
  • C2+ C2+ has no electrons to donate, so it cannot be a reducing agent.

11.1-13. Redox Couple

An oxidizing agent and the reducing agent it becomes are called a redox couple.
Electron transfer converts a reducing agent (Red1) into an oxidizing agent (Ox1) and converts an oxidizing agent (Ox2) into a reducing agent (Red2). The oxidizing agent and the reducing agent it becomes form a redox couple. Redox reactions involve two redox couples. A redox couple is written as the oxidized form, a slash, and then the reduced form (Ox/Red). For example, the two redox couples in the reaction of copper(II) and iron are expressed as follows:
Exercise 11.4:

Consider the following redox reaction:
Ag + VO21+ + 2 H1+ Ag1+ + VO2+ + H2O
Indicate any subscripted characters with an underscore (_) and any superscripted characters with a carat (^). For example, NH_4^1+ for NH41+.
What atom is oxidized? Give the symbol. o_Ag_ins Ag atoms are oxidized to Ag1+ ions. Ag atoms are oxidized to Ag1+ ions.
The oxidation couple (Ox/Red) is o_Ag^1+/Ag_s Writing the oxidized form of silver first, we obtain Ag1+/Ag for the oxidation couple.
What atom is reduced? Give the symbol. o_V_s V atoms are reduced from a +5 in VO21+ to +4 in VO2+.
The reduction couple (Ox/Red) is o_VO_2^1+/VO^2+_s Writing the oxidized form of vanadium first, we obtain VO21+/ VO2+ for the reduction couple.

11.2 Half-Reactions

Introduction

Redox reactions can be broken down into two half-reactions, an oxidation and a reduction, that show the loss and gain of electrons explicitly. The total reaction is the sum of the two half-reactions. Using half-reactions simplifies the writing of balanced redox reactions and helps us better quantify the driving force behind a redox reaction.

Objectives

11.2-1. Half-Reactions

A redox reaction can be broken down into two half-reactions: an oxidation half-reaction and a reduction half-reaction. The electrons gained are shown as reactants in the reduction half-reaction, and the electrons lost are shown as products in the oxidation half-reaction. The net redox reaction is simply the sum of the two half-reactions. The two half-reactions in the redox reaction between Fe and Cu2+ are:
oxidation half-reaction:FeFe2+ + 2 e1–
reduction half-reaction:Cu2+ + 2 e1– Cu
net reaction:Fe + Cu2+ Fe2+ + Cu
Note that the electrons gained in the reduction equals the number lost in the oxidation, so the electrons cancel in the sum to yield the net reaction. The net reaction never contains electrons because the number of electrons gained in the reduction must be the same as the number lost in the oxidation.
Exercise 11.5:

Consider the following reaction:
3 Sn + 2 Cr3+ 3 Sn2+ + 2 Cr
Represent the oxidation half-reaction as z Red z Ox + n e1–. Indicate any subscripted characters with an underscore (_) and any superscripted characters with a carat (^). For example, NH_4^1+ for NH41+. Include any coefficients other than 1. Omit any spaces.
z Red z Ox + n e1−
o_3Sn_s The reducing agent is 3Sn.
        
o_3Sn^2+_s The oxidized form is 3Sn2+.
    +    
o_6e^1-_s Three Sn atoms each lose two electrons, so
n = 3 × 2 = 6
electrons.
Represent the reduction half-reaction as y Ox + n e1– y Red. Indicate any subscripted characters with an underscore (_) and any superscripted characters with a carat (^). For example, NH_4^1+ for NH41+. Include any coefficients other than 1. Omit any spaces.
y Ox + n e1− y Red
o_2Cr^3+_s The oxidizing agent is 2Cr3+.
    +    
o_6e^1-_s Two Cr3+ ions each gain three electrons, so
n = 2 × 3 = 6
electrons.
         
o_2Cr_s The reduced form is 2Cr.

11.2-2. Summing Half-Reactions

Net redox reactions are usually constructed from tabulated half-reactions, which are always in the form of reductions in this course. The procedure is the following: Use the following two reduction half-reactions to write the chemical equation that explains what happens when aluminum metal is placed in a strong acid (H1+) to produce H2 gas.
Al3+ + 3 e1– equilibrium arrow Al
Reaction 1
2 H1+ + 2 e1– equilibrium arrow H2
Reaction 2

Reduction half-reaction

H1+ is a reactant and H2 is a product, so we use Reaction 2 as is.
2 H1+ + 2 e1− H2
reduction
Protons are reduced to hydrogen gas in the reduction half-reaction. Strong acids are the source of H1+ ions, thus strong acids can function as oxidizing agents. Indeed, this half-reaction represents the redox reaction of most strong acids. The nitrate ion in nitric acid is a stronger oxidizing agent than H1+, so this half-reaction is not used for nitric acid.

Oxidation half-reaction

Aluminum metal is oxidized to aluminum ion, so the reduction half-reaction given by Reaction 1 must be reversed to make it an oxidation.
Al Al3+ + 3 e1−
oxidation
Aluminum is a good reducing agent, and it reacts vigorously with acid. Most metals are oxidized to their ions by acid and some by water.

Electrons transferred

The reduction half-reaction involves a two-electron gain, while the oxidation half-reaction involves a three-electron loss. The LCM is therefore 6. The net equation is a six-electron transfer.

Electrons gained = electrons lost

The reduction half-reaction is multiplied by three, to make it a six-electron gain, and the oxidation half-reaction is multiplied by two to make it a six-electron loss.
reduction3 × 2 H1+ + 2 e1–H2
oxidation2 × AlAl3+ + 3 e1–

Sum half-reactions

Summing the two half-reactions yields the net equation for the reaction. The number of electrons gained equals the number lost, so the electrons cancel in the sum. Note that both the number of atoms and the total charge, which is +6 on each side, balance.
reduction6 H1+ + 6 e1–3 H2
oxidation2 Al2 Al3+ + 6 e1–
net equation6 H1+ + 2 Al3 H2 + 2 Al3+

11.2-3. Practice Writing Redox Equations

Exercise 11.6:

Use the following reduction half-reactions to write balanced redox reactions below:
Na1+ + e1− equilibrium arrow Na
Cr3+ + 3 e1− equilibrium arrow Cr
2 H2O + 2 e1− equilibrium arrow H2 + 2 OH1−
Ni2+ + 2 e1− equilibrium arrow Ni
Ag1+ + e1− equilibrium arrow Ag
Indicate any subscripted characters with an underscore (_) and any superscripted characters with a carat (^). For example, NH_4^1+ for NH41+. Include any coefficients other than 1. Omit any spaces.


Chromium metal is placed in a solution of Ag1+ ions.
Oxidation half reaction:

o_Cr_s Oxidation half reaction: Cr equilibrium arrow Cr3+ + 3e1–
    equilibrium arrow    
o_Cr^3+_s Oxidation half reaction: Cr equilibrium arrow Cr3+ + 3e1–
    +    
3_0__ Oxidation half reaction: Cr equilibrium arrow Cr3+ + 3e1– e1–
Red. 1                                 Ox. 1
Reduction half reaction:

o_Ag^1+_s Reduction half reaction: Ag1+ + e1– equilibrium arrow Ag
    +    
1_0__ Reduction half reaction: Ag1+ + e1– equilibrium arrow Ag e1–
    equilibrium arrow    
o_Ag_s Reduction half reaction: Ag1+ + e1– equilibrium arrow Ag
Ox. 2                                                                                Red. 2
LCM of transferred electrons:
3_0__ The lowest common multiple (LCM) of the electrons gained in the reduction and lost in the oxidation is 3.
Write the net equation (if any).
o_Cr_s Net Equation: Cr + 3 Ag1+ Cr3+ + 3 Ag
    +    
o_3Ag^1+_s Net Equation: Cr + 3 Ag1+ Cr3+ + 3 Ag
        
o_Cr^3+_s Net Equation: Cr + 3 Ag1+ Cr3+ + 3 Ag
    +    
o_3Ag_s Net Equation: Cr + 3 Ag1+ Cr3+ + 3 Ag
Red. 1                          Ox. 2                                  Ox. 1                              Red. 2
Metallic nickel is placed in a solution of Ag1+ ions.
Oxidation half reaction:

o_Ni_s Oxidation half reaction: Ni equilibrium arrow Ni2+ + 2e1–
    equilibrium arrow    
o_Ni^2+_s Oxidation half reaction: Ni equilibrium arrow Ni2+ + 2e1–
    +    
2_0__ Oxidation half reaction: Ni equilibrium arrow Ni2+ + 2e1– e1–
Red. 1                                 Ox. 1
Reduction half reaction:
o_Ag^1+_s Reduction half reaction: Ag1+ + e1– equilibrium arrow Ag
    +    
1_0__ Reduction half reaction: Ag1+ + e1– equilibrium arrow Ag e1–
    equilibrium arrow    
o_Ag_s Reduction half reaction: Ag1+ + e1– equilibrium arrow Ag
Ox. 2                                                                                  Red. 2
LCM of transferred electrons:
2_0__ The lowest common multiple (LCM) of the electrons gained in the reduction and lost in the oxidation is 2.
Write the net equation (if any).
o_Ni_s Net Equation: Ni + 2 Ag1+ Ni2+ + 2 Ag
    +    
o_2Ag^1+_s Net Equation: Ni + 2 Ag1+ Ni2+ + 2 Ag
        
o_Ni^2+_s Net Equation: Ni + 2 Ag1+ Ni2+ + 2 Ag
    +    
o_2Ag_s Net Equation: Ni + 2 Ag1+ Ni2+ + 2 Ag
Red. 1                          Ox. 2                                  Ox. 1                                  Red. 2
Chromium metal is placed in a solution of Ni2+ ions.
Oxidation half reaction:

o_Cr_s Oxidation half reaction: Cr equilibrium arrow Cr3+ + 3e1–
    equilibrium arrow    
o_Cr3+_s Oxidation half reaction: Cr equilibrium arrow Cr3+ + 3e1–
    +    
3_0__ Oxidation half reaction: Cr equilibrium arrow Cr3+ + 3e1– e1–
Red. 1                                 Ox. 1
Reduction half reaction:
o_Ni^2+_s Reduction half reaction: Ni2+ + 2e1– equilibrium arrow Ni
    +    
2_0__ Reduction half reaction: Ni2+ + 2e1– equilibrium arrow Ni e1–
    equilibrium arrow    
o_Ni_s Reduction half reaction: Ni2+ + 2e1– equilibrium arrow Ni
Ox. 2                                                                                  Red. 2
LCM of transferred electrons:
6_0__ The lowest common multiple (LCM) of the electrons gained in the reduction and lost in the oxidation is 6.
Write the net equation (if any).
o_2Cr_s Net Equation: 2Cr + 3Ni2+ 2Cr3+ + 3Ni
    +    
o_3Ni^2+_s Net Equation: Cr + 3 Ag1+ Cr3+ + 3 Ag
        
o_2Cr^3+_s Net Equation: 2Cr + 3Ni2+ 2Cr3+ + 3Ni
    +    
o_3Ni_s Net Equation: 2Cr + 3Ni2+ 2Cr3+ + 3Ni
Red. 1                          Ox. 2                                  Ox. 1                                  Red. 2
Metallic sodium is placed in H2O.
Oxidation half reaction:

o_Na_s Oxidation half reaction: Na Na1+ + e1–
    equilibrium arrow    
o_Na^1+_s Oxidation half reaction: Na Na1+ + e1–
    +    
1_0__ Oxidation half reaction: Na Na1+ + e1– e1–
Red. 1                                 Ox. 1
Reduction half reaction:
o_2H_2O_s Reduction half reaction: 2H2O + 2e1– equilibrium arrow H2 + 2OH1–
    +    
2_0__ Reduction half reaction: 2H2O + 2e1– equilibrium arrow H2 + 2OH1– e1–
    equilibrium arrow    
o_H_2_s Reduction half reaction: 2H2O + 2e1– equilibrium arrow H2 + 2OH1–
    +    
o_2OH^1-_s Reduction half reaction: 2H2O + 2e1– equilibrium arrow H2 + 2OH1–
Ox. 2                                                                      Red. 2
LCM of transferred electrons:
2_0__ The lowest common multiple (LCM) of the electrons gained in the reduction and lost in the oxidation is 2.
Write the net equation (if any).
o_2Na_s Net Equation: 2Na + 2H2O 2Na1+ + H2 + 2OH1–
  +  
o_2H_2O_s Net Equation: 2Na + 2H2O 2Na1+ + H2 + 2OH1–
    
o_2Na^1+_s Net Equation: 2Na + 2H2O 2Na1+ + H2 + 2OH1–
  +  
o_H_2_s Net Equation: 2Na + 2H2O 2Na1+ + H2 + 2OH1–
  +  
o_2OH^1-_s Net Equation: 2Na + 2H2O 2Na1+ + H2 + 2OH1–
Red. 1                      Ox. 2                          Ox. 1                         Red. 2

11.3 Galvanic Cells

Introduction

We have seen that electrons transfer spontaneously from the donor to the acceptor if the transfer lowers their free energy. Furthermore, the free energy that is released can be used to do work, as is done in a battery. In this section, we show how the relative free energies of the electrons in different redox couples are determined.

Prerequisites

Objectives

11.3-1. Definitions

Electrons flow spontaneously to the redox couple at higher electrical potential.
We now combine electron transfer reactions and electrical conduction through a circuit, which is the field of electrochemistry. We begin by recalling from Section 9.7 that –ΔG is the maximum amount of work (w) that can be obtained from any reaction at constant T and P. Extending that to redox reactions, we can write the following:
( 11.1 )
w = −ΔG
Maximum Work Done by Redox Electrons
If the redox reactants are in direct contact, the released free energy is lost as heat. However, the free energy can be harnessed by separating the reactants in an electrochemical cell and forcing the electrons through an external circuit. Electrochemical cells in which electrons flow spontaneously
(ΔG < 0)
are called galvanic cells.
Electrochemical cells were investigated about 60 years prior to the discovery of the electron, but it was known that the charge in a cell always moved from a lower or more negative electrical potential toward higher or more positive potential, so the early electrochemists expressed the work done in terms of the flow of charge and electrical potential rather than the flow of negatively charged electrons and energy change. The absolute value of the charge on a mole of electrons is called the faraday (ℱ). A faraday is simply Avogadro's number times the absolute value of the charge on a single electron.
1 ℱ = (6.022 × 1023 electrons/mol)(1.602 × 10−19 C/electron) = 96,500 C/mol
The work done by n moles of electrons (a charge of nℱ coulombs) being transferred through an electrical potential ℰ is given as
( 11.2 )
w = nℱℰ
Work and Charge
nℱ is the magnitude of the charge in coulombs that is transferred, and ℰ is the electrical potential difference through which the electrons move expressed in volts. A volt (V) is a joule per coulomb (1V = 1 J/C), so Equation 11.2 gives the work done by the electrons in joules. Rearranging shows that the electrical potential equals the work done by the electrons divided by the number of coulombs. In other words, ℰ is the work that can be done by each coulomb of charge. Combining Equation 11.2 with Equation 11.1 gives us the following relationship between the free energy of the redox reaction and the voltage that would be measured in the corresponding electrochemical cell:
( 11.3 )
ΔG = −nℱℰ = −96,500n
Free Energy in Redox Reactions
Equation 11.3 indicates that the cell potential (ℰ) becomes more positive as the free energy change of the redox reaction becomes more negative, which means that the amount of work that can be done by each electron increases as the cell potential becomes more positive.

11.3-2. Half-Cells

Just as a redox reaction is comprised of two half-reactions, an electrochemical cell is divided into two half-cells: These two cells are separated but connected through an electrical circuit, so electrons can flow from the anode to the cathode while keeping the two reactants separated. The two half-cells each develop an electrical potential, which is know as the half-cell potential and designated ℰanode and ℰcathode. Cell potentials are defined as the final electrical potential of the electron minus its initial electrical potential; i.e.,
cell = ℰfinal − ℰinitial
During a redox process, the electrons flow from the species that is to be oxidized in the anode to the species that is reduced in the cathode, so ℰinitial = ℰanode and ℰfinal = ℰcathode, and we may write
cell = ℰcathode − ℰanode
That is, the cell potential is the difference in the half-cell potentials. If all substances in the cell are in their standard state, then all potentials are standard potentials, and the standard cell potential can be expressed as the difference between the standard half-cell potentials as shown in Equation 11.4.
( 11.4 )
ℰ°cell = ℰ°cathode − ℰ°anode
Standard Cell Potential
The extent of reaction depends upon ΔG°, which can be determined with Equation 11.5.
( 11.5 )
ΔG° = −nℱℰ°
Standard Free Energy
ΔG° < 0 if ℰ°cell > 0, so a reaction is extensive if the standard cathode half-cell potential (ℰ°cathode) is greater than the standard anode half-cell potential (ℰ°anode). Thus, we could predict the extent of a redox reaction if we knew the relative values of the standard half-cell potentials of the redox couples.

11.3-3. Cell Description

Galvanic cells convert the chemical potential stored in high-energy electrons into electrical potential that can be used to do work.
A typical galvanic cell consists of four components: Each compartment contains the reactants in solution and a piece of metal immersed into a solution. The metal is called an electrode. Electrodes provide a surface at which electrons move between the circuit and the reactant. Active electrodes participate in the reaction (the atoms are reactants), while passive electrodes provide only the surface and do not participate in the reaction.
galvanic cell
Figure 11.5: An Electrochemical Cell

11.3-4. Anode

The anode is the oxidation half-cell. A typical anode half-reaction has the following form:
Red Ox + ne1−
The following points should be noted for the anode, which is boxed in red in Figure 11.6.
galvanic cell - anode compartment
Figure 11.6
The anode compartment, outlined by the red box, consists of a solution of Fe2+ ions and a piece of Fe metal.

11.3-5. Cathode

The cathode is the reduction half-cell. The cathode half-reaction can be expressed as
Ox + ne1− Red
The following points should be noted for the cathode, which is boxed in red in Figure 11.7.
galvanic cell - cathode compartment
Figure 11.7
The cathode compartment, outlined by the red box, consists of a solution of Cu2+ ions and a piece of metallic Cu.

11.3-6. Bridge

A liquid junction or salt bridge completes the circuit of an electrochemical cell by allowing ions to migrate between the compartments.
As reaction proceeds, Cu2+ ions are consumed as electrons enter the cathode, and Fe2+ ions are created as electrons leave the anode. However, all reaction would cease under these conditions because the two compartments would become electrically charged. The liquid junction maintains electrical neutrality by allowing ions to pass between the two compartments, while separating the oxidant and reductant. A salt bridge is a liquid junction that contains a saturated solution of a strong electrolyte, such as KCl. The flow of electrons into the cathode must be balanced by either a flow of anions out of the compartment or cations into the compartment. The charge balance in the cell shown in the figure can be maintained by sulfate ions leaving the cathode and entering the salt bridge or by potassium ions entering the cathode from the bridge. Note that two K1+ must enter for every one Cu2+ that is consumed. The flow of electrons out of the anode is balanced by a flow of anions into the compartment or cations out of the compartment. Thus, chloride ions migrate from the salt bridge into the anode or Fe2+ ions migrate from the anode into the salt bridge.
galvanic cell - liquid junction
Figure 11.8
The liquid junction in this electrochemical cell is a salt bridge, which consists of a saturated solution of KCl.

11.3-7. Load

The load is a device that uses the energy released by the transferred electrons.
The load might be the filament of a light bulb, a power tool, the starter of an automobile, a toy, or anything that requires a battery to operate. In the experiment discussed here, the load is a voltmeter. A voltmeter measures the electrical potential difference between the two electrodes. The two terminals of a voltmeter are labeled as follows: Note that a negative cell potential simply means that the terminals have been reversed; i.e., that the red terminal is actually connected to the cell at the more negative potential. In this case the electrons flow from the Hi (+) to the Lo (–) terminal. The voltage (cell potential) measured by the voltmeter is defined as the following:
cell = ℰHi − ℰLo
Recall that electrons move from lower to higher potential (from more negative toward more positive charge) and from the anode to the cathode. Consequently, the anode should be connected to the black (Lo or –) terminal and the cathode should be connected to the red (Hi or +) terminal to get a positive voltage. Thus, the cell potential can also be expressed as the following.
Ecell = EcathodeEanode
galvanic cell - load
Figure 11.9
The load in the figure is a voltmeter.

11.3-8. Example

Exercise 11.7:

The electrical potential of the Ag1+/Ag couple is about 1 V higher than that of the Ni2+/Ni couple.
Indicate the oxidizing and reducing agents in the electron transfer that takes place when Ag1+, Ag, Ni2+, and Ni are mixed.
    Oxidizing agent (oxidant)
  • Ag1+ The Ag1+/Ag couple is at the higher potential, so electrons flow spontaneously to it. Ag1+ is the oxidized species (OX) in the couple, so it is the oxidizing agent.
  • Ag The Ag1+/Ag couple is at the higher potential, so electrons flow spontaneously to it. However, Ag is the reduced form of the couple, i.e., the redox orbitals are full in Ag. Consequently, it cannot accept additional electrons, so it cannot be the oxidizing agent.
  • Ni2+ The Ni2+/Ni couple is at lower potential, so electrons flow spontaneously away from the couple. Consequently, the Ni2+/Ni couple is the oxidation couple. The oxidizing agent is the reactant in the reduction couple.
  • Ni The Ni2+/Ni couple is at lower potential, so electrons flow spontaneously away from the couple. Consequently, the Ni2+/Ni couple is the oxidation couple. The oxidizing agent is the reactant in the reduction couple.
    Reducing agent (reductant)
  • Ag1+ The Ag1+/Ag couple is at higher potential, so electrons flow spontaneously to the couple. Consequently, the Ag1+/Ag couple is the reduction couple. The reducing agent is the reactant in the oxidation couple.
  • Ag The Ag1+/Ag couple is at higher potential, so electrons flow spontaneously to the couple. Consequently, the Ag1+/Ag couple is the reduction couple. The reducing agent is the reactant in the oxidation couple.
  • Ni2+ The Ni2+/Ni couple is at lower potential, so electrons flow spontaneously away from it and it is the oxidation couple. However, Ni2+ is the oxidized form of the couple, so it does not contain the electrons that must be transferred.
  • Ni The Ni2+/Ni couple is at lower potential, so electrons flow spontaneously away from it and it is the oxidation couple. Ni is the reduced species (RED) in the oxidation couple, so it is the reducing agent.
Write the reaction. Indicate any subscripted characters with an underscore (_) and any superscripted characters with a carat (^). For example, NH_4^1+ for NH41+. Include any coefficients other than 1. Omit any spaces.
Oxidation half reaction:
o_Ni_s Oxidation half reaction: Ni Ni2+ + 2e1–
    equilibrium arrow    
o_Ni^2+_s Oxidation half reaction: Ni Ni2+ + 2e1–
    +    
2_0__ Oxidation half reaction: Ni Ni2+ + 2e1– e1–
Red. 1                          Ox. 1
Reduction half reaction:
o_Ag^1+_s Reduction half reaction: Ag1+ + 1e1– equilibrium arrow Ag
    +    
1_0__ Reduction half reaction: Ag1+ + 1e1– equilibrium arrow Ag e1–
    equilibrium arrow    
o_Ag_s Reduction half reaction: Ag1+ + 1e1– equilibrium arrow Ag
Ox. 2                                                                                 Red. 2
The number of electrons transferred (or the LCM) is:
2_0__ The lowest common multiple (LCM) of the electrons gained in the reduction and lost in the oxidation is 2.
Write the net equation (if any).
o_Ni_s Net Equation: Ni + 2Ag1+ Ni2+ + 2Ag
  +  
o_2Ag^1+_s Net Equation: Ni + 2Ag1+ Ni2+ + 2Ag
    
o_Ni^2+_s Net Equation: Ni + 2Ag1+ Ni2+ + 2Ag
  +  
o_2Ag_s Net Equation: Ni + 2Ag1+ Ni2+ + 2Ag
Red. 1                      Ox. 2                                  Ox. 1                          Red. 2

11.4 Standard Reduction Potentials

Introduction

The standard reduction potential of a redox couple is a measure of the electrical potential of the redox electron in that couple relative to the potentials of the redox electrons in other couples under standard conditions. Thus, standard reduction potentials can be used to determine cell potentials and to predict the spontaneity of redox processes.

Objectives

11.4-1. Standard Reduction Potentials Video

  • 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.

11.4-2. Reference Electrodes

Although we cannot measure a half-cell potential, we can measure the potential difference between two half-cells. Relative half-cell potentials can be obtained by defining one half-cell as a reference against which all other half-cells can be measured. Which half-reaction we choose for the reference half-cell and the value we assign its potential are arbitrary. The decision, made many decades ago, was to use the standard hydrogen electrode (SHE) as the reference and assign it a value of exactly 0 V. The SHE half-reaction is

Ox + ne1– equilibrium arrow Red ℰ° (V)
2 H1+ + 2 e1– equilibrium arrow H2 0.00
By convention, the half-cell to be measured is connected to the "Hi" terminal, so it is assumed to be the cathode or the reduction half-reaction. The SHE is then connected to the "Lo" terminal, so it is assumed to be the anode or oxidation half-reaction. If both cells are in their standard states, then the resulting cell potential is the standard reduction potential of the half-cell. It is a "reduction potential" because the half-cell to be measured is connected to the "Hi" terminal. Consider the setup in Figure 11.10.
standard hydrogen electrode experiment
Figure 11.10: Determining the Standard Reduction Potential of M
None of the reactants or products of the SHE reaction are solids, so an inert metal is used as the electrode. Platinum is used as the electrode in the figure. Electrons enter and leave the solution on the surface of the Pt.

11.4-3. Determining Standard Reduction Potentials

Standard Reduction Potential of Copper

Determining the standard reduction potential of Cu2+:

Ox + ne1– equilibrium arrow Red ℰ° in V
Cu2+ + 2 e1– equilibrium arrow Cu +0.34
When the Cu2+/Cu couple is connected to a half-reaction at a potential that is less than 0.34 V, the electrons move from the lower potential reactant and reduce Cu2+ to Cu, but when the couple is connected to a half-reaction at higher potential, the electrons leave the Cu and reduce the species at higher potential.
Figure 11.11: Determining the Standard Reduction Potential of Cu

Standard Reduction Potential of Iron

Determining the standard reduction potential of Fe2+:

Ox + ne1– equilibrium arrow Red ℰ° in V
Fe2+ + 2 e1– equilibrium arrow Fe –0.44
When the Fe2+/Fe couple is connected to a half-reaction at a potential more negative than –0.44 V, the electrons move from the lower potential reactant and reduce the Fe2+ to Fe. When the couple is connected to a half-reaction at a higher potential, the electrons leave the Fe to reduce the species at higher potential. The Fe2+/Fe couple (–0.44 V) is more negative than the H1+/H2 couple (0 V), so electrons transfer to the H1+/H2 couple and convert H1+ into H2.
Figure 11.12: Determining the Standard Reduction Potential of Fe

11.4-4. Calculating Cell Potentials

A standard cell potential is the potential difference between the cathode and the anode. The relative potentials of the two half-reactions are given by their standard reduction potentials, so standard reduction potentials can be used to determine the difference. Consequently, the standard cell potential of any electrochemical cell can be expressed as
ℰ°cell = ℰ°cathode − ℰ°anode
where Note that the standard reduction potential is a measure of the relative electrical potential of the half-cell. Consequently, standard reduction potentials are independent of the number of electrons transferred and the direction in which they are transferred. The Fe2+/Fe couple is at –0.44 V whether iron atoms are reduced or iron(II) ions are oxidized, i.e., whether it is the cathode or the anode. To determine the cell potential for the Fe + Cu2+ Fe2+ + Cu reaction, follow these steps.
1
Determine ℰ°cathode. Cu2+ is reduced, so ℰ°cathode is the standard reduction potential of the Cu2+/Cu couple: ℰ°cathode = +0.34.
2
Determine ℰ°anode. Fe is oxidized, so ℰ°anode is the standard reduction potential of Fe2+/Fe couple: ℰ°anode = –0.44 V. Note that the sign is NOT changed even though this is the half-reaction for the oxidation couple.
3
Determine the cell potential, ℰ°cell.
ℰ°cell = ℰ°cathode − ℰ°anode = ℰ°Cu − ℰ°Fe
ℰ°cell = +0.34 − (−0.44) = +0.78 V
The potential of the cathode is 0.78 V higher than that of the anode. A positive standard cell potential means that the reaction is extensive.

11.4-5. Reduction Potentials

Electron transfer is spontaneous when the standard reduction potential of the reducing agent is more negative than that of the oxidizing agent.
The standard reductions that we have considered thus far are shown in the table of standard reduction potentials below.

Ox + ne1– equilibrium arrow Red ℰ° (V)
Fe2+ + 2 e1– equilibrium arrow Fe –0.44
2 H1+ + 2 e1– equilibrium arrow H2 0
Cu2+ + 2 e1– equilibrium arrow Cu +0.34
Table 11.2
Note that the half-reactions are listed so that ℰ° is most negative at the top and most positive at the bottom. This choice places the electrons of greatest free energy at the top. The donor electrons are on the reducing agents (Red), while the acceptor orbitals are on the oxidizing agents (Ox). Spontaneous electron transfer occurs from a donor at higher free energy (lower or more negative potential) to an acceptor at lower free energy (higher or more positive potential). Thus, a redox reaction is spontaneous when the reducing agent lies above the oxidizing agent in a table of standard reduction potentials ordered in this manner.
Exercise 11.8:

Use Table 11.2 to answer the following questions.
    Which of the following is the best oxidizing agent?
  • H2 The best oxidizing species has acceptor orbitals (is on the right side of a reduction half-reaction) at the highest electrical potential (lowest free energy).
  • H1+ H1+ is an oxidizing agent, but its acceptor orbitals are not the lowest in free energy (most positive potential) of the three oxidizing agents listed here.
  • Fe The best oxidizing species has acceptor orbitals (is on the right side of a reduction half-reaction) at the highest electrical potential (lowest free energy).
  • Fe2+ Fe2+ is an oxidizing agent, but its acceptor orbitals are at the highest free energy (most negative potential) of the three oxidizing agents listed here.
  • Cu The best oxidizing species has acceptor orbitals (is on the right side of a reduction half-reaction) at the highest electrical potential (lowest free energy).
  • Cu2+ The acceptor orbitals on Cu2+ are at the lowest free energy (highest potential), so electrons transfer spontaneously from either H2 or Fe into the Cu2+ orbitals to produce copper atoms.
    Which of the following is the best reducing agent?
  • H2 The best reducing species is the one that has donor orbitals at the lowest (most negative) electrical potential (highest free energy).
  • H1+ A species on the right side of a reduction half-reaction contains the acceptor orbitals, so the substance is an oxidizing agent not a reducing agent.
  • Fe The iron orbitals are the donor orbitals at the most negative electrical potential (highest free energy), so iron is the best reducing agent.
  • Fe2+ A species on the right side of a reduction half-reaction contains the acceptor orbitals, so the substance is an oxidizing agent not a reducing agent.
  • Cu The best reducing species is the one that has occupied orbitals at the lowest (most negative) electrical potential (highest free energy).
  • Cu2+ A species on the right side of a reduction half-reaction contains the acceptor orbitals, so the substance is an oxidizing agent not a reducing agent.

11.5 Writing Redox Reactions

Introduction

We now use our understanding of half reactions and standard reduction potentials to write net equations for redox reactions.

Objectives

11.5-1. Reactivity and Standard Reduction Potentials Video

  • 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.

11.5-2. Writing REDOX Chemical Equations Video

  • 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.

11.5-3. Standard Reduction Potentials

Electrons transfer spontaneously from a reducing agent that is above and to the right of the oxidizing agent.
Electron transfer is spontaneous when the reactants are the stronger oxidizing and reducing agents and the products are the weaker reducing and oxidizing agents. In this discussion, we will express a general redox reaction between species 1 and 2 as:
Red1 + Ox2 Ox1 + Red2
The two half-reactions have the relative positions in a Table of Standard Reduction Potentials that are shown in Figure 11.13.
red and ox couples in a table of standard reduction potentials
Figure 11.13: Relative positions of the Ox1/Red1 and Ox2/Red2 couples in a Table of Standard Reduction Potentials.
Ox1 is above Ox2 in the table of standard reduction potentials, so we can conclude that
spontaneous reactions based on a table of standard reduction potentials
Figure 11.14: Electron transfer from upper right to lower left is extensive.
We also conclude that
determining spontaneity of a reaction from a table of standard reduction potentials
Figure 11.15: Electron transfer from lower right to upper left is NOT extensive.

11.5-4. Predicting Reactions Exercise

Exercise 11.9:

Note: sulfate ion is a common spectator ion in redox reactions, so ignore it in the following. Also, "metallic" implies that the metal is in the zero oxidation state, so it can be oxidized but not reduced. Use the Table of Standard Reduction Potentials to determine whether a reaction takes place when the following are mixed.
    Metallic tin is added to 1 M CuSO4.
  • Yes
  • No Sn does reduce Cu2+. Sn, the reducing agent, lies above Cu2+, the oxidizing agent, in the table so the reaction is extensive. Alternatively, electrons flow spontaneously from the reducing agent to the oxidizing agent because the oxidizing agent is at a more positive potential.
    Metallic silver is added to 1 M CuSO4.
  • Yes Ag does not reduce Cu2+. Ag, the reducing agent, lies below Cu2+, the oxidizing agent, in the table so no reaction takes place. Alternatively, electrons do not flow spontaneously from the reducing agent to the oxidizing agent because the oxidizing agent is at a more negative potential.
  • No
    Metallic tin is added to 1 M FeSO4.
  • Yes Sn does not reduce Fe2+. Sn, the reducing agent, lies below Fe2+, the oxidizing agent, in the table so no reaction takes place. Alternatively, electrons do not flow spontaneously from the reducing agent to the oxidizing agent because the oxidizing agent is at a more negative potential.
  • No
    Metallic iron is added to 1 M NiSO4.
  • Yes
  • No Fe does reduce Ni2+. Fe, the reducing agent, lies below Ni2+, the oxidizing agent, in the table so no reaction takes place. Alternatively, electrons flow spontaneously from the reducing agent to the oxidizing agent because the oxidizing agent is at a more positive potential.
    Metallic copper is added to water.
  • Yes H2O does not oxidize Cu to Cu2+ (otherwise we could not use copper pipes for water lines). Cu, the reducing agent, lies below H2O, the oxidizing agent, in the table so no reaction takes place. Alternatively, electrons do not flow spontaneously from the reducing agent to the oxidizing agent because the oxidizing agent is at a more negative potential.
  • No
    Metallic sodium is added to water.
  • Yes
  • No H2O oxidizes Na to Na1+. Na lies at the top of the table, so it reduces all of the OX species in the table.
    Metallic copper is added to 1 M HCl (hydrochloric acid).
  • Yes Copper does not dissolve in strong acids. All strong acids (except nitric acid) are represented by the H1+ at 0 V. Cu, the reducing agent, lies below H1+, the oxidizing agent, in the table so no reaction takes place. Alternatively, electrons do not flow spontaneously from the reducing agent to the oxidizing agent because the oxidizing agent is at a more negative potential.
  • No
    Metallic copper is added to 1 M HNO3 (nitric acid).
  • Yes
  • No Copper does dissolve in nitric acid. Nitric acid (H1+ + NO31–, o = +0.96 V) is a stronger oxidizing agent than the other strong acids due to the presence of the nitrate ion in acid. Cu, the reducing agent, lies above H1+ + NO31–, the oxidizing agent, in the table so reaction takes place. Alternatively, electrons flow spontaneously from the reducing agent to the oxidizing agent because the oxidizing agent is at a more positive potential.

11.5-5. Net Equations

Exercise 11.10:

Use the Table of Standard Reduction Potentials to write the net equation for the process described below and determine its standard cell potential. Indicate any subscripted characters with an underscore (_) and any superscripted characters with a carat (^). For example, NH_4^1+ for NH41+. Include any coefficients other than 1. Omit any spaces.
Metallic tin is placed in 1 M CuSO4.
Oxidation half reaction:
o_Sn_s Oxidation half reaction: Sn equilibrium arrow Sn2+ + 2e1–
    equilibrium arrow    
o_Sn^2+_s Oxidation half reaction: Sn equilibrium arrow Sn2+ + 2e1–
    +    
2_0__ Oxidation half reaction: Sn equilibrium arrow Sn2+ + 2e1– e1–
Red. 1                          Ox. 1
Reduction half reaction:
o_Cu^2+_s Reduction half reaction: Cu2+ + 2e1– equilibrium arrow Cu
    +    
2_0__ Reduction half reaction: Cu2+ + 2e1– equilibrium arrow Cu e1–
    equilibrium arrow    
o_Cu_s Reduction half reaction: Cu2+ + 2e1– equilibrium arrow Cu
Ox. 2                                                                              Red. 2
The number of electrons transferred (or the LCM) is:
2_0__ The lowest common multiple (LCM) of the electrons gained in the reduction and lost in the oxidation is 2.
Write the net equation (if any).
o_Sn_s Net Equation: Sn + Cu2+ Sn2+ + Cu
  +  
o_Cu^2+_s Net Equation: Sn + Cu2+ Sn2+ + Cu
    
o_Sn^2+_s Net Equation: Sn + Cu2+ Sn2+ + Cu
  +  
o_Cu_s Net Equation: Sn + Cu2+ Sn2+ + Cu
Red. 1                          Ox. 2                              Ox. 1                              Red. 2
The standard cell potential in volts is:
0.48___ The cathode is the Cu2+/Cu couple, so ℰcathode = +0.34 V.
The anode is the Sn2+/Sn couple, so ℰanode = –0.14 V.
The cell potential is ℰ = ℰcathode – ℰanode = 0.34 –(–0.14) = 0.48 V.
V

11.6 Common Batteries

Introduction

Batteries are devices that utilize the free energy from spontaneous redox reactions to generate electrical energy, i.e., they are electrochemical cells. Three of the more common batteries are described in this section.

Objectives

11.6-1. Alkaline Battery

All batteries are galvanic cells.
The alkaline battery is commonly used in flashlights. The half-reactions are: cell = 1.5 V
alkaline battery inside view
Figure 11.16: Alkaline Battery
The cathode reaction shown above is simply representative of a much more complicated process involving manganese species. The zinc base is an active anode, and a graphite rod serves as a passive cathode. The MnO2 is present in a moist paste of electrolyte (KOH, H2O, and KOH).

11.6-2. Silver (Button) Battery

The button battery is used in calculators, cameras, watches, etc. The half-reactions are: Note: Hg is sometimes used in place of Ag, in which case the cathode reaction is
HgO + H2O + 2 e1− Hg + 2 OH1−.
cell = 1.6 V when Ag2O is the oxidizing agent and 1.3 V when HgO is the oxidizing agent. The porous separator serves as the liquid junction.
Figure a: button battery outside view. Figure b: button battery inside view
Figure 11.17: Button Battery

11.6-3 Lead Acid Battery

Almost all automobiles are started with the power of the lead-acid battery. A 12 V battery actually contains six identical electrochemical cells, each producing 2.0 V. The two half-reactions of each cell are The electrodes, spongy lead anodes, and powdered PbO2 cathodes, are immersed in ~4.5 M H2SO4. A major advantage of the lead-acid battery is that it can be recharged by reversing the half-reactions with an external power supply. The alternator in an automobile, which is driven by the car's engine, generates electrical energy that is applied across the battery in the reverse direction. The electrical energy is used to drive the reactions in the reverse direction. Consequently, the starting materials are regenerated for the next time the car is started. We take up the recharging process in more detail in our discussion of electrolytic cells.
lead storage battery
Figure 11.18: Lead Storage Battery
The potential of the cathode of each cell is 2 V higher than that of its anode, but metallic conductors keeps the anode of each cell at the same potential as that of the previous cell. Consequently, the potential of the cathode increases by 2 V in each cell. The figure shows a battery with three cells, so it is a 6 V battery. A 12 V battery would have six such cells.

11.7 Corrosion

Introduction

Corrosion is the unwanted oxidation of a metal. Approximately 25% of new steel production is devoted to the replacement of corroded steel. This very costly process is a series of electrochemical reactions that take place in the iron and the water that is on it.

Objectives

11.7-1. Definition and Example

Corrosion is the unwanted oxidation of a metal. Consider the following observations regarding the corrosion of iron. These observations are explained in the figure, which shows the rusting of iron as an electrochemical cell in which the iron is both an active anode and a passive cathode. Note that oxygen and other gases from the atmosphere dissolve in rain drops, and some of these gases react with water to produce H1+, which makes the drops acidic. For example, SO3, which is an industrial pollutant, reacts with water to produce sulfuric acid (H2SO4), and CO2 reacts in a similar manner to produce carbonic acid (H2CO3).
corrosion of iron
Figure 11.19: Corrosion of Iron
A drop of water with dissolved oxygen and some acid sitting on a piece of iron is an electrochemical cell.

11.7-2. Protection Against Corrosion

Metals can be protected from corrosion by being galvanized or passivated.


Process Definition Examples
Galvanization Protecting one metal by sacrificing another more reactive metal, which reacts instead of the protected metal.
  • The standard reduction potentials of iron and zinc are –0.44 and –0.76 V, respectively, so zinc is oxidized more easily than iron. Consequently, automobile bodies are galvanized with a zinc coating, which prevents the corrosion of the iron frame.
  • Magnesium (ℰ° = –2.36 V) and aluminum (ℰ° = –1.66 V) rods are connected to underground iron (ℰ° = –0.44 V) pipes to protect them. The reactive metals are sacrificed before the iron corrodes.
Passivation Protection of a reactive metal by the formation of an oxide layer. Some of the metal at the surface reacts, but the product of the oxidation is an oxide layer that is impervious to further attack.
  • Aluminum is a reactive metal, but an aluminum chair can be left outdoors because the reactive metal forms a coating of aluminum oxide on the surface. The layer protects the underlying metal.
  • The iron in an automobile body is also passivated by ZnO, which forms from the sacrificial Zn that was added in the galvanization process. The ZnO coating further protects the underlying iron.

11.8 Electrolytic Cells

Introduction

We have seen that spontaneous redox reactions can be used to generate electrical energy in an electrochemical cell, and, in this section, we demonstrate the reverse of that process, using electrical energy to drive redox reactions that are not spontaneous.

Objectives

11.8-1. Electrolytic Cells

Electrolytic cells convert electrical potential energy into chemical potential energy.
Thus far, we have concerned ourselves with galvanic cells, i.e., cells in which the electron transfer is spontaneous. However, a major advantage of electrochemistry is that we can vary the free energy of the electrons in the electrodes and force electrons to transfer uphill in free energy. The cell potential is a measure of the amount of work we can get out of a galvanic cell (ℰcell > 0), but if it's negative it indicates the amount of work that an external power source must provide to an electrolytic cell (ℰcell < 0) in order to force the reaction. That is, galvanic cells convert chemical potential energy into electrical potential energy, while electrolytic cells convert electrical potential energy into chemical potential energy. Forcing a nonspontaneous reaction to take place by the application of a voltage from an external source is called electrolysis. Electrolysis is used to extract many metals from their ores, and to coat one metal (frequently iron) with another in a process called plating (silver and gold plated jewelry, and chrome or nickel plated sink fixtures).
direction of energy flow in galvanic and electrolytic cell
Figure 11.20
For example, the redox electrons are at much higher potential energy on Na than on Cl1–. Consequently, the reaction of Na with Cl2 in an electrochemical cell would produce a large voltage as the chemical potential energy of the electrons is converted to electrical potential energy. When Na1+ and Cl1– are brought into contact no reaction takes place unless electrical potential energy from a power supply is added to drive the electrons from the lower potential energy of the chloride ion to the higher potential energy of the sodium atom. The nonspontaneous reaction is called electrolysis. In this example, NaCl is electrolyzed to Na and Cl2.

11.8-2. Lead Storage Batteries

A car battery operates like a galvanic cell when it starts a car and like an electrolytic cell when it is recharging.
Recharging batteries is also a function of electrolytic cells. The lead-acid battery is one of the most common examples. Cranking power is given in amps, which is a measure of the rate at which electrons flow through the circuit, and every two moles of electrons that flow through the circuit produce two moles of PbSO4. A battery with 550 amp of cranking power produces 1.7 g PbSO4 per second. At this rate, the battery would run out of reactants after only a few starts. However, cars are equipped with generators (or alternators) that force the spontaneous discharge reaction to reverse while the car is running, i.e., they recharge the battery. Thus, lead-acid batteries are galvanic while a car is starting and electrolytic while it is running.

Starting Galvanic
  • Anode: Pb + SO42– PbSO4 + 2 e1–
  • Cathode: PbO2 + 4 H1+ + SO42– + 2 e1– PbSO4 + 2 H2O

Electron transfer is spontaneous and develops 2 V per cell. Electrons move through the starter to get from the anode to the cathode. The free energy change of the electrons is used to start the car.

Driving Electrolytic
  • Anode: PbSO4 + 2 H2O PbO2 + 4 H1+ + SO42– + 2 e1–
  • Cathode: PbSO4 + 2 e1– Pb + SO42–

Electrons must be forced from the anode to the cathode because the cathode is now at lower potential (higher free energy). In this case, electrons flow through the alternator to get from the anode to the cathode, and the battery recharges.

11.9 Exercises and Solutions

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