Anda di halaman 1dari 18

Chemistry 3830 Lecture Notes

5.

Page 76

Dr. M. Gerken

Redox chemistry and multiple oxidation states

So far we have only mentioned the fact that transition elements have multiple oxidation states, but we have ignored the
impact of this on chemical behavior. We will do so now, and then we will continue to use redox principles to treat the
chemistry of the p-block elements. Redox chemistry is probably the most interesting (and complex) aspect of the chemical
behavior of the elements, and careful study of this material will greatly reward the effort expended. The majority of the
reactions which you do in Part II of the lab are in some sense redox reactions.
As background to this section, you should review how to balance redox reactions and other basic electrochemical
principles from General Chemistry.

5.1

Review of basic principles from Chemistry 2000

Oxidation-Reduction reactions, better known by the short-form redox reactions, involve the transfer of electrons between
atoms and molecules during the reaction. They represent one of the major classes of chemical reactions. We can detect redox
reactions by monitoring the oxidation states of the atoms in the reactants and products. Any change in these formal oxidation
states during the reactions means that a redox reaction has taken place. Oxidation states were introduced in Chem2000, and
you should go through as many exercises on assigning oxidation states as possible. You may wish to review the section in your
Chem2000 textbook.

5.1.1. Definition of Oxidation States


For reaction involving the transformation of a neutral atom into an ion, e.g., the oxidation of Na to Na+, the identification
of the electron transfer (and the assignment of oxidation states) is obvious. The assignment of oxidation states in covalent
compounds is far more complex. In order to see which atom in a compound looses/gains electrons, we formally (on paper)
decompose a covalent compound into monatomic ions. This is done by formal heterolytic cleavage of bonds according to the
electronegativity difference. For example: the H-Cl bond is formally broken by assigning the bonding electron pair to the more
electronegative bonding partner (Cl). This gives the fragments: H+ and Cl-. The charges of these formal fragments are the
oxidation states of the corresponding atoms in the covalent compound. In the case of element-element bonds, homonuclear
bonds, the bonds are cleaved homolytically (one electron of a single bond (two electrons of a double bond) goes to each of the
bonding partners. Oxidation states are usually denoted by Roman numerals.
Example: H2O2
1.step: Lewis structure
2.step: Formal bond cleavage

.. ..
H O
.. O
.. H

.. ..
H( O
.. )( O
.. )H

3. step: Charges of fragments

+
H

...
:O
..

.. .O :
..

+
H

..
.
:O
..

..
.O :
..

4. step: Indicate oxidation states in Lewis structure

+I -I.. -I
.. +I
H O
.. H
.. O

5.1.2 Guidelines for Determining Oxidation Numbers


Usually, you do not have to go through the formal bond cleavage, as shown above. Simple rules has been developed to
calculate oxidation states for most of the known compounds whithout drawing Lewis structures. However, by using these rules,
you assume a certain Lewis structure.
1.
Each atom in a pure element has an oxidation number of 0. The oxidation number of Cu in metallic copper is 0
and is zero for each atom in I2 or S8. [always valid]
2.
For ions consisting of a single atom, the oxidation number is equal to the charge on the ion. Elements of periodic
Groups 1, 2 and 13 form monatomic ions with a positive charge and oxidation number equal to the group number.
Aluminum therefore forms Al3+ , and its oxidation number is +III. (See Section 3.3.) [always valid]
3.
Fluorine is always I in compounds with other elements. [always valid]
4.
Cl, Br, and I are always I in compounds except when combined with more electronegative elements, such as O
or F. This means that Cl has an oxidation number of I in NaCl (in which Na is +I, as predicted by the fact that it
is an element of Group 1). In the ion ClO the Cl atom has an oxidation number of +I (and O has an oxidation
number of II; see Guideline 5).
5.
The oxidation number of H is +I and of O is II in most compounds. Although this statement applies to many,
many compounds, a few important exceptions occur.

Chemistry 3830 Lecture Notes

Dr. M. Gerken

Page 77

6.

When H forms a binary compound with a metal, the metal forms a positive ion and H becomes a hydride ion,
H. Thus, in CaH2 the oxidation number of Ca is +II (equal to the group number) and that of H is I. [The
deciding factor remains the electronegativity of the bonding partner]
Oxygen can have an oxidation number of I in a class of compounds called peroxides, compounds based on
the O22 ion. For example, in H2O2, hydrogen peroxide, H is assigned its usual oxidation number of +I, and
so O is I. [If O is bonded to fluorine you have to use the definition given in 7.1.1]
The algebraic sum of the oxidation numbers in a neutral compound must be zero; in a polyatomic ion, the sum
must be equal to the ion charge. Examples of this rule are the previous compounds and others found in Example
4.8.

5.1.3 Oxidation and Reduction - a working definition


The following basic definition is used to describe redox processes:
If X loses one or more electrons, it is oxidized, and is the reducing agent.
X Xn+ + n e
If Y gains one or more electrons, it is reduced, and is the oxidizing agent.
Y + n e Yn
These are definitions that must be learned cold. They are the oui and non of electrochemistry. All the other ideas are
based on them. The origin of the term oxidation comes from the fact that combination with oxygen (e.g. in combustion) is one
common form of an oxidation reaction. The element oxidized loses electrons to the oxygen atom. For example, when
magnesium is burnt:
Mg Mg2+ + 2 e
while oxygen gains electrons:
O2 + 2 e O2
and the net reaction is:
Mg + O2 MgO
Reduction originates in the concept of reducing an ore, usually a metal oxide, to the elemental form, usually the pure metal.
For example when iron is made from iron ore, the iron reaction is:
Fe2+ + 2 Fe3+ + 8 e 3 Fe
and the carbon monoxide reaction is:
4 CO 4 CO2+ + 8 e
Here the oxide ions are spectators, transferring from the iron to the oxidized carbon. The net reaction therefore becomes:
Fe3O4 + 4 CO 3 Fe + 4 CO2

5.1.4 Balancing redox equations in solution


These examples illustrate the common technique of separating redox reactions into two complementary redox halfreactions. In the above examples, the Mg/Mg2+ reaction is an oxidation half reaction. In such reactions, electrons are
always among the products. The CO/CO2+ reaction is also an oxidation half reaction. The other two are reduction halfreactions, for which electrons will always be reactants. Although simple reaction such as these are easy to balance at sight,
and we often dont stop to think that they are in fact redox reactions, more complicated redox reactions need a rigorous
approach for successful balancing. We will concentrate on reactions that take place in solution, and these will usually be either
basic or acidic solutions. Since electrochemistry is a branch of thermodynamics, we usually will be dealing with reactions
under conditions of standard state, which for solution chemistry means 1 M concentrations at 25C. Thus acidic solutions
will have [H3O+] = 1.00 M, while basic solutions will have [OH] = 1.00 M. The usual Kw relationship between hydroxide and
hydronium ion will always hold.
The rules for balancing redox reactions are as follows. These are better learned by doing many examples than by actual
memorization of the rules!
Step 1. Assign oxidation states and recognize the reaction as an oxidation-reduction.
Step 2. Separate the overall process into half-reactions.
Step 3. Balance each half-reaction for material (first balance oxidized/reduced element, then balance O, then balance H).
In acid solution, add H2O to the side requiring O atoms.
Then add H+ to balance any remaining unbalanced H atoms.
In basic solution, add H2O to the side requiring O atoms.
Then add H2O to the side requiring H atoms, and one OH to the other side. Add H2O/OH- for every
unbalanced H atom.
Step 4. Balance the half-reactions for charge by inserting the correct number of electrons. Consider the number of
oxidized/reduced atoms.
Step 5. Multiply the balanced half-reactions by appropriate factors to obtain the same number of electrons in both halfreactions.
Step 6. Add the balanced half-reactions.

Chemistry 3830 Lecture Notes

Page 78

Dr. M. Gerken

Step 7. Simplify the equation by eliminating common reactants and products.


Step 8. Check the final result for material and charge balance.

5.1.5 Some examples of balancing redox reaction equations


The following reaction takes place in standard acid solution (1 M H+)
MnO4(aq) + HSO3(aq) Mn2+(aq) + SO42(aq)
We recognize that Mn changes oxidation state from +7 to +2, while S changes from +4 to +6. The two half reactions are:
HSO3 SO42
Oxidation:
Material balance
HSO3 + H2O SO42 + 3 H+
Electron balance
HSO3 + H2O SO42 + 3 H+ + 2 e
Multiply by 5
5 HSO3 + 5 H2O 5 SO42 + 15 H+ + 10 e
Reduction:
Material balance
Electron balance
Multiply by 2
Overall reaction:

MnO4
Mn2+

MnO4
+ 8 H+ Mn2+ + 4 H2O

MnO4
+ 8 H+ + 5 e Mn2+ + 4 H2O

2 MnO4
+ 16 H+ + 10 e 2 Mn2+ + 8 H2O

2 MnO4(aq) + 5 HSO3(aq) + H+(aq) 2 Mn2+(aq) + 5 SO42(aq) +

3 H2O

Finally, check that the atoms and charges balance at the right and at the left
The following reaction takes place in standard basic solution (1 M OH-)
Fe(OH)2(s)
+ CrO42(aq) Fe2O3(s)
+ Cr(OH)4(aq)
We recognize that Fe changes oxidation state from +2 to +3, while Cr changes from +6 to +3. The two half reactions are:
Fe(OH)2
Fe2O3
Oxidation:
Material balance
2 Fe(OH)2 + 2 OH Fe2O3 + 3 H2O
Electron balance
2 Fe(OH)2 + 2 OH Fe2O3 + 3 H2O + 2 e
Multiply by 3
6 Fe(OH)2 + 6 OH 3 Fe2O3 + 9 H2O + 6 e
Reduction:
Material balance
Electron balance
Multiply by 2
Overall reaction:

CrO42
Cr(OH)4
2
CrO4
+ 4 H2O Cr(OH)4 + 4 OH
2
CrO4
+ 4 H2O + 3 e Cr(OH)4 + 4 OH
2
2 CrO4
+ 8 H2O + 6 e 2 Cr(OH)4 + 8 OH

6 Fe(OH)2(s) + 2 CrO42(aq) 3 Fe2O3(s) + 2 Cr(OH)4(aq) + 2 OH(aq) +

H 2O

Finally, check that the atoms and charges balance at the right and at the left

5.1.6 Spontaneous reactions


Gibbs Free Energy of reaction tells us whether a certain reaction is product favoured or reactant-favoured in the forward
direction. The electrochemical analogue to this concept is called the cell potential given by the symbol E. E is expressed in
the common electrical unit of volts, but is really just the Gibbs free energy in disguise! The exact relationship between cell
potential and Gibbs energy is given by the relationship:

G = nFE in general, and for standard conditions; G 0 = nFE 0 ;


n = number of electrons transferred, F = Faradays constant = 96485 C mol-1
Because of the minus sign in the equation, the cell voltage E has the opposite sign convention to that of G:
Product-Favoured Reaction:
Reactant-Favoured Reaction:

ve G
+ve G

or
or

+ve E
ve E

Why do we have such a confusing convention? The origin of this discrepancy comes from the way that physicists define
current flow: although electrical current is entirely due to the migrations of electrons in a circuit, a forward current flow is
defined in standard theories of electricity as the direction of positive charge flow. It is when harmonizing standard electrical
conventions with chemical definitions that these apparently contradictory conventions are generated. In electrochemistry, a
reaction which is product favoured produces a voltage, and is therefore called a Voltaic cell, in honor of Alessandro Volta. A

Chemistry 3830 Lecture Notes

Dr. M. Gerken

Page 79

reaction which is reactant favoured can be driven forward by the application of a greater opposite voltage; such reactions are
called electrolytic cells, and the process is named electrolysis.
We now want to remember another aspect of the meaning of G, which is that it tells for any given reaction the maximum
amount of the total energy change that may be harnessed for useful work. The electrochemical equivalent of this idea is the
maximum voltage of the electrochemical cell. This maximum potential is obtainable only from a fully charged cell running
under zero load. This means that we can calculate the maximum cell voltage from standard free energies for the reaction.
Consider the reaction:

Zn2+(aq)
+
Cu(s)
Zn(s) +
Cu2+(aq)
1
fG (kJ mol )
0
65.52
147.0
0
Since rG = {fG (products)} {fG(reactants)}, rG = (-147.0 kJ/mol) (65.52 kJ/mol) = 212.5 kJ mol1, and we
would predict this to be a spontaneous reaction. It is spontaneous, but frankly quite useless unless you need some finely
divided metallic copper and have only these reactants at hand. Let us now see if it is indeed possible to harness the useful
work from this reaction.

5.1.7 The electrochemical cell


To do this we must perform exactly the same
chemical reaction but using an electrochemical cell.
A typical cell-design is shown in the figure below. It
divides the redox reaction into to compartments, each
of which contains one of the two redox halfreactions. What voltage does our device measure?
What voltage should it measure? Well, rG =
212.5 kJ mol1, and the value of n = 2, as can be seen
by breaking the reaction into its constituent
electrochemical half-reactions:
Zn(s) Zn2+(aq) + 2 e Oxidation
Cu2+(aq) +
and so:

2 e

Cu(s)

Reduction

G 0
212.5 kJ mol 1
E =
=
= +1.10V
nF
2 96.485 kC mol 1
0

Any discrepancy between the actual measured


potential of a given cell and this value is usually due
to non-standard conditions, primarily that the
concentrations are not exactly one Molar.
Diagram of an electrochemical cell to harness the Zn/Cu reaction

5.1.8 Electrode or Half-cell potentials


Although the reaction only requires a
copper(II) salt and metallic zinc, we have
built a cell in which both metallic zinc
and copper are present, and copper and
zinc sulfate. The reason for this is to
make the reactions reversible. It turns
out that accurate potentials can only be
measured for reversible electrochemical
cells. Electrochemists have developed a
short-hand notation to indicate the
complete
make-up
of
a
given
electrochemical cell. It is based on the
balanced redox reaction, but includes
both the reactants and the products. For
our cell the correct notation would be:
Zn(s) | Zn2+(1 M) || Cu2+(1 M) | Cu(s)
In this notation, the cathode reaction is always put on the right hand side, and the anode reaction at the left. The cathodic
process is always reduction, while the anodic process is always oxidation. Thus the anode half-cell will always be written at

Chemistry 3830 Lecture Notes

Dr. M. Gerken

Page 80

the left hand side of the notation. A convenient mnemonic device to remember this convention is Right Red Cat, i.e. that the
cathode process is a reduction process and is placed at the right of the cell notation. The vertical lines, |, in the notation
indicate a phase boundary, such as between a solid and a liquid. Double lines, ||, indicate double boundaries, such as
commonly occur when a salt bridge is placed between the two half cells. If the components of a redox reaction co-exist in
solution without a phase interface, they are listed together with a comma separating oxidized and reduced forms. This is often
the case when the current is introduced into a solution via an inert electrode (usually platinum metal); an example is the
oxidation of iron(II) to iron (III) for which the notation could be: Pt(s) | Fe2+(1 M), Fe3+(1 M).
The operation of a voltaic cell is summarized in Figure 21.4. It represents a complete electrical circuit. In the wires
external to the cell, the current is carried exclusively by a moving stream of electrons. However, within the solution, the
current must be carried by migrating ionic species, such as in our example cell Zn2+ and Cu2+ ions. Within the salt bridge, if
one is used, inert ions such as K+ and Cl carry the charges and provide for current flow.
Standard Half-Cell Reduction Potentials
From the construction of our electrochemical demonstration cell, and from the cell notation, it should now be obvious to
you that electrochemical half-reactions, which we arbitrarily introduced as a tool to balance redox reactions, have some
degree of physical reality they can after all be built into electrochemical half-cells. It should, for example, be possible to
uncouple the copper/zinc cell and construct other voltaic cells from them, for example a copper/lithium or a silver/zinc cell, etc.
This is indeed possible. It would also be nice to be able to assign a voltage to each half-cell, but this turns out to be
impossible, since the voltage depends on a complete electrical circuit being established. Nonetheless, this idea of a half-cell
potential is so attractive, that chemists have figured out a way to do so. We do it by arbitrarily setting one redox half-reaction
to zero, and measuring all other cells with respect to this arbitrary zero reference point. The universal reference standard for
electrochemistry is the standard hydrogen electrode, or SHE. For example, if the zinc half-cell is combined with the SHE, as
shown in the Figure below, a voltage is measured as +0.76 V. We assign this cell voltage entirely to the zinc half cell, since
SHE is zero. Alternatively, we combine the SHE with the copper half cell, and get +0.34 V. Then we can establish the
potential of the copper zinc cell as E = + 0.76 + 0.34 V = +1.10 V.

Since for any given combination of two half-cells, we can never be sure which half will act as the anode and which as the
cathode, electrochemists have a standardized way of expressing half-cell potentials, and that is always to depict them as
reductions. This leads to the compilation of the standard reduction potentials in aqueous solution which are compiled in
the Table below.
We can use this table in numerous ways. First of all, we can now construct an electrochemical cell out of any
combination of half cell reactions. Thus we can go and obtain potentials for the examples cited above:

Chemistry 3830 Lecture Notes

Dr. M. Gerken

Page 81

Table of Standard Reduction Potentials in Aqueous Solution at 25 C


Reduction Half-Reaction
F2(g) + 2 e
2 F (aq)
+

H2O2(aq) + 2 H3O (aq) + 2 e


4 H2O(l)
PbO2(s) + SO4 2 (aq) + 4 H3O + (aq) + 2 e
PbSO4(s) + 6 H2O(l)
MnO4 (aq) + 8 H3O + (aq) + 5 e
Mn 2+ (aq) + 12 H2O(l)
3+

Au (aq) + 3 e
Au(s)
Cl2(g) + 2 e
2 Cl (aq)
Cr2O7 2 (aq) + 14 H3O + (aq) + 6 e
2 Cr 3+ (aq) + 21 H2O(l)
+

O2(g) + 4 H3O (aq) + 4 e


6 H2O(l)
Br2(l) + 2 e
2 Br (aq)
NO3 (aq) + 4 H3O + (aq) + 3 e
NO(g) + 6 H2O(l)
OCl (aq) + H2O(l) + 2 e
Cl (aq) + 2 OH (aq)
Hg 2+ (aq) + 2 e
Hg(l)
Ag + (aq) + e
Ag(s)
Hg2 2+ (aq) + 2 e
2 Hg(l)
Fe 3+ (aq) + e
Fe 2+ (aq)

I2(s) + 2 e
2 I (aq)

O2(g) + 2 H2O(l) + 4 e
4 OH (aq)
2+

Cu (aq) + 2 e
Cu(s)
Sn 4+ (aq) + 2 e
Sn 2+ (aq)

2 H3O + (aq) + 2 e
H2(g) + 2 H2O(l)
Sn 2+ (aq) + 2 e
Sn(s)
Ni 2+ (aq) + 2 e
Ni(s)
V 3+ (aq) + e
V 2+ (aq)
PbSO4(s) + 2 e
Pb(s) + SO4 2 (aq)
2+

Cd (aq) + 2 e
Cd(s)
Fe 2+ (aq) + 2 e
Fe(s)
Zn 2+ (aq) + 2 e
Zn(s)
2 H2O(l) + 2 e
H2(g) + 2 OH (aq)
3+

Al (aq) + 3 e
Al(s)
Mg 2+ (aq) + 2 e
Mg(s)
Na + (aq) + e
Na(s)
K + (aq) + e
K(s)
Li + (aq) + e
Li(s)

In volts (V) versus the standard hydrogen electrode.

E (V)
+2.87
+1.77
+1.685
+1.52
+1.50
+1.360
+1.33
+1.229
+1.08
+0.96
+0.89
+0.855
+0.80
+0.789
+0.771
+0.535
+0.40
+0.337
+0.15
0.00
0.14
0.25
0.255
0.356
0.40
0.440
0.763
0.8277
1.66
2.37
2.714
2.925
3.045

5.1.9 Constructing cell potentials from standard half-cell potentials


copper/lithium cell
From the table we get the half-reactions:
Cu 2+ (aq) + 2 e Cu(s)
E = +0.337 V
+

Li (aq) + e Li(s)
E = 3.045 V
We now combine them in such a way as to get a positive overall cell potential. This means we must reverse the lithium
equation, making it the anode (where the oxidation will take place.)
E = +3.045 V
Li(s) Li + (aq) + e
now we have the cell potential for the overall reaction
Cu(s) + 2 Li + (aq)
Ecell = +3.382 V
Cu 2+ (aq) + 2 Li(s)
Note here very carefully, that the voltage was not doubled when the coefficients are doubled. However, n for this
reaction = 2. Cell voltages are therefore independent of stoichiometry. The stoichiometric information is stored in the
value of n.

Chemistry 3830 Lecture Notes

Page 82

Dr. M. Gerken

silver/zinc cell
From the table we get the half-reactions:
Zn 2+ (aq) + 2 e Zn(s)
Ag + (aq) + e Ag(s)

E = 0.763 V
E = +0.80 V

We now combine them in such a way as to get a positive overall cell potential. This means we must reverse the zinc
equation, making it the anode (where the oxidation will take place.)
E = +0.763 V
Zn(s) Zn 2+ (aq) + 2 e
now we have the cell potential for the overall reaction
2Ag(s) + Zn2+(aq)
Ecell = +1.563 V
2Ag+ (aq) + Zn(s)
These examples should be sufficient to show how cell potentials are determined for standard reduction potentials. Remember
that voltaic cells must have a positive cell potential.

5.1.10 Other uses for the table of electrode potentials


In this course, we will not be concerned much with constructing electrochemical cells. But we do want to discuss the
chemical behaviour of the elements. We can use the data in the Table directly in the description of chemical properties.

a)

Spontaneity of reaction

Combine any two standard reduction reactions by reversing one of the couples and adding the resulting potentials together.
If the E cell is positive, the reaction as written is spontaneous. If negative, the reverse reaction is spontaneous.

b)

Identifying reducing agents

Any species with a negative Eelectrode will tend to be a reducing agent, and elements with large negative electrode
potentials are strong reducing agents.

c)

Identifying oxidizing agents

Any species with a positive Eelectrode will tend to be an oxidizing agent, and elements with large positive electrode
potentials are strong oxidizing agents.

Chemistry 3830 Lecture Notes

5.2

Page 83

Dr. M. Gerken

Latimer diagrams

The main disadvantage of the Table of Redox Potentials is that it is organized by the voltage of the process, rather than by
the chemical species involved. This is great if you want to build a battery, or to use an electrochemical process in an analytical
technique, e.g. to monitor the concentration of a given species. But it is not great if you want to understand the complex redox
chemistry of an element such as iron, copper or manganese. To address this issue, Latimer devised a very helpful way to
display redox potentials for the elements. These are given in an Appendix in the text by Shriver, Atkins and Langford.
We illustrate both the need for, and the method of construction, of Latimer diagrams with the following example: What
happens when Fe(s) is reacted with a strong, non-oxidizing, acid (for example, hydrochloric or perchloric acid)? Now there
are two common oxidation states of Fe, 2+ and 3+. We thus need to decide which of these two species is predicted by
thermodynamics to form in acid solution?
Reminder: thermodynamic data will predict which reactions ought to occur, but cannot determine whether they
happen at an observable rate or not. Most of the redox reactions of inorganic compounds are rapid reactions, but there are
many times when thermodynamics predicts more than one possible product, and where the actual product is selected by the
rate of reaction. We will say a bit more about this later on. The example dealt with here can be decided unambiguously by
thermodynamics.
From the table of reduction potentials, we pick out the data that is available for iron. We then remember that the free energies
of chemical reactions, i.e. G, obey algebra (often called Hess' Law of Heat Summation). This allows us to combine the two
equations to obtain the unknown redox potential relating iron in the 3+ state to the element. The calculations are summarized:
Reaction
Potential
G = nFE
2+
Fe + 2e Fe
-0.440 V
-2 F -0.440
Fe3+ + e- Fe2+
+0.771 V
-1 F +0.771
Fe3+ + 3e- Fe
NOT +0.331V!!!!
= +0.109 F
= -3 F -0.036
Thus we have obtained the desired potential, the so-called skip potential, for the direct conversion of elemental iron to
iron(III):
E = -0.036 V
Fe3+ + 3e- Fe
and we can answer the original question. When Fe is dissolved in acid, we consider the reverse of reactions 1 and 3. We see
that the potential for the formation of Fe(II) is greater than
for the formation of Fe(III). Thus, even though the
3+
formation of Fe3+ is overall favored, the reaction has an even
+0.036 V
Fe
3+
Fe
Fe
greater tendency to stop at the Fe(II) stage. We can diagram
2+
G

Fe
this effectively by plotting the free energy terms (as Fe
+0.440 V
2+
multiples of the Faraday constant) as a function of oxidation
Fe
state. In effect, the reaction will stop in the "well" at Fe2+,
0
+I
+II
+III
because it would cost energy to rise up from the Fe(II) state
oxidation state
up to the Fe(III) state.
Latimer diagrams are convenient for the display of all redox potentials relating to a given element, including the skip
potentials. Thus the three redox potentials we have considered so far for iron are displayed as follows:
-0.036

+0.771

Fe

3+

Fe

2+

-0.440

Fe

For clarity I have added arrows to the Latimer diagram. Thus the numbers as written are standard reduction potentials, i.e.
the sign of the redox potential refers to the reaction as it proceeds from left to right. If you need to consider the reverse
reaction, as in our question above, the sign must be reversed. Latimer diagrams do not normally have such arrows, so that you
need to remember this convention!

5.2.1. Construction and use of the Latimer diagram for copper


Let's look at another example of such reasoning. Consider the two half-reactions:
Reaction
Potential
G = nFE
+
Cu + e Cu
+0.520 V
-1 F 0.520
Cu2+ + e- Cu+
+0.159 V
-1 F 0.159
Cu2+ + 2e- Cu
NOT 0.679!
= -0.679F
-2 F +0.340

Chemistry 3830 Lecture Notes

Page 84

Dr. M. Gerken

Cu

The free energy diagram is plotted at right, and the Latimer diagram becomes:

+0.340

+1.8

Cu

a)

3+

Cu

2+

+0.159

Cu

Cu

2+

Cu

+0.520

Cu

+I

+II

Oxidation of elemental copper

What happens when Cu(s) is placed in water or acid? The reverse of the reactions leading to Cu+ and Cu2+ are both
negative! Under standard conditions, copper will not be oxidized by acid! Remember that in your lab experiment on the
reactions of metals, you had to use HNO3 to dissolve copper wire. In other words, we needed to add a reagent capable of
supplying the energy required to overcome the barrier depicted in the above graph. This depends on the additional redox halfreaction:
E = +0.803 V
2 NO3- + 4 H+ + 2 e- N2O4 + 2 H2O
The overall reaction is:
Cu + 2 NO3- + 4 H+ Cu2+ + N2O4 + 2 H2O E = +0.463
You may also ask why the product of the nitric acid oxidation is Cu(II) rather than Cu(I). We need to work out the
potential for an overall reaction to form Cu(I):
Cu + NO3- + 4 H+ Cu+ + N2O4 + H2O E = +0.283
However, the overall reaction to Cu+ has E > 0; since this is a less positive EMF, the reaction proceeds directly to Cu2+ and
does not halt at Cu+. This is of course also clear from the free-energy diagram above.
NOTE: be sure that you can obtain such balanced redox reactions from the three redox half-reactions involved in this question.
Past experience has shown that even the best students consistently fail this task!

b)

Disproportionation of Copper (I)

Copper(I) in acid solution has an interesting property. Consider the possible reactions of Cu+: it can be reduced to Cu or
oxidized to Cu2+. Let us consider both of these processes.
Reaction
Potential
G = nFE
+
Cu + e Cu
+0.520 V
-1 F +0.520
Cu+ Cu2+ + e-0.159 V
-1 F -0.159
2Cu+ Cu + Cu2+
+0.3611
= - 0.361F
-1 F +0.361
NOTE: The combination of these two redox half reactions describe a balanced redox reaction, i.e. effectively the final reaction
describes an electrochemical cell. Whenever this is the case, the potentials obtained via the free energy calculation is the same
as just adding the redox potentials (with the correct sign!) together.
What we have demonstrated is that Cu+ is unstable towards disproportionation, the process in which a compound
undergoes an autoredox reaction to produce forms of the element with higher and lower than the original oxidation state. The
reverse of this equation, the conversion of Cu and Cu2+ to 2 Cu+ is called comproportionation. Thermodynamics tells us that
for copper, the disproportionation reaction is product-favored, whereas comproportionation is not product-favored.

5.2.2 Generalized Latimer diagrams


Consider the Latimer diagram for iron in acid solution shown below in the form that it is found in SAL. It is extended by a
branch describing iron in the presence of both 1 M [CN-] and 1 M acid. The potentials change because coordination with CNligands alters the free energy of both the Fe(II) and Fe(III) species.
Acidic solution
+III
+II
0
Fe 3+

0.771

-0.44

Fe 2+

Fe

-0.04

3
[Fe(CN)6 ] -

0.361

-1.16

[Fe(CN)6 ] 3+
We remember from our discussion of acid/base chemistry that Fe in acid solution is a short-hand representation for the aqua
complex, [Fe(OH2)6]3+, while Fe2+ is really [Fe(OH2)6]2+. Thus the two branches of the diagram actually reflect identical redox
2

Chemistry 3830 Lecture Notes

Page 85

Dr. M. Gerken

processes for iron in the presence of two ligand systems. We can now see at a glance that the hexacyano complex of Fe(III) is
more stable than the aqua ion with respect to reduction to Fe(II). Also, the oxidation of Fe(0) to Fe(II) is considerably more
favorable in the cyanide/acid mixture than in aqueous acid. A Latimer diagram can thus be extended to include any number of
related redox systems. We could just as well construct a branch in which iron was coordinated to ammonia or chloride ions,
for example.
In general, anything which alters the free energy of the system will change the redox potentials. The following factors all
affect the size of G :
(1)
Concentration
(2)
Temperature
(3)
Other reagents which are not inert
(4)
pH (a special case of (3))
In practice, the most important of these two examples for aqueous element chemistry is pH changes. It has become
conventional to construct Latimer diagrams for the two extremes of pH = 0 and pH = 14 (respectively 1 M acid and 1 M base).
This is shown for the element manganese below:
+VII

+VI

+V

+IV

+III

+II

acidic solution
MnO4-

0.90

HMnO4-

1.28

(H3MnO4)

2.9

MnO2

0.95

Mn3+

1.5

Mn2+

-1.18

Mn

basic solution
MnO4-

0.56

MnO42-

0.27

MnO43-

0.93

MnO2

0.146

Mn2O3

-0.234

Mn(OH)2

-1.56

Mn

Note that dramatic differences in redox potentials occur for these two sets of conditions. The reason for this is that large
numbers of H+ and OH ions are usually involved in the redox half reactions. Anything that affects the concentration of these
ions will therefore have a dramatic effect on the redox potentials. This is such an important consideration that it has led to the
wide-scale use of a graphical presentation of the free energy changes that accompany redox reactions. We now consider such
Frost diagrams.

5.3

Frost diagrams

Frost diagrams are essentially the same as the graphs of free energy against oxidation state, where G is given in units of nF
(i.e. in Volts). They are simply quantitative versions of the graphs we have already been considering. We demonstrate the
construction of complex Frost diagrams for the element manganese, which has as complicated a redox chemistry as any
element known.

5.3.1 Constructing the Frost diagram for Manganese


A Frost diagram relates the free energy of any given redox state to the energy of the elemental form. We calculate the
value nE for the overall redox couple:
X(N)/X(0)
and plot each couple against the oxidation state N. The oxidation state scale must be linear, even if the element does not exist
in all possible oxidation states. We therefore need to calculate the value nE for each species in the Latimer diagram. This is
most economically done step-wise, as shown in the table below. Thus we start by calculating the two-electron process between
Mn2+ and Mn, which works out to 2.36 V. For the next step, we recognize that Mn+3 to Mn2+ is an additional one-electron
step, so that +1.5 V must be added to the previous step to get the value that relates the energy of Mn3+ to that of elemental
manganese.
+VII

+VI

+V

+IV

+III

+II

acidic solution
MnO4-

0.90

HMnO4-

1.28

(H3MnO4)

2.9

MnO2

0.95

Mn3+

1.5

Mn2+

-1.18

Mn

basic solution
MnO4-

0.56

MnO42-

0.27

MnO43-

0.93

MnO2

0.146

Mn2O3

-0.234

Mn(OH)2

-1.56

Mn

Chemistry 3830 Lecture Notes

Page 86

Dr. M. Gerken

OxSt

Species

Calculation

value of nE
(V)
0.0

Mn

+II

Mn2+

2 1.18

2.36

+III

Mn3+

1 1.5 + 2.36

0.86

Frost Diagram for Manganese


6
(MnO4)-

5
(HMnO4)4

MnO2

1 0.95 + 0.86

0.09

+V

H3MnO4

1 2.9 + 0.09

2.99

+VI

HMnO4

1 1.28 + 2.99

4.27

+VII

MnO4

1 0.90 + 4.27

5.17

nE (V)

+IV

H3MnO4

Acid
1

Base
Mn

MnO2

The results of our calculation are plotted on the graph


shown at the right. In addition, the graph also includes
the results of a similar type of calculation using the
Latimer diagram for basic solution above. This allows
us to directly compare the redox behavior of the element
in acid and basic solution.

II

III

-1

IV
Mn3+

VI

(MnO4)2-

VII
(MnO4)-

-2
Mn2+
-3

NOTE: you should test your ability to construct Frost


diagrams by doing the base calculation and thereby
confirming the points on the graph for basic solution.

(MnO4)3MnO2

Mn(OH)2
Mn2O3
-4

Oxidation State

5.3.2 Reading the Frost diagram for


manganese
Most reactions in aqueous solution occur between the extremes of 1 M acid and 1 M base. For neutral solution, a line
approximately intermediate between the two lines plotted will be observed. It is rarely necessary to actually calculate the
intermediate state in order to explain what happens in a neutral-solution reaction. There are several characteristic features of
Frost diagrams that you should be aware of in order to be able to interpret chemical behavior quickly from the graphs. (NOTE:
there is unfortunately no universal agreement as to whether the oxidation state in such diagrams are plotted as increasing from
left to right, or vice versa. The two text books used for Chemistry 2810, Shriver-Atkins-Langford and Rayner-Canham, differ
in this regard. In these notes I will consistently use the most-positive-at-right convention (that used in Shriver-AtkinsLangford), and I will provide alternatives to all the reversed diagrams presented in Rayner-Canham's book.

a)

Identifying oxidizing and reducing agents

Species lying high on the diagram are oxidizing agents towards species on their left, while species high on the diagram act
as reducing agents to other oxidizing agents on their right. Another way to visualize this is by considering the lines connecting
the higher and lower lying species. If the line has a postitive slope, the higher-lying species is an oxidizing agent. If the line
has a negative slope, the higher-lying species is a reducing agent.
Thus for Mn in both acid and base solution, MnO4 is an oxidizing agent (the line has a positive slope), being reduced to
several possible manganese species of lower oxidation state. In acid solution, the remaining forms of manganese down to Mn3+
are all potential oxidizing agents, while in base the lowest oxidizing agent is MnO2. Elemental manganese, (i.e. manganese
metal) is a reducing agent (the line has a negative slope), being itself oxidized most readily to Mn2+ in acid solution, and Mn2O3
in base. In base only, Mn(II) can also act as a reducing agent.

b)

Identifying strong and weak agents

Once we have identified potential oxidizing and reducing agents from the slopes of the lines, their relative strength can be
determined by the steepness of the slope of the lines. The steeper the slope, the stronger the agent. Thus MnO4 is a much
more powerful oxidizing agent in acidic solution than in basic solution. On the other hand, metallic manganese is a slightly
stronger reducing agent in basic solution.

Chemistry 3830 Lecture Notes

c)

Dr. M. Gerken

Page 87

Identifying redox products (unreactive redox states)

Species at the bottom of the graph have low free-energy, thus little tendency to react. The lowest species on the graph are
the thermodynamic final product(s) of the redox reactions involving that element. Note that many Frost diagrams display a
thermodynamic well. This is the case for manganese, for which the well is Mn2+ in acid and Mn2O3 (i.e. Mn(III)) in base.

d)

Identifying species likely to undergo disproportionation

If a species lies above the line connecting its neighbors, it is thermodynamically unstable towards disproportionation. This
has been described as a point lying along a concave line. For example, in basic solution MnO43 lies on a point which is above
the line connecting MnO42 and MnO2. This means that the reaction:
2 MnO43 + 2 H2O MnO42 + MnO2 + 4 OH
is predicted to be product-favored.

e)

Identifying species likely to undergo comproportionation

Species likely to undergo comproportionation to a third species are located to left and right of a point which lies below the
line connecting the two species. Thus in acid solution, a mixture of MnO2 and Mn are expected to react together to form Mn2+.
The rate of this reaction may be hindered by the insolubility of both species, but thermodynamically it is favored. Similarly, in
base, a mixture of MnO2 and Mn(OH)2 should comproportionate to Mn2O3. Note an important difference between
disproportionation and comproportionation. The former process renders a single species unstable, in that an autoredox reaction
can occur at any time. Comproportionation, on the other hand, requires both reactants to be present at the same time, and if
one is missing the other species remains stable.

5.3.3 An overview of transition metal redox behavior in the first period


One of the great advantages of the Frost concept is the ability to compare the relative behavior of different species. We are
now finally ready to take a more detailed look at the redox behavior of the transition element, which was one of the goals we
set for ourselves when we introduced redox tools. Consider the following Frost diagram for all the first-row transition
elements in acid solution. Despite the large number of compounds, the voltages of the redox processes are quite different, so
that the points do not on the whole obscure each other.

Chemistry 3830 Lecture Notes

Dr. M. Gerken

Page 88

We see that all the metals are potent reducing agents, with the exception of copper, for which the oxidized forms have a
higher free energy than the element. The reducing strength of the metals goes down smoothly from calcium to nickel,
across the period, with nickel being only a mild reducing agent. The jump to copper is fairly large, but its behavior is
consistent with the trend towards weaker reducing power - copper simply has none.
The earlier transition elements favor the +3 oxidation state as the most stable form (bottom of the diagram), while for the
latter elements +2 is more stable, sometimes (as for cobalt) considerably more stable.
The elements in the middle of the series - Mn and Fe - have the largest range of accessible oxidation states. But the free
energies of the highest oxidation states of these elements are extremely high, and they are all potent oxidizing agents. In
fact, only Ti in its highest state (+4) has virtually no oxidizing power (if we ignore the calcium and scandium ions, which
immediately form stable noble-gas configurations forms and are extremely stable and totally non-oxidizing.)

5.3.4 Trends in the redox behavior down Group 6


The Frost diagram at the right picks on one of the elements in the above
diagram, still in acid solution, and compares its behavior to the remainder of the
transition elements in this group. This may be taken as fairly representative of the
redox trends down any d-block group. Note that this diagram has not included
the chemical form of the species involved. It is
The highest oxidation state for Cr, +6, is strongly oxidizing. For Mo, +6 is mildly
oxidizing, but for W it is completely non-oxidizing.
The most stable oxidation states are +3 for Cr, +4 for Mo, and +6 for W. Higher
oxidation states become more favorable in the 2nd and especially the 3rd transition
series.
In the +5 state, the energy of Mo and W species are the same, which leads to very
similar behavior for the two elements in this oxidation state.

Chemistry 3830 Lecture Notes

5.4

Page 89

Dr. M. Gerken

Kinetics of redox processes

A detailed consideration of the kinetics of redox processes is beyond the scope of this course. However, in order to make
sense of redox reactions, particularly those of the non-metals, we will require to take account of a few kinetic principles.

5.4.1 Overpotentials for gas formation


Reactions which produce gases, especially H2, O2 and N2 often have an overpotential associated with them, and tend to
be kinetically slow. Overpotentials are simply the extra energy required to overcome activation energies for reactions to occur
at appreciable rates. Consider the reaction to initiate the reduction of H3O+ to H2:
H3O+ + e- 4H2 + H2O
Thus the overpotential is simply the voltage needed to sustain a particular rate of electron transfer.
Overpotential (V) for gas evolution at 25 C
100 A/m2
10 A/m2
Electrode
H2
O2
H2
O2
Platinized Pt 0.015
0.398
0.030
0.521
Smooth Pt
0.024
0.721
0.068
0.85
Cu
0.479
0.422
0.584
0.580
Ag
0.475
0.588
0.762
0.729
Au
0.241
0.673
0.390
0.963
Graphite
0.599
0.778
Sn
0.856
1.076
Pb
0.52
1.090
Zn
0.716
0.746
Cd
0.981
1.134
Hg
0.9
1.0
Fe
0.403
0.577
Ni
0.563
0.353
0.747
0.519
Source: International Critical Tables, 6, 339 (1929).

H2
0.040
0.288
0.801
0.875
0.588
0.977
1.223
1.179
1.064
1.216
1.1
0.818
1.048

1000 A/m2
O2
0.638
1.28
0.660
0.984
1.244

0.726

H2
0.048
0.676
1.254
1.089
0.798
1.220
1.230
1.262
1.229
1.254
1.1
1.292
1.241

10000 A/m2
O2
0.766
1.49
0.793
1.131
1.63

0.853

For the simple reaction of hydrogen evolution by a metal dissolving in water or acid, the overpotential is ~0.6 V. This
explains why Zn, Fe, Ni and Pb do not evolve hydrogen gas when placed in water. Although the E for Zn is 0.76 V in 1 M
acid, this is reduced considerably when the [H3O+] is 1 10-7 in neutral solution. [HINT: use the Nernst equation!] Thus we
found in the lab that Zn did not react with boiling water, but produced H2 rapidly from 6 M H3O+.
ASIDE: The presence of overpotentials is often exploited in the design of electrochemical experiments and industrial cells.
Thus redox reactions can be studied at voltages where H2 or O2 evolution from water should be occurring, but does not due to
the overpotential. This is where the table from Harris becomes useful in finding an electrode material with in this case the
highest possible overpotentials. On the other hand, there is a great deal of interest in photoelectrochemistry, in which
sunlight is harnessed directly in a suitable electrochemical cell to split water into H2 and O2. Such cells require electrodes with
the lowest overpotential, and thus they often use platinized Pt (Pt covered with a colloidal deposit of fresh Pt on the surface) as
the electrode material.

5.4.2 Outer-sphere electron transfer


If a redox process can occur with no change in the atomic composition of the reacting species, electron transfer is often
very fast, and the rates of reaction closely follow the thermodynamic predictions of G. For example, consider the reaction:
[Fe(phen)3]3+ + [Fe(CN)6]4- [Fe(phen)3]2+ + [Fe(CN)6]3This is an example of a reaction which proceeds by an outer-sphere electron transfer, in which the electron is transferred
between the two species much as a baton is between two participants in a relay race.

5.4.3 Inner-sphere electron transfer


If a redox reaction requires the transfer of atoms from one reactant to the other, reaction tends to be much slower due
to significant activation energies associated with the atom transfer process. The mechanism of such reactions are inner-sphere
electron transfers, in which the composition of the coordination sphere of a complex changes during the reaction:
[CoCl(NH3)5]2+ + [Cr(OH2)6]2+ + 5 H2O + 5 H+ [Co(OH2)6]2+ + [CrCl(OH2)5]2+ + 5NH4+

Chemistry 3830 Lecture Notes

Dr. M. Gerken

Page 90

The inner-sphere mechanism is common for redox reactions involving oxoanions. For example, the reduction of
oxoanions by NO2- occurs by attack of N on the oxygen atom of the oxoanion:
NO2- + OCl- NO3 + ClThe rate of reduction are found to be:
ClO4- < ClO3- < ClO2- < ClOand
ClO4- < SO42- < HPO42Thus the lower the oxidation state of the central atom, the faster the reaction is found to be. Why? Because the O-E bond is
strongest for the highest oxidation state, and this bond must be broken for the reaction to proceed. Further evidence in favor
of this hypothesis is the effect of size:
ClO4- < BrO4- < IO4The strength of the bond is reduced as the central atom size is increased.

5.4.4 Noncomplementary redox reactions


Very slow kinetics are observed when the change in oxidation states in the oxidizing and reducing agent are not the same,
e.g.:
2 Fe3+ + Tl+ 2 Fe2+ + Tl3+
If this reaction is to proceed stepwise, then one Fe3+ reacts with one Tl+ to give a very unfavorable Tl2+ion. Alternatively the
reaction must be fully termolecular, requiring an activated complex containing two Fe3+ ions and a Tl+ ion. This is also very
unlikely, so the reaction tends to be slow.

5.5

Pourbaix diagrams

We now wish to discuss the fate of the metallic elements in aqueous solution in somewhat more detail. This ties in with a lot
of the work in the lab so far, and has important implications particularly for environmental chemistry.

5.5.1 Oxidation of the elements by water


The basic reaction we will consider is:
M + H2O M+ + H2 + OHThis is the reaction involved in oxidation of metals by water, i.e. the process commonly known as rusting. The tables of
standard reduction potentials tells us for which metals this reaction goes in 1 M H3O+. We must allow for the overpotential,
~0.6 V. We must also correct for pH if we want to discuss neutral solutions!
Consider Mg:
E = -2.37 V is the standard reduction potential in 1 M acid, i.e.
Mg + 2H3O+ Mg2+ + 2H2O + H2
E = +2.37
2+
[H ][Mg ]
0.0591
Now apply the Nernst equation:
log 2 + 2
E = E0
[H ]
2
0
.
0591
1
For pH 7 this becomes:
E = 2.37
log 7 2 = +1.956
2
(10 )
So here is a reaction that has enough voltage to overcome the overpotential, so long as a fresh surface is kept exposed to
the water such that the oxidation can proceed. This latter condition is what protects Al: a tough coating of Al2O3 protects
bulk aluminum from air and water oxidation. This is also why the Mg had to be put in boiling water in the lab for it to react at
appreciable rates. A good rule of thumb then is a metal with a reduction potential of ~ {0.4 + 0.6} = 1V or greater will
oxidize in water at appreciable rates in the absence of air. NB: dissolved O2 will change this picture, of course! We all know
that iron rusts in aerated water.)

5.5.2 Reduction of elements by water


The redox reaction involved in acid solution is:
O2 + 4 H+ + 4 e- 2 H2O E = +1.23 V
In basic solution it becomes:
O2 + 2 H2O + 4 e- 4 OH- E = +0.40 V
Some strong oxidizing agents, e.g. Co3+, are reduced by water. The overall reaction in acid is:
4 Co3+ + 2 H2O 4 Co2+ + O2 + 4 H+ E = 0.59 V
This reaction is thus at the overpotential boundary. It becomes fully favored in basic solution, with E = 1.42 V

Chemistry 3830 Lecture Notes

Dr. M. Gerken

Page 91

5.5.3 Stability field of water


We can combine the reduction and oxidation of water with the pH
With overpotential
+1.6
dependence and construct a diagram which represents the stability field
for water. This is shown in the diagram at right.
pH = 4
+1.2
The region boxed in the middle is the normal range found for
O 2/H 2O
natural waters, in which water is not oxidized or reduced, and those are
pH = 9
the pH ranges found in the various types of natural waters. Specific
+0.8
E/pH zones for different kinds of environments are indicated by the
Fresh surface water
circles drawn into the stability field diagram.
+0.4
Ocean water
Well aerated natural waters near the surface contain enough
Bog
E(V)
Organic-rich
dissolved oxygen to get close to the oxidation of water to O2.
water
0
lake water
Eutrophic lake water contains sufficient dissolved organic matter to
Organic-rich
approach closely the H+ reduction line. Ocean waters are relatively
waterlogged
-0.4
soils
basic, and may be oxidizing if saturated in dioxygen, or reducing if
Organic-rich
saline water
H2O/H2
saturated in organic matter (i.e. in stagnant lagoons, etc.)
-0.8
Fresh waters are considerably more acidic (because of dissolved
With
carbon dioxide), but again they can be oxidizing if saturated in oxygen,
overpotential
or reducing if too much organic matter is consuming all the oxygen.
This acidity is greatly enhanced in bogs and organic-laden soils, due to
2
4
6
8
10
high humic acid content. (Humic acids are complex organic acids
pH
occurring in the soil and in bituminous substances formed by the
decomposition of dead vegetable matter.)
The stability field for water
These are strongly reducing conditions, and explain the formation
of CH4 in marshes. Methane was first discovered from this source; Alessandro Volta was one of the first to identify this gas.
Ammonia and hydrogen sulfide as well as the inflammable phosphine, PH3, can also emanate from swamps. (There are many
rumors of eerie glows emanating in misty swamps; such tales are likely rooted in phosphine being released and burning above
the surface of the waters.) If you remember that bogs and marshes release such compounds, in other words very reduced
chemical compounds, it is should help you to remember the acidic, reducing character of bog water.

5.5.4 Pourbaix diagrams


The final type of diagram we want to consider for redox chemistry is a type of combination redox/pH predominance
diagram developed by the French electrochemistry Pourbaix. The diagrams are usually named after him. These diagrams are
closely related to the stability field for water just discussed. Indeed, most Pourbaix diagrams include either the main E/pH
lines from the water diagram, or both the main and overpotential lines. What these diagrams add are the relationships between
the redox activity and the Brnsted acidity of the elements. They are thus related both to Latimer diagrams, and to
predominance diagrams for acid/base reactions.
Consider as an example the Pourbaix diagram for
iron. Note that some of the lines from the stability field
of water are drawn into this diagram, since this diagram
applies to aqueous solutions of iron compounds.
The vertical axis plots the standard reduction
potential, and the horizontal the pH. Remember what
a predominance diagram was: simple vertical
boundaries where the most abundant species
altered.
The bottom of the diagram refers to reduced species,
i.e. Fe(s), or to the type of conditions that lead to
reduction.
The top of the diagram to oxidized species and/or
oxidizing conditions.
Vertical lines indicate changes in acid base chemistry
independent of E, e.g. Fe3+/Fe(OH)3.
Horizontal lines indicate redox changes unaffected
by pH, e.g. Fe/Fe2+ below pH 6.
In the more general case, the lines slope, since both E
and [H+] or [OH-] affect the redox process.

Chemistry 3830 Lecture Notes

Dr. M. Gerken

Page 92

To test your ability to read Pourbaix diagrams, see if you can find answers to the following:
(a) The form of iron which is the strongest oxidizing agent: FeO42- at [H+] 1 M
(b) The form of iron which is the strongest reducing agent: Fe(0), elemental iron
(c) The predominant form at pH = 7 and E = 0.0 V:
Fe(OH)3 predominates, but close to Fe2+/Fe(OH)2
3+
2(d) E for (acid) reduction of FeO4 to Fe :
1 M, 0 pH = 0 so E = 2.2 V
(e) E for reduction of Fe2+ to Fe(s):
E = 0.5 V (This MUST be an acid process. Why?)
On the diagram, dashed lines d and e represent respectively the normal and overpotential for oxygen evolution according to:
2 H2O 4 H+(aq) + O2 + 4 eE = +1.229 V.
The actual E for hydrogen evolution is given by f, while the overpotential is given by line g:
E = 0.00 V
2H+ + 2e- H2

5.5.6 How elements behave in natural waters


We can now take any of the other Pourbaix diagrams, and compare them to the natural water limits, and predict what
forms may exist in various
environments.
Several
lanthanide
and
actinide
element Pourbaix diagrams
are shown in the figure at
right.
What form will Yb take in
natural
lake
water?
Answer: Yb(OH)3
Can
uranium
be
solubilized in sea water?
Answer: Yes as UO22+.
Would you expect to find
cerium metal free in
nature?
Answer:
Plutonium is highly toxic,
as well as being strongly
radioactive. What would
happen if plutonium was
released into a lake or a
stream?
Answer:
Imagine that plutonium
oxides from a nuclear
weapons processing centre
had been dumped into a
small, well aerated lake
where the pH = 6-8 and E
= 0.0-0.5 V. Over the
course of 20 years, the
lake converted to a bog,
where the pH = 4 and E =
0.1 V.
Discuss the
environmental concerns in
the initial stages and the
final stages of the lake
"storage", remembering
that plutonium would be a
serious toxic hazard if it
entered the food stream.
Answer:

Chemistry 3830 Lecture Notes

Dr. M. Gerken

Page 93

Note that there are some limitations to this approach. First of all, the concentration of the elements in natural waters is
often much lower than standard conditions. Furthermore, these data are only valid for pure water plus the element in
question. For example, on this basis we would say gold cannot exist in sea water. However, it does, as a chloro complex,
which is soluble. This is due to a complexation equilibrium. Nevertheless the combined E/pH diagrams provide a very
comprehensive insight into the behavior of the elements in aqueous solution, and this is clearly the most important set of
conditions relevant to the terrestrial environment.

Anda mungkin juga menyukai