Sei sulla pagina 1di 36

Chapter 4 Oxidation-Reduction

4.1 The significance of Oxidation-Reduction

These are reactions involving a change in the oxidation state of the


reactants. This requires the transfer of electrons from one species to
another. Manahan uses the example:

Cd+2 + Fe Cd + Fe+2

This reaction is divided into two parts, the reduction part, where cadmium
receives two electrons:

Cd+2 + 2e- Cd

and the oxidation, where iron give up two electrons:

Fe Fe+2 + 2e-
The oxidation of organic matter by oxygen is also an oxidation reduction
reaction. Organic matter, here represented by CH2O, reacts with O2:

(CH2O) + O2 → CO2 + H2O

This reaction, a simplified form of the Biochemical Oxygen Demand


(BOD) reaction, consumes oxygen. Aqueous oxygen may be so depleted
that anoxic conditions result. This may lead to fish kills, and
eutrophication.

We can write the oxidation side of the above reaction:

C0H2+1O-2 → C+4 + O-2 +4e- + 2 H+ In this reaction, C starts at 0 and is


oxidized to +4

O20 + 4e- → 2 O-2 Here, O starts at 0 and is reduced to -2

We add the two reactions:


(CH2O) + O2 → CO2 + H2O The electrons cancel out, the O-2 react
with either C or H+. The “ins’ equal the
“outs.”
The foregoing reaction may be used to describe the oxidation of
Biochemical Oxygen Demand (BOD). This is of great interest with respect
to the treatment of wastewater, and with the discharge of waste into natural
water supplies. BOD is expressed as the number of mg of O2 per liter of
water required to oxidize the organic carbon in that liter of water. Various
BODs are given below:

Typical BOD for Various Wastewaters


Type of Discharge BOD (mg O2/liter of Wastewater)
Domestic Sewage 165
All Manufacturing 200
Chemicals and allied products 314
Paper 372
Food 747
Metals 13
From: Chemistry of the Environment, 2nd ed., Spiro and Stigliani, Prentice Hall, 2003. page 309
From the table above, we can see that Food manufacturing produces high
strength wastewater, while Metals manufacturing produces much lower
strength wastewater, at least with respect to BOD.

Some other important Redox reactions that take place in water:

Fe(OH)3(s) + 3 H+ + e- → Fe+2 + 3 H2O

In a reducing environment (oxygen deficient), insoluble Fe+3 is reduced to


soluble Fe+2. This soluble iron becomes a contaminant in drinking water
supplies, and must be removed.

NH4+ + 2 O2 → NO3- + 2H+ + H2O

The oxidation of ammonium to nitrate, which can be assimilated by algae in


water is another important redox reaction that takes place in water.
The tendency of chemical species to undergo redox reactions is governed by
the “electron activity.” This is also true when the chemical reactions are
catalyzed by microbes, as in the case of BOD removal and the oxidation of
ammonia to nitrate, shown above.

The “electron activity” can be described in terms analogous to “hydrogen


ion activity,” i.e., pH. We can define a pE function that is analogous to pH.
That is, we can state the electron activity as the negative logarithm of the
electron activity. In this context, high pE values correspond to an oxidizing
environment, while low pE values correspond to a reducing environment.
Note that a high pE value means a low electron activity, and a low pE value
means a high electron activity. This, of course, corresponds to a low pH
representing a high hydrogen activity, and vice versa.

We can further complicate matters by considering species that can exchange


both hydrogen ions and electrons. Manahan gives the following example:

Fe(H2O)6+3 Fe(H2O)5OH+2 + H+
and

Fe(H2O)6+3 + e- Fe(H2O)6+2

In the first case, one of the water molecules that has complexed with the
Fe+3 loses an H+ (this is how iron acidifies soil and water), however, the
oxidation state of the iron remains the same. In the second case, the water
molecules are not affected; the Fe+3 gains an electron, forming the more
soluble species Fe+2.

However, the real world tends towards maximum complexity. The transfer
of electrons is frequently accompanied by the transfer of hydrogen ions.
Manahan gives the example:

Fe(H2O)6+2 Fe(OH)3 + e- + 3 H2O + 3 H+

At pH 7, the reaction given is favored. Soluble Fe II is oxidized to form Fe


III hydroxide, which is insoluble.
Environmental Chemistry, 7th Edition, Stanley E. Manahan. Lewis Publishers, Boca Raton, 1999.

Note the following:


At the surface, where gases can exchange and the O2 concentration is
highest, the pE is highest and, generally, elements are in higher oxidation
states.
• Carbon is in the +4 state (CO2)
• Nitrogen is in the +5 state (NO3-)
• Sulfur is in the +6 state (SO4-2)
• Iron is +3 (Fe(OH)3 ,although limited by solubility)
• Manganese is +4 (MnO2) -- This is important with respect to its
ability to bind metals and phosphate.

At the bottom, where the O2 concentration is very low:


• Carbon is in the 0 oxidation state (CH4)
• Nitrogen is in the -3 oxidation state (NH4+)
• Sulfur is in the -2 oxidation state (H2S)
• Iron is in the +2 state (much more soluble)
• Manganese is in the +2 state, releasing sorbed heavy metals and
phosphate

It is also important to note that the figure depicts chemical species at


equilibrium. In natural water systems, this is seldom the case.
4.2 The Electron and Redox Reactions: All redox reactions involve changing
the oxidation states of at least some of the species that take part in the
reaction.

To start with, we consider Fe(III) and Fe(II), in a medium that is sufficiently


acidic to prevent the precipitation of Fe(OH)3. If we treat the solution with
hydrogen gas, in the presence of an appropriate catalyst, Fe(III) can be
reduced (made to gain electrons) to Fe(II):

2 Fe+3 + H2 2 Fe+2 + 2 H+

As the reaction proceeds to the right, hydrogen is oxidized (made to lose


electrons). The oxidation state (or number) of hydrogen is changed from 0
to +1. Note that any element, in its pure state (in this case, H2) has an
oxidation number of 0. Fe(III) gains the electrons lost by the hydrogen.
Thus, Fe(III) is reduced to Fe(II). We can write the half-reactions:

2 Fe+3 +2 e- 2 Fe+2
and
H2 2 H+ + 2 e-

When we add the two half-reactions together, we get the overall reaction:

2 Fe+3 + H2 2 Fe+2 + 2 H+

Note that the electrons on each side of the half-reactions must cancel; there are
not electrons shown in the overall reaction.

The reaction shown above could go to either the left or right; we can deduce the
direction from the initial half-reactions. If we assume that the initial activities of
Fe+2, Fe+3, H+ are 1 (i.e., they are present at 1 molar concentration), and that H2
is present at 1 atmosphere pressure, hydrogen would be oxidized to H+, and Fe+3
would be reduced to Fe+2. We can visualize the half-reactions by separating
them schematically, as in figure 4.2.
Environmental Chemistry, 7th Edition, Stanley E. Manahan. Lewis Publishers, Boca Raton, 1999.

We see that hydrogen is supplied to a hydrogen electrode, where it is oxidized to


H+. Ions travel through the salt bridge to the other cell so that electroneutrality
is maintained. Electrons travel through the connection wire. If we insert a
voltmeter in the wire (with sufficiently high impedance), the flow of electrical
current will stop. The reaction will also stop. The voltage registered will
indicate the relative tendencies of the two half-reactions to occur, and the sign of
the voltage will indicate the direction in which the overall reaction will occur.
The left hand cell where, in this case, electrons are released from the hydrogen,
will be negative relative to the right hand cell. The right hand cell will collect
-
electrons, using them to drive the reaction to the right (that is, Fe+3 + e → Fe+2).
So the right hand electrode will have a positive potential. The difference in the
potential of the right-hand side and that of the left-hand side is the Driving
Force of the reaction. Since we’ve specified that all the reactants are present at
unit activity, the potential difference will be 0.77 V. This is the standard
-
reduction potential (E0) for the reaction Fe+3 + e → Fe+2. We don’t have to
consider the left side of the equation because the standard reduction potential of
the standard hydrogen electrode (SHE) is 0.00 Volts. If we used a different
left-hand cell, we would have to subtract the two potentials to find the overall
cell potential.
4.3 Electron Activity and pE: In environmental chemistry we use pE and pE0
instead of E and E0. This is because we are interested in the behavior of
redox equilibria over many orders of magnitude. This method is almost
exactly analogous to pH. Recall the rigorous definition of pH:

pH = - log(aH+) Of course, we usually assume that aH+ [H+]

Analogously, pE is defined as:

pE = -log(ae-) where ae- is the “activity” of electrons. Since there are no


electrons in solution and, if there were, we wouldn’t know how to
approximate them, this is a bit of a thorny concept. Of course, there are
also no free protons in solution, either. In order to make some sense of this
concept we’ll have to stagger down the path for a while, until it becomes
clear.

First, we will define:


E E
pE = = (at 25° C)
2.303 R T 0.0591
F

and

E° E°
pE° = = (at 25° C)
2.303 R T 0.0591
F
R is the molar gas constant, 8.3144 joule/˚K-mole, T is the absolute
temperature, and F is the Faraday constant, 96,493 Coulombs/ equiv. The
2.303 term is the conversion from natural log to base 10 log. It will take a
minute to explain why it’s included, so please be patient.

Now, we will state that for the half-reaction:


-
2 H+ (aq) + 2 e H2 E0 = 0.00 Volts and pE0 = 0.00
This is just a statement of the Standard Hydrogen Electrode (SHE) as the
reference point for all other reduction potentials. This statement means that,
when H+(aq) is present at unit activity and is in equilibrium with hydrogen
gas at 1 atmosphere pressure, the electron activity is exactly 1.00 and the pE
is 0.0. Recall that the log (1) = 0. If we were to reduce the H+(aq) activity
to 0.1, then the electron activity would become 10, and the pE would be
-1.0. This will make more sense as we continue.

4.4 The Nernst Equation: Look back at figure 4.2 on p. 103. If the
concentration of Fe+3 ion is higher than the concentration of Fe+2, the
tendency of the electron deficient Fe+3 to acquire electrons will make the
electrode potential more positive. If the converse is true, than the potential
will become less positive. So the concentration of the various species
affects the electrode potentials.
-
Fe+3 + e → Fe+2 E0 = 0.77 Volts and pE0 = 13.2
We want to quantify the effect of the change in concentration of the
reactants, so we write:

2.303 R T [ Fe +3 ] 0.0591 [ Fe +3 ]
E=E + 0
log +2
=E +
0
log
nF [ Fe ] n [ Fe +2 ]

Now, because we can express the Nernst equation as

1 [ Fe+3 ]
pE = pE + log
0

n [ Fe+ 2 ]

This is why we used the “natural log to base 10 log conversion earlier.”
Later, we’ll see why we have a natural log involved.

If, for example, [Fe+3] is present at 10-3 M, and [Fe+2] is present at 10-5 M,
then the value of pE is:
1 [10−3 ]
pE = pE + log
0
= 15.2
n [10−5 ]

We can see that increasing the electron deficient species (Fe+3) relative to
the electron-rich species (Fe+2) increased the cell potential, and pE. Since
pE = -log(ae-), this means that the “activity” of the electron has decreased.
All of this should make sense.

4.5 Reaction Tendency: Whole Reaction from Half-Reactions:

In the case of the previous reaction, where Fe reduction was paired with
hydrogen oxidation, the whole reaction potential is the same as the Fe cell
reaction potential. For cases other than reduction by hydrogen, we must
find the difference between two cells in order to find the whole reaction
potential, and hence the reaction direction and equilibrium.

For the case of a solution of Cu+2 flowing through a lead pipe, the following
reaction can be written:
Cu+2 + Pb Cu + Pb+2

For the half-reaction Cu+2 + 2 e- Cu, pE0 = 5.71, while for the reaction
Pb+2 +2 e- Pb, pE0= -2.13.
If we assume that all reactants are present at unit activity, we can obtain the
overall reaction potential by subtracting the two reactions. The reason we
subtract is that, one of the reactions must go in the opposite direction from
what is written above. That is, one of the species must be oxidized in order
for the other to be reduced.

Cu+2 + 2 e- Cu pE0 = 5.71

-(Pb+2 +2 e- Pb) -(pE0 = -2.13)

Cu+2 + Pb Cu + Pb+2 pE0 =7.84


The positive value of pE0 tells us that the reaction will go to the right, as
written. We will see why this is so presently. Of course, if the activities of
the two ionized species are not unity, we would write the equation:

1 [Cu +2 ]
pE = pE + log
0

n [ Pb + 2 ]

Note that the activities of the solid Cu and Pb are always 1.

4.6 The Nernst Equation and Chemical Equilibrium: Realize that as long as we
have inserted a volt meter into the circuit shown in Fig. 4.3 on p. 107, the
reaction cannot proceed, as electrons cannot flow through the volt meter. If
we were to bypass the meter with a wire, the reaction could proceed until it
reached equilibrium. The equilibrium constant for the reaction would be
written:
[ Pb+2 ][Cu] [ Pb+2 ]
K eq = +2
or, since the activities of both Pb and Cu are 1, K eq =
[ Pb][Cu ] [Cu + 2 ]

Note that the Equilibrium constant is the inverse of the logarithmic term in
the Nernst equation. We can now write:

1 [Cu +2 ] 1 1
pE = pE + log
0
= pE 0
+ log
n [ Pb + 2 ] n K eq

To find the equilibrium concentration, we must find the point where pE=0:

1 1 1 [ Pb +2 ] 1
pE = 0.00 = pE + log
0
= 7.84 − log +2
= 7.84 − log K
n K eq 2 [Cu ] 2

We can solve this to yield


1
( )
7.84 = log K or n pE 0 = log K
2

4.7 The Relationship of pE to Free Energy: So far, we’ve said nothing about
where the Nernst equation comes from, or why it includes a logarithmic
term, which must be converted from natural log to base 10 log. We won’t
say too much here either. Recall from basic chemistry that for the reaction:
aA + bB cC +dD

{C}c {D}d
∆G = ∆G + R T ln
0
a b
= ∆G 0
+ R T ln Q
{ A} {B}

Where G is the Gibbs Free Energy, and G0 is the Gibbs Free Energy of
Formation. When G is equal to 0, the reaction is at equilibrium. In order
for this to happen, G0 must be equal to -RT ln Q. When the activities of
A, B, C, and D are of the correct magnitude to cause -RT ln Q to be equal
to G0, then Q is equal to the equilibrium constant, Keq. So, at equilibrium:
∆G 0 = − R T ln K eq
Now, we can see why we have to include the conversion from natural log to
base 10 log. Backing up, and continuing,

∆G = ∆G 0 + R T ln Q Eqn 1

We also know that the work obtained from a chemical reaction is:

∆G = −n F E
and
∆G 0 = −n F E 0
We know this because n is the number of electrons per mole of
reactant, F is the electric charge of a mole of electrons, and E is the
potential driving this electrical current. We know that the work obtained
from an electric current is:
w =V i
where V is the potential, in joules/coulomb (volts), and i is the electrical
current in coulombs/second (amperes). In the present case, this work, in
joules/second (Watts) is the Free Energy, G.

Now, if we divide equation 1 by - nF

∆G ∆G 0 RT
= + ln Q
− nF − nF − nF

we obtain

joules
8.3144 298° K
RT deg mole 0.0591
E = E0 + ln Q = E 0 − 2.303 ln Q = E 0 − log Q
− nF coulomb n
n 96,493
equiv
Because we know that, at equilibrium, G = 0, so E = 0, and Q must be
equal to Keq. So:

∆G 0 RT 0.0591
E =
0
= ln K eq = log K eq
−n F n F n

Now we know that:


0.0591 0.0591 0.0591 0.0591 Q
E = E0 − log Q = log K eq1 − log Q = − log
n n n n K eq

Keeping in mind that E has the opposite sign of G, if Q/Keq is >1, than G is
>0, and E <0. So the reaction proceeds to the left. Conversely, if Q/Keq is <1,
than G is < 0, and E >0. So the reaction proceeds to the right, as written.

4.8 Reactions in Terms of One Electron-Mole: Different reactions entail the


transfer of different numbers of electrons. Your text uses the two examples:
NH4+ + 2O2 NO3- + 2 H+ + H2O pE0 = 5.85

and

4 Fe+2 + O2 + 10 H2O 4 Fe(OH)3(s) + 8 H+ pE0=7.6


In the case of the first example, we are oxidizing N from the -3 state to the +5
state. Thus, N loses 8 electrons. In the second case, each Fe loses 1 electron.
All this means is that, in order to compare the free energy of the two reactions,
and determine which reaction is favored thermodynamically, we need to re-write
the reactions so that we can calculate G on a the basis of one electron mole.
Since we’ve already established that G= -nFE = -2.303n RT(pE0), we can
calculate the free energy change for each reaction. However, if we want to
compare the free energy change per mole of electrons transferred, we can restate
the reactions so that they are based on the transfer of one electron. That is:

1/8 NH4+ + ¼O2 1/8 NO3- + ¼H+ + 1/8 H2O pE0 = 5.85
and

Fe+2 + ¼ O2 + 5/2 H2O Fe(OH)3(s) + 2 H+ pE0=7.6

Now, each reaction involves the transfer of one electron. This allows us to
consider these equilibria based on the number of electrons (oxygen molecules)
available to perform oxidation. Let’s look at the equilibrium expression for
the first reaction:
Recall that log(Keq) = npE0. If n = 1, then log(Keq) = pE0. So

1 1
[ NO3− ] 8 [ H + ] 4
K eq = 1 1
+
[ NH ] PO24
4
8

and the second:


[ H + ]2
K eq = 1
+2
[ Fe ]PO24
The two Keqs are, of course, not identical. Note that, in this case, I have set the
activity of Fe(OH)3 to 1, as it is insoluble. Some Fe+3 would really be in
solution. If we were treating this in a rigorous fashion, we would include the
solubility of Fe+3, which is a function pH. Things quickly become complicated
and messy.

Look at Table 4.1 on page 111 of your text. The table lists a series of important
reactions in natural water, and their associated pE0. Also shown is pE0(W). This
value is also frequently expressed as EH, the potential measured in a water
system. As the reactions given above show, the H+ concentration in water
affects the reaction. This value is the pE0 for a pH of 7.

Some Problems: Based on the above, one might get the impression that we can
just measure the potential present in a natural water system and, based on its
value, calculate the equilibrium concentrations of all the species present.
Unfortunately, nothing is ever that simple. Most of the reactions described in
Table 4.1 are calculated from thermodynamic data, rather than from actual
measurements. In reality, most of the reactions do not behave in a Nernstian
manner. First of all, the kinetics are often so slow that the reactions do not reach
equilibrium. Secondly, the reactions are frequently not reversible. Even more
unfortunately, the definition of a reversible reaction is not clear-cut. In fact,
there are at least 3 different definitions that apply. They are:

1. Chemical Reversibility. Consider the following electrochemical cell:

Pt/H2/H+,Cl-/AgCl/Ag

As a galvanic cell (no external voltage applied) the reaction proceeds as:

H2+2 AgCl → 2 Ag + 2H+ + 2 Cl- with E0 = 0.222 V

If sufficient voltage is applied to drive the cell the other way:

2 Ag + 2H+ + 2 Cl-→ H2+2 AgCl


The reaction will proceed as written, with no new reactions. Thus, the reaction
is chemically reversible.

On the other hand,

Zn/H+,SO4-2/Pt

is not chemically reversible. The zinc electrode is negative compared to the Pt


electrode, so the reaction

Zn→ Zn+2+ 2e-

proceeds, while at the Pt electrode,

2H++ 2e- → H2

so the net cell reaction is:


Zn + 2H+→ Zn+2 + H2

However, the reaction does not go in reverse. Instead,

at the Zn electrode: 2H++ 2e- → H2

and at the Pt electrode: H2O →1/2 O2 + 2 H+ + 2e-

In short, we electrolyze water. Therefore, this reaction is not chemically


reversible.

2. Thermodynamic Reversibility. A process can only be thermodynamically


reversible if it proceeds at an infinitesimal rate, and is always in equilibrium.
This never occurs in nature, however, some situations come closer than
others. It is important to note that a cell that is chemically irreversible is also
thermodynamically irreversible. But a chemically reversible cell may, or
may not approximate thermodynamically reversibility.
3. Practical Reversibility: This is not an absolute term. In the context of
electrochemistry, it means that the reaction obeys the Nernst equation. That
is, the reaction must never be very far from equilibrium, and the system must
respond to any perturbation by attaining equilibrium in a reasonably short
time. If this definition seems less than satisfying, it should.

So, measuring the potential in a natural water system will give us a number.
However that number usually cannot be related to any absolute values of
concentration, as found from the Nernst equation. However, it can give an idea
of the relative position of the system. The following figure illustrates this:
From: Chemistry of the Environment, 2nd ed., Spiro and Stigliani, Prentice Hall, 2003.
4.9 The Limits of pE in Water:

Water oxidized: 2 H2O O2 + 4 H+ + 4 e-

Water reduced: 2 H2O + 2 e- H2 + 2 OH-

We can rewrite these in terms of pE:

1
pE = pE + log pO 24 [ H + ]
0

pE = 20.75 − pH
and, given that the activity of H2 gas, at one atmosphere, is 1, as is the
activity of water:

(
pE = pE 0 + log [ H + ] )
pE = 0.00 − pH = − pH
What these mean is that if pE exceeds the first value, water is oxidized, and
if it is less than the second value, water is reduced. Thus, water can only
exist in the pE region bounded by the two equations. In practice, water
electrolyzed very slowly, and in the case of chlorine addition, the pE can
exceed the theoretical value calculated.

4.10 pE Values in Natural Water Systems: As stated, it is usually not possible to


obtain an accurate EH value by measuring the potential of a natural water
system. However, we can make some amusing calculations that at least
suggest what reality might be.
If water is in equilibrium with the atmosphere, then PO2 = 0.21 atm, and
[H+] = 1 x 10-7. We can plug this into the equation for the oxidation of
water:

pE = 20.75 + log(0.21¼ x 1 x 10-7) = 13.8

So this is the pE of oxygen saturated water at pH = 7.


Now, imagine a reducing environment, where microbes are fermenting CO2
to CH4 by the reaction:

1/8 CO2 + H+ + e- 1/8 CH4 + ¼ H2O

Your text chooses to assume that PCH4 = PCO2 and that the pH is 7. We can
write the Nernst equation:
1
P 8
CO2 [H + ]
pE = 2.87 + log 1
= 2.87 + log[ H + ] = −4.13
P 8
CH 4

Now, we can plug this pE value back into

pE=-4.13 = 20.75 + log(O2¼ x 1 x 10-7)

This yields an oxygen partial pressure of 3.0 x 10-72 atm. Since we can’t possibly
measure a value this low, we can never confirm this (although it’s very
unlikely). However, we can assume that, under conditions so reducing as to
ferment carbon dioxide to methane, the dissolved oxygen concentration is very,
very low.

So we can see that pE, or EH, can give us some idea of what reactions are most
likely to occur, and how they relate to one another. But drawing quantitative
conclusions about precise equilibria is, at best, risky.

Potrebbero piacerti anche