
Chambers, Holliday. Modern inorganic chemistry
.pdf96 |
ACIDS AND BASES: OXIDATION AND REDUCTION |
||
oxygens, each |
-2. total |
— 8 ; overall charge on ion = -L hence |
|
the |
oxidation |
state of |
Mn = x 7, i.e. manganate(Vll) ion. (b) |
Chlorine in the chlorate ions, ClO^ —there are three oxygens,each
— 2, total -6; overall charge |
on ion = —1, hence the oxidation |
state of Cl = -1-5 and the ion |
is a chlorate(V) ion.Chromium in |
the dichromate ion C^O^"; there are seven oxygens each —2, total = — 14; overall charge on ion = —2, hence chromium atoms share 12 formal positive charges and so the oxidation state of chromium is 4-6, and the ion is dichromate(VI).
Oxidation states can be used to establish the stoichiometry for an equation. Consider the reaction between the manganate(VII) (permanganate) and ethanedioate (oxalate) ions in acidic solution. Under these conditions the MnOjfaq) ion acts as an oxidising agent and it is reduced to Mn2 r (aq), i.e.
Mnv n + 5e~ -> Mn2+
The full half equation is
(i) MnO4(aq) + 8H3O+ + 5e~ -> Mn2+(aq) + 12H2O
The ethanedioate (oxalate) ion C2O|"(aq) is oxidised to carbon dioxide, i.e.
(ii)C2O|-(aq)-+2CO2 + 2e~
To maintain electrical neutrality in the reaction we need to multiply
(i) by 2 and (ii)by 5, ten electrons being transferred. The overall reaction then becomes
2MnO4(aq) + 16H3O+ |
+ 10e" -> 2Mn2 + (aq) + 24H2O |
||
5C2O2"(aq)-> 1QCQ2 + |
10g~ |
|
_______ |
2^6"4laq)"TT6H^OT"~+ 5C2Of:r^ |
2MnIT(aq) |
||
|
|
-f |
24H2O + 10CO2 |
Consider also the oxidation of iron(II) ions by dichromate(VI) ions in acidic solution. The Q^O^" is reduced to Cr3+(aq)
Cr2O?"(aq) -h 6e~ -» 2Cr3+ (aq)
The full half equation is
(i)Cr2O?~(aq) -f 6e" 4- 14H3O+ ^ 2Cr3+ (aq) -h 21H2 O The Fe2 + (aq) is oxidised to Fe3^(aq), i.e.
(ii)Fe2 + (aq) -» Fe3 + (aq) -he"
|
ACIDS AND BASES: OXIDATION AND REDUCTION |
97 |
|||||
Thus the equation for the reaction is: |
|
|
|
|
|
||
Cr2Or(aq) + 6Fe2+(aq) + 14H3O+ |
-> 2Cr3+(aq) + |
|
|
||||
|
|
|
|
|
21HO |
|
|
STANDARD REDOX |
POTENTIALS |
|
|
|
|
||
When the reaction between zinc and copper(II) sulphate was carried |
|
||||||
out in the form of an electrochemical |
cell (p.94), a potential |
differ- |
|
||||
ence between the copper and zinc |
electrodes was noted. |
This |
|
||||
potential resulted from the differing tendencies of the two metals to |
|
||||||
form ions. An equilibrium is established |
when any metal is placed |
|
|||||
in a solution of its ions. |
|
|
|
|
|
|
|
The enthalpy changes A/f involved in this equilibrium are (a) the |
|
||||||
heat of atomisation of the metal, (b) the ionisation energy of the |
|
||||||
metal and (c) the hydration enthalpy of the metal ion (Chapter 3). |
|
||||||
For copper and zinc, these quantities have the values (kJ moPJ): |
|
||||||
|
H eat of |
Sum of 1st and 2nd |
Hydration |
AH |
|
|
|
|
atomisation |
ionisation energies |
enthalpy |
|
|
||
|
|
|
|
||||
Cu |
339 |
2703 |
|
-2100 |
+942 |
|
|
Zn |
126 |
2640 |
|
-2046 |
+720 |
|
|
For the equilibrium M(s) ^ M2+(aq) + 2e , it might then be |
|
||||||
(correctly) assumed that the equilibrium for copper |
is further to the |
|
|||||
left than for zinc, i.e. copper has less |
tendency to form ions in |
|
|||||
solution than has zinc. The position of equilibrium (which depends |
|
||||||
also on temperature and concentration) |
is related |
to the relative |
|
||||
reducing powers of the metals when two different metals in solutions |
|
||||||
of their ions are connected (as shown in Figure 4.1 for the copper- |
|
||||||
zinc cell) ; a potential difference is noted because |
of the differing |
|
|||||
equilibrium positions. |
|
|
|
|
|
|
Since it is not possible to measure a single electrode potential, one electrode system must be taken as a standard and all others measured relative to it. By international agreement the hydrogen electrode has been chosen as the reference:
This electrode, shown diagrammatically in Figure 4.4, is assigned zero potential when hydrogen gas at one atmosphere bubbles over platinised platinum in a solution of hydrogen ions of concentration 1 mol P * (strictly, at unit activity).
98 ACIDS AND BASES. OXIDATION AND REDUCTION
^=—H5
Figure 4.4. The hydrogen electrode
Standard redox potentials for metals (usually called electrode potentials), E"9", are measured at 298 K relative to a standard hydrogen electrode for the pure metal in a solution containing Imoir1 of its ions and at pH = 0 (i.e. containing ImolT1 hydrogen ions). (The importance of pH is stressed later, p. 101.) If the metal is a better reducing agent than hydrogen the metal will
Table 4.2
STANDARD REDOX POTENTIALS OF SOME COMMON METALS
|
Reaction |
|
E~~(V\ |
||
Li + (aq) + e |
|
~» Li(s) |
|
|
|
K+ (aq) + e~ -»K(s) |
|
|
|||
Ba2+(aq) + 2e~ -> Ba(s) |
|
|
|||
Ca2 + (aq) + |
2e~~ |
-»Ca(s) |
|
-2.87 |
|
|
|
>Na(s) |
|
-2.71 |
|
Mg2+(aq) |
2e~ |
* Mg(s) |
|
-2.37 |
|
Al3+ (aq) + |
3e |
|
* Al(s) |
|
-1.66 |
Zn2 + (aq) + 2e |
|
-* Zn(s) |
|
-0.76 |
|
Fe2+(aq) + 2e |
|
* Fe(s) |
Increasing |
-0.44 |
|
|
|
|
* Ni(s) |
reducing |
-0.25 |
|
|
|
* Sn(s) |
power |
-0.14 |
Pb2 + (aq) + 2e |
|
H. Pb(s) |
|
-0.13 |
|
|
|
|
»Fe(s) |
|
-0.04 |
|
|
|
H20(l) |
|
0.00 |
Cu*' + (aq) + |
2e |
|
-»• Cu(s) |
|
+ 0.34 |
Ag+ (aq) + e~ |
* Ag(s) |
|
+0.80 |
||
Au3*(aq) -I- |
3f |
|
- Hg(s) |
|
+ 0.86 |
|
-* Au(s) |
|
+1.50 |
ACIDS AND BASES: OXIDATION AND REDUCTION 99
lose electrons more readily than hydrogen and, therefore, be negative with respect to the hydrogen electrode. Table 4.2 gives the standard redox potentials of some common metals. By convention the oxidised state is always written on the left-hand side.
Redox half-reactions are often written for brevity as, for example, Li"" + e' Li with the state symbols omitted. The electrode system represented by the half-reaction may also be written as Li+ / Li. The standard redox potentials for ion-ion redox systems can be determined by setting up the relevant half-cell and measuring the potential at 298 K relative to a standard hydrogen electrode. For example, the standard redox potential for the half-reactions
can be determined by measuring the potential of a half-cell, made 1 molar with respect to both iron(II) and iron(III) ions, and in which a platinised platinum electrode is placed, relative to a standard hydrogen electrode at 298 K.
Saturated KCl solution
Mercury
Mercury(l)
chloride WJj~~ Hole
^^ sleeve Ring
Figure 4.5
For many purposes the hydrogen electrode is not convenient and it can be replaced by another cell of known standard electrode Potential. A well-known example is the calomel cell shown in figure 4.5.
A number of redox potentials for ion-ion systems are given in Table 4.3; here again, state symbols are often omitted.
100 ACIDS AND BASES: OXIDATION AND REDUCTION
Table 4.3
Increasing oxidising power
REDOX POTENTIALS FOR ION-ION SYSTEMS (ACID SOLUTIONS)
|
|
£(V) |
Sn4+ (aq) + 2e" -^Sn2^(aq) |
|
+ 0.15 |
yI2(s) -f e~ -> I~(aq) |
|
+ 0.54 |
Fe3 + (aq) + e~ -> Fe2 + (aq) |
|
+ 0.76 |
iBr2(l) + e~ -> Br"(aq) |
Increasing -(-1.07 |
|
IOj(aq) -f (jH3O^ ^ 5e"" -+ iI2(s) + 9H2O |
reducing |
4-1.19 |
02(g) + 4HJO + 4e" -^6H 20 |
power |
+1.23 |
Cr2O?"(aq) -f 14H3O+ -f 6 r -*2Cr3+(aq) |
|
+ 1.33 |
H-21H2O |
|
|
iCl2(g) + e'- -* CT(aq) |
|
+ 1.36 |
MnO^Caq) f 8H3O+ -h 5^~ -M2+(aq) |
|
|
+ 12H2O |
|
+ 1-52 |
iF2(g) 4- e~ - F-(aq) |
|
+ 2.80 |
THE EFFECT OF CONCENTRATION AND
TEMPERATURE ON POTENTIALS
Changes in ion concentration and temperature influence redox potentials by affecting the equilibrium
M(s) ^ M*+(aq) + ne~
The change in the redox potential is given quantitatively by the Nernst equation :
RT
where £ is the actual electrode potential, E^ isthe standard electrode potential, R the gas constant, Tthe temperature in K, F the Faraday constant and n the number of electrons.
Substituting for R and F and for a temperature of 298 K this equation approximates to :
The redox (electrode) potential for ion-ion redox systems at any concentration and temperature is given by the Nernst equation in the form
^ |
RT |
rOxidised state"! |
+ |
~nF |
ge [TedSoedltate] |
ACIDS AND BASES: OXIDATION AND REDUCTION 101
(Note that the equation for metal-metal ion systems is a special case of this general equation since the reduced state is the metal itself and the concentration of a solid is a constant and omitted from the equation.)
THE EFFECT OF CHANGE OF LIGAND AND pH ON REDOX POTENTIALS
The data in Tables 4.2 and 4.3 refer to ions in aqueous acid solution; for cations, this means effectively [M(H2O)X]"+ species. However, we have already seen that the hydrated cations of elements such as aluminium or iron undergo 'hydrolysis' when the pH is increased (p. 46). We may then assume (correctly), that the redox potential of the system
Fe3 + (aq) + e~ -> Fe21aq)
will change with change of pH. In fact, in this example, change of pH here means a change of ligand since, as the solution becomes more alkaline, the iron(III) species in solution changes from [Fe(H2O)6]3+ to [Fe(OH)3(H2O)3] (i.e. iron(III) hydroxide). The iron(II) species changes similarly. The redox half-reaction then becomes
[Fe(OH)3(H2O)3] + e~ -> [Fe(OH)2(H2O)4] + OH~
for which E^ is —0,56 V. compared with E^ = 4- 0,76 V in acid solution; thus in alkaline conditions, iron(II) becomes a good reducing agent, i.e. is easily oxidised.
When the water ligands around a cation are replaced by other ligands which are more strongly attached, the redox potential can change dramatically, for example for the cobalt(II)-cobalt(III) system we have
(i)[Com(H2O)6]3+ +* -- > [Co"(H2O)6]2 + :E^ = + 1.81V
(ii)[Com(NH3)6]3+(aq) + e~ -> [Con(NH3)6]2+(aq):£^ - +0.1 V
(iii)[Corn(CN)6]3-(aq) + e~ -> [Co"(CN)5(H2O)]3-(aq) + CN~:
E*= - 0.83 V
Half-reaction (i) means that Co(II) in aqueous solution cannot be oxidised to Co(III); by adding ammonia to obtain the complexes in (ii), oxidation is readily achieved by, for example, air. Similarly, by adding cyanide, the hexacyanocobaltate(II) complex becomes a sufficiently strong reducing agent to produce hydrogen from water!
102 ACIDS AND BASES: OXIDATION AND REDUCTION
When either hydrogen ions or hydroxide ions participate in a redox half-reaction, then clearly the redox potential is affected by change of pH. Manganate(VII) ions are usually used in well-
acidified |
solution, where (as we shall see in detail later) they oxidise |
|
chlorine |
ions. If the pH |
is increased to make the solution only |
mildly acidic (pH = 3-6), |
the redox potential changes from 1.52 V |
|
to about |
1.1 V, and chloride is not oxidised. This fact is of practical |
use; in a mixture of iodide and chloride ions in mildly acid solution. manganate(VII) oxidises only iodide; addition of acid causes oxidation of chloride to proceed.
Other important effects of ligand and pH changes on redox potentials will be given under the appropriate element.
USES OF REDOX POTENTIALS
Reaction feasibility predictions
When the e.m.f. of a cell is measured by balancing it against an external voltage, so that no current flows, the maximum e.m.f. is obtained since the cell is at equilibrium. The maximum work obtainable from the cell is then nFE J, where n is the number of electrons transferred, F is the Faraday unit and E is the maximum cell e.m.f. We saw in Chapter 3 that the maximum amount of work obtainable from a reaction is given by the free energy change, i.e.
— AG. Hence
-AG =nFE
or
AG - - nFE
For a half-cell under standard conditions this becomes
where AG^ and E* are the free energy and redox potential under standard conditions. In Chapter 3 we also noted that for a reaction to be energetically feasible the total free energy must fall, i.e. AG must be negative. An increase in free energy indicates that the reaction cannot proceed under the stated conditions. The relationship AG = -nFE can now be used to determine reaction feasibility. Let us consider first the oxidation of iron(II) to iron(III) by bromine in aqueous solution, i.e.
2Fe2 + (aq) 4- Br2(aq) -> 2Fe3 + (aq) + 2Br~(aq)
ACIDS AND BASES: OXIDATION AND REDUCTION 103
We can determine the energetic feasibility for this reaction from the two half-reactions:
Reaction |
E~~(\) |
|
AG"^ — —nFE^ |
||
Fe3 + (aq)-4- e |
-»Fe2 + (aq) |
+ 0.76 |
-1 x 96487 x (+0.76) |
||
Fe2'(aq)-» Fe3+ (aq) + e~ |
-0.76 |
-1 |
x 96487 |
x (-0.76) |
|
;. (i)2Fe2 +(aq )^2Fe3 "(aq) 4- 2e~ |
-0.76 |
-2 x 96487 |
x (-0.16) |
||
iBr2(aq) + ?~ |
->< Br(aq) |
+1,07 |
AG(7; = + 146.7kJ |
||
-1 |
x 96487 |
x (+ 1.07) |
|||
;. (ii) Br2(aq) -\-2e~ -»2Br~(aq) |
1.07 |
-2 |
x 96487 |
x ( + 1.07) |
|
|
|
|
AG^ = - 206.5 kJ |
Hence (i) and (ii) give
2Fe2 + (aq) 4- Br2(aq) -» 2Fe3 + (aq) -f 2Br~(aq) AG = AG^ -f GJ^}
- + 146.7 -h ( - 206.5) = - 59.8 kJ
Thus the reaction is energetically feasible and does indeed take place. It is interesting at this point to investigate the reasons why iron(II) ions in aqueous solutions are quantitatively estimated by titration using potassium manganate(VII) (permanganate) when chloride ions are absent but by potassium dichromate(VI) when chloride ions are present. The data for the oxidation of chloride ions to chloride by (a) manganate(VII) and (b) dichromate(VI) ions under standard conditions are given below:
(a) 2MnO4 (aq) -f 10Cl~(aq) + I6H3O+
|
|
|
4- 24H2O 4- 5Cl2(g) |
|
Reaction |
E~(V) |
A G * = - W J*E |
||
MnO 4 (aq)+ 5e~ -f 8H^O+ |
+1.52 |
-5 x 96487 x ( + 1.52) |
||
-> Mn-(aq) + |
12H2O |
4- 1.52 |
AG(t = - 10 x96487 |
|
(i) 2MnO4 (aq) + |
I0e~ + 16H3O* |
|||
—» 2Mn2 "^(aq) 4- 24H2O |
4-1.36 |
x (+1.52) = -1467kJ |
||
iC!2(aq) + e~ |
- > C l ~ ( a q ) |
- 1 x 96487 x (-1.36) |
||
CP(aq) -»• 4Cl2 (aq) + e" |
-1.36 |
|||
(ii) lOCl'(aq) -> 5Cl2(aq) 4- We~ |
-1.36 |
AGJ^, = — 10 |
x 96487 |
|
|
|
|
x (-1.36) = |
+ 1312kJ |
Hence (i) and (ii)give
2MnC>4 (aq) + 10C1" (aq) + 16H3O+ ->
2Mn2 + (aq) + 24H2O 4- 5Cl2(aq)
104 ACIDS AND BASES: OXIDATION AND REDUCTION
lor which AG = AG(7; + AG(%
= ( - 1467) + ( -f 1312) - - 155 kJ
Thus chloride ions are oxidised to chlorine by manganate(VII) under standard conditions
(b) |
Cr2 CH~(aq) + 6C1"(aq) + |
14H3O+ |
-> 2Cr3 + (aq) |
||||
|
|
|
|
|
|
|
4- 21H2O -f 3Cl2(aq) |
|
Reaction |
|
|
|
|
AG~* = -nh'c |
|
(i) Cr2 O= (aq) + 6e |
+ 14H3CT |
+1,33 |
jf} |
= ^6 x 96487 x (+1.33) |
|||
-> 2Cr3 '(aq) + 21H2O |
" |
|
|
|
|
||
k12|aq) + f " |
-> CP(aq) |
+1.36 |
|
|
|||
Cl'(aq) ->• K'Maq) + f ' |
|
-1,36 |
-1 |
x 96 487 x ( - 1,36) |
|||
(ii) 6C1 |
(aq) ->~3Cl2(aq) + |
6t>" |
-1.36 |
AG(T*» = -6 x 96487 x (-1,36) |
Hence (i) and (ii) give
6CP(aq)
21H2O + 3Cl2(aq)
for which AG - AGg -f
= |
( -769) ( + 787) |
= |
+ 18kJ |
Thus under standard conditions chloride ions are not oxidised to chlorine by dichromate(VI) ions. However, it is necessary to emphasise that changes in the concentration of the dichromate(VI) and chloride ions alters their redox potentials as indicated by the Nernst equation. Hence, when concentrated hydrochloric acid is added to solid potassium dichromate and the mixture warmed, chlorine is liberated.
Equilibrium constants from electrode potentials
We have seen that the energetic feasibility of a reaction can be deduced from redox potential data. It is also possible to deduce the theoretical equilibrium position for a reaction. In Chapter 3 we saw that when AG = 0 the system is at equilibrium. Since AG = — nFE, this means that the potential of the cell must be zero. Consider once again the reaction
+ Zn(s) -* Cu(s) + Zn2 f (aq)
ACIDS AND BASES: OXIDATION AND REDUCTION 105
At equilibrium at 298 K the electrode potential of the half-reaction for copper, given approximately by
must equal the electrode potential for the half-reaction for zinc, given approximately by
Thus,
Efn + ^- log10[Zn2 + (aq)] = Eg, -f ^- log10[Cu2+(aq)]
— |
z. |
Hence, |
|
loglo[Zn2 + (aq)] - loglo[Cu2 + (aq)] = (Eg, - EfJ x -|- |
|
Substituting for Eg, = |
+ 0.34, and Efn=- 0.76we have: |
Hence
2 +
=
This is in fact the equilibrium constant for the reaction
Cu2 + (aq) + Zn(s) -> Cu(s) + Zn2 + (aq)
and its high value indicates that the reaction goes effectively to completion.
Similar calculations enable the equilibrium constants for other reactions to be calculated.
Potentiometrie titrations
The problem in any quantitative volumetric analysis for ions in solution is to determine accurately the equivalence point. This is often found by using an indicator, but in redox reactions it can often