978 94 011 0691 7
978 94 011 0691 7
978 94 011 0691 7
ELECTROCHEMI~TRY
SECOND EOITION
PHILIP H. RIEGER
Preface ix
Chapter 1: Electrode Potentials 1
1.1 Introduction 1
1.2 Electrochemical Cell Thermodynamics 5
1.3 Some Uses of Standard Potentials 13
1.4 Measurement of Cell Potentials 27
1.5 Reference and Indicator Electrodes 31
1.6 Ion-Selective Electrodes 35
1.7 Chemical Analysis by Potentiometry 39
1.8 Batteries and Fuel Cells 44
References 54
Problems 55
Chapter 2: The Electrified Interface 59
2.1 The Electric Double Layer 59
2.2 Some Properties of Colloids 68
2.3 Electrokinetic Phenomena 73
2.4 Electrophoresis and Related Phenomena 81
2.5 Electrode Double-Layer Effects 85
2.6 Debye-Hiickel Theory 90
References 105
Problems 106
Chapter 3: Electrolytic Conductance 109
3.1 Conductivity 109
3.2 Conductance Applications 125
3.3 Diffusion 128
3.4 Membrane and Liquid Junction Potentials 136
References 146
Problems 147
Chapter 4: Voltammetry of Reversible Systems 151
4.1 Diffusion-Limited Current 152
4.2 Experimental Techniques 165
4.3 A Survey of Electroanalytical Methods 174
4.4 Cyclic Voltammetry 183
viii Contents
ix
x
SI units have been employed throughout the book. References to
older units are given in footnotes where appropriate. In most cases, the
use of SI units eliminates unit conversion problems and greatly
simplifies numerical calculations. The major remaining source of
units ambiguity comes from concentrations. When a concentration is
used as an approximation to an activity, molar units (mol L-l) must be
used to conform to the customary standard state. But when a
concentration acts as a mechanical variable, e.g., in a diffusion problem,
the SI unit, mol m- 3 , should be used. The mol m- 3 concentration unit is
equivalent to mmol L-l and, in a sense, is a more practical concentration
scale since voltammetric experiments often employ substrate
concentrations in the millimolar range.
Several topics have been added or expanded in the second edition.
In particular, coverage of microelectrode voltammetry has been much
expanded, and previous discussions of steady-state voltammetry with
rotating-disk electrodes have been modified to include microelectrodes;
spectroelectrochemistry (electron spin resonance and infrared
spectroscopy) is now discussed as an aid to deducing mechanisms of
electrode processes; the discussion of cyclic voltammetry has been
expanded to include adsorption effects and derivative, semi-derivative
and semi-integral presentation; the discussion of organic
electrosynthesis has been considerably expanded; and many new
examples of work from the literature have been added to illustrate the
techniques discussed.
It has been said that no book is ever finished, it is just abandoned.
The truth of that aphorism is never more apparent than to an author
returning to a previously abandoned project. There has been more than
one instance when I have been appalled at the state in which I left the
first edition of this book. I have labored mightily to correct the errors of
commission and at least a few of the errors of omission, but the awful
truth is that the book must be abandoned again with topics which should
have been covered more completely or more clearly.
I am particularly grateful to my wife, Anne L. Rieger, for her
patience in listening to my problems and for her encouragement in
times of discouragement. My colleague, Dwight Sweigart, has been an
invaluable source of expertise and encouragement during the
preparation of the second edition. I am indebted to Petr Zuman for some
valuable suggestions after publication of the first edition and to Nancy
Lehnhoff for a stimulating discussion of microelectrodes which greatly
clarified the presentation. I am deeply grateful to Barbara Goldman of
Chapman and Hall for her thoughtful suggestions and timely support in
this project.
Thanks are still due to those who helped with the first edition: to
David Gosser, who listened to my ideas and offered many helpful
suggestions-the cyclic voltammogram simulations of Chapters 4 - 6 are
his work; to my colleagues at Brown who offered advice and
xi
encouragement, most particularly Joe Steim, John Edwards, and Ed
Mason; to Bill Geiger, who provided a stimulating atmosphere during
my 1985 sabbatical and gave some timely advice on electro analytical
chemistry; to James Anderson of the University of Georgia, Arthur Diaz
of IBM, San Jose, Harry Finklea of Virginia Polytechnic Institute, and
Franklin Schultz of Florida Atlantic University for their careful reading
of the first edition manuscript and numerous helpful suggestions.
The first edition was produced using the IBM Waterloo SCRIPT
word-processing system and a Xerox 9700 laser printer equipped with
Century Schoolbook roman, italic, bold, bold italic, greek, and
mathematics fonts. Seven years later, that system is hopelessly obsolete
and the present edition has been completely redone using Microsoft
Word on a Macintosh computer with equations formatted with
Expressionist. To maintain a semblence of continuity, the principal font
is again New Century Schoolbook. The figures all have been redone
using CA-Cricket Graph III, SuperPaint, and ChemDraw. Figures
from the literature were digitized with a scanner and edited with
SuperPaint.
Philip H. Rieger
May 1993
1
ELECTRODE
POTENTIALS
1.1 INTRODUCTION
(a)
~ T (b) (c)
~k~%%Zy;:;~?/.~::;z~2$~Z0.Z2~-':~J.?~nZlJ bWff//ff/k;wuu~
Fermi level
Figure 1.1 The conduction bands of two dissimilar metals (a) when the
metals are not in contact; (b) at the instant of contact; and (c) at equilibrium.
1
2 Electrode Potentials
higher Fermi level into the metal with the lower Fermi level. This
electron transfer results in a separation of charge and an electric
potential difference across the phase boundary. The effect ofthe electric
potential difference is to raise the energy of the conduction band of the
second metal and to lower the energy of the conduction band of the first
until the Fermi levels are equal in energy; when the Fermi levels are
equal, no further electron transfer takes place. In other words, the
intrinsically lower energy of electrons in the conduction band of the
second metal is exactly compensated by the electrical work required to
move an electron from the first metal to the second against the electric
potential difference.
A very similar process occurs when a metal, say a piece of copper,
is placed in a solution of copper sulfate. Some of the copper ions may
deposit on the copper metal, accepting electrons from the metal
conduction band and leaving the metal with a small positive charge and
the solution with a small negative charge. With a more active metal, it
may be the other way around: a few atoms leave the metal surface as
ions, giving the metal a small negative charge and the solution a small
positive charge. The direction of charge transfer depends on the metal,
but in general charge separation occurs and an electric potential
difference is developed between the metal and the solution.
When two dissimilar electrolyte solutions are brought into contact,
there will be a charge separation at the phase boundary owing to the
different rates of diffusion of the ions. The resulting electric potential
difference, called a liquid junction potential, is discussed in §3.4.
In general, whenever two conducting phases are brought into
contact, an interphase electric potential difference will develop. The
exploitation of this phenomenon is one of the subjects of
electrochemistry.
Consider the electrochemical cell shown in Figure 1.2. A piece of
zinc metal is immersed in a solution of ZnS04 and a piece of copper
metal is immersed in a solution of CUS04. The two solutions make
contact with one another through a fritted glass disk (to prevent mixing),
and the two pieces of metal are attached to a voltmeter through copper
wires. The voltmeter tells us that a potential is developed, but what is its
origin? There are altogether four sources of potential: (1) the copper-
zinc junction where the voltmeter lead is attached to the zinc electrode;
(2) the zinc-solution interface; (3) the junction between the two solutions;
and (4) the solution-copper interface. The measured voltage is the sum of
all four interphase potentials.
In the discussion which follows, we shall neglect potentials which
arise from junctions between two dissimilar metals or two dissimilar
solutions. This is not to say that such junctions introduce negligible
potentials; however, our interest lies primarily in the metal-solution
interface and solid or liquid junction potentials make more or less
constant additive contributions to the measured potential of a cell. In
§ 1.1 Introduction 3
Zn Cu
glass frit
(salt bridge)
Origins of Electrochemistry
The electrochemical cell we have been discussing was invented in
1836 by John F. Daniell. It was one of many such cells developed to
supply electrical energy before electrical generators were available.
Such cells are called galvanic cells, remembering Luigi Galvani, who in
1791 accidentally discovered that static electricity could cause a
convulsion in a frog's leg; he then found that a static generator was
unnecessary for the effect, that two dissimilar metals (and an electrolyte
solution) could also result in the same kinds of muscle contractions.
Galvani thought of the frog's leg as an integral part of the experiment,
but in a series of experiments during the 1790's, Alessandro Volta
showed that the generation of electricity had nothing to do with the frog.
Volta's work culminated in the construction of a battery (the voltaic pile)
from alternating plates of silver and zinc separated by cloth soaked in
salt solution, an invention which he described in a letter to Sir Joseph
Banks, the President of the Royal Society of London, in the spring of 1800.
Banks published the letter in the Society's Philosophical Transactions
that summer, but months before publication, the voltaic pile was well
known among the scientific literati of London.
Among those who knew of Volta's discovery in advance of
publication were William Nicholson and Sir Anthony Carlisle, who
constructed a voltaic pile and noticed that bubbles of gas were evolved
from a drop of water which they used to improve the electrical contact of
the leads. They quickly showed that the gases were hydrogen and oxygen
4 Electrode Potentials
Electrical Work
1 Because the potential of the Weston cell, 1.0180 V at 25°C, is very reproducible, it
has long been used as a standard potential source.
8 Electrode Potentials
According to convention, the free energy change for the cell reaction is
negative if the reaction proceeds spontaneously to the right and,
according to eq (1.7), the cell potential should then be positive, i.e., the
right-hand electrode (Hg) should be positive with respect to the left-hand
electrode (Cd).
1 The identification of the cathode with the reduction process and the anode with the
oxidation process is common to both galvanic and electrolysis cells and is a better
§ 1.2 Electrochemical Cell Thermodynamics 9
definition to remember than the electrode polarity, which is different in the two
kinds of cells.
1 We will use the 1 M standard state in this book, but another common choice is 1
molal, 1 mole solute per kilogram of solvent. Although activity coefficients are
unitless, they do depend on the choice of the standard state (see §2.6).
10 Electrode Potentials
Half-Cell Potentials
1 The details of this variation of potential with distance are considered in Chapter 2.
12 Electrode Potentials
Hg
The Nernst equation for the Weston cell, eq (1.20), gives us the cell
potential as a function of the standard potential and of the activities of all
participants in the cell reaction. We would like to break eq (1.20) into two
parts which represent the electrode-solution potential differences shown
in Figure 1.5. It is customary in referring to these so-called half-cell
potentials to speak of the potential of the solution relative to the electrode.
This is equivalent to referring always to the electrode process as a
reduction. When, in the actual cell reaction, the electrode process is an
oxidation, the contribution to the cell potential will then be the negative of
the corresponding reduction potential. Thus in the case of the Weston
cell, we can write for the cell potential
Ecell = -ECd 2+/Cd + EHg2S04/Hg (1.21)
Breaking the standard cell potential into two components in the same
way, the half-cell potentials defined by eq (1.21) can be written in the form
of the Nernst equation such that substitution in eq (1.21) gives eq (1.20).
Thus
ECd2+/Cd = EOCd2+/Cd - RT In (acd)
2F (aCd 2+)
E Eo RI..1 (aHg)(asol)
Hg2S0JHg = Hg2S0JHg - 2F n ( )
aHg2so•
Unfortunately, there is no way of measuring directly the potential
difference between an electrode and a solution, so that single electrode
potentials cannot be uniquely defined. 1 However, since the quantity is
not measurable we are free to assign an arbitrary standard potential to
one half-cell which will then be used as a standard reference. In effect
this establishes a potential zero against which all other half-cell
potentials may be measured.
By universal agreement among chemists, the hydrogen electrode
was chosen as the standard reference half-cell for aqueous solution
1 We will see in §2.5 that some properties (the electrode-solution interfacial tension
and the electrode-solution capacitance) depend on the electrode-solution potential
difference and so could provide an indirect means of establishing the potential zero.
§ 1.2 Electrochemical Cell Thermodynamics 13
Pt
1 See §4.2 for discussion of the problem of reference potentials for nonaqueous
solutions.
14 Electrode Potentials
12 + 2 e- -t 2 1- EOI21I- = 0.536 V
To obtain the overall cell reaction, we multiply the first
equation by 2 and subtract the second:
2 Fe3+ + 2 1- -t 2 Fe2+ + 12
EO cell =EOFe3+/Fe2+ - EOI2 /I - = 0.235 V
fl.Go = -- 2FEo
t-.Go = -45.3 kJ mol- 1
Thus the oxidation of iodide ion by ferric ion is spontaneous
under standard conditions. The same standard free energy
change should apply for the reaction under nonelectro-
chemical conditions. Thus we can use the standard free
energy change computed from the cell potential to calculate the
equilibrium constant for the reaction of Fe3 + with 1- to give Fe2 +
and 12.
In K = -t-.GoIRT
Formal Potentials
Standard potentials refer to standard states, which for solution
species are the hypothetical 1 M ideal solutions. Very dilute solutions
can be assumed ideal and calculations using standard potentials are
then reasonably accurate without activity coefficient corrections. For
electrolyte concentrations less than about 0.01 M, activity coefficients can
be computed reasonably accurately using Debye-Huckel theory (§ 2.6), but
for more concentrated solutions, empirical activity coefficients are
required.
One way around the problem of activity coefficients is through so-
called formal potentials. A formal half-cell potential is defined as the
potential of the half-cell when the concentration quotient of the Nernst
equation equals 1. Consider the Fe(III)IFe(II) couple. The Nernst
equation gives
or
E = EOFe3+IFe2+ In YF
D'f1
_ll.L
2+ D'f1T [Fe 2+]
_e__ ll.L I n - -
F YFe3 + F [Fe3+]
Thus when the concentrations of Fe3 + and Fe 2+ are equal, the last term
on the right-hand side is zero and the formal half-cell potential is
RT YFe2+
EO' = EOFe 3+IFe 2+- - I n - -
F YFe3 +
As we will see in §2.6, activity coefficients depend primarily on the total
electrolyte concentration (ionic strength) of the solution, so that in a
solution where the ionic strength is determined mostly by a high
concentration of an inert electrolyte, the activity coefficients are nearly
constant. Molar concentrations can then be used, together with formal
potentials appropriate to the medium, in calculations with the Nernst
equation. A representative sample of formal potentials for 1 M HCI04, 1
M HCI, and 1 M H2S04 solutions is given in Table A.6.
The formal potential of the Fe(III)IFe(II) couple in 1 M HCI04 is EO'
= 0.732 V, significantly different from the standard potential of 0.771 V,
18 Electrode Potentials
suggesting that the activity coefficient ratio is about 0.22 in this medium.
In a medium with coordinating anions, e.g., aqueous HCI, Fe(II), and
Fe(III) form a variety of complexes. In order to compute the half-cell
potential from the standard potential, we would have to know not only
the activity coefficients but the formation constants of all the complexes
present. The formal potential of the Fe(III)lFe(II) couple in 1 M HCI, EO'
= 0.700 V, thus differs from the standard potential both because of activity
coefficient effects and because of chloro complex formation. As long as
the medium is constant, however, the relative importance of the various
complexes is constant and the formal potential can be used as an
empirical parameter to compute the overall Fe(II)lFe(III) concentration
ratio.
Latimer Diagrams
When we are interested in the redox chemistry of an element, a
tabulation of half-cell potential data such as that given in Table A.4 can
be difficult to assimilate at a glance. A lot of information is given and it
is not organized to give a qualitative understanding of a redox system.
Nitrogen, for example, exists in compounds having nitrogen oxidation
states ranging from -3 (NH3) to +5 (N03') and all intermediate oxidation
states are represented. One way of dealing with complex systems like
this is to use a simplified diagram introduced by Latimer (HI) and
usually referred to with his name. An example of a Latimer diagram is
shown in Figure 1.7 for the aqueous nitrogen system. In the Latimer
diagram, any H+, OH-, or H20 required to balance the half-cell reaction
is omitted for clarity. Thus, if we wish to use the half-cell potential for
the N03-/N204 couple, for example, we must first balance the equation
N03-+2H++e- ~ ~N204+H20 EO=0.80V
Figure 1.7 also includes a Latimer diagram for nitrogen species in basic
aqueous solution. The Latimer diagram tells us that one-electron
reduction of nitrate ion again produces N204, but the half-cell potentials
differ by 1.66 V. The reason is clear when we write the half-cell reaction.
The difference between the half-cell reactions is
2H20 ~ 2H++20H-
the Gibbs free energy of which is -2RT In K w , 160 kJ mol-lor 1.66 eV.
N03- + H20 + e- ~ ~ N204 + 2 OH- EO = -0.86 V
of interest is formed from the element in its most stable form at 25°C and
from electrons, H+ or OH- ions, water, or other solution species. Thus we
could form HN02 from N2(g) by the following half-cell process:
2"1 N2(g) + 2 H20 -1 HN02 + 3 H+ + 3 e-
The free energy change for this process can be obtained from half-cell
potential data such as those given in Figure 1.7. Since the reductions of
HN02 to NO, NO to N20, and N20 to N2 all involve one electron per
nitrogen atom, the potentials can be added directly to obtain the
(negative) overall free energy change for the three-electron reduction;
changing the sign gives t'1.Go for the oxidation:
t'1.Go =F(1.77 + 1.59 + 1.00) =4.36 F
t'1.Go is usually expressed in kJ mol-I, but for the present purpose it is
easier to think of t'1.GolF, which is equivalent to putting !1Go in units of
electron-volts (1 eV = 96.485 kJ mol-I).
Similar calculations for the free energies of the other nitrogen
oxidation states can be done using the data of Figure 1. 7. These free
energies are plotted us. nitrogen oxidation number in Figure 1.8. Free
energy - oxidation state diagrams were introduced by Frost (1) and often
are called Frost diagrams. These diagrams were popularized in Great
Britain by Ebsworth (2) and sometimes are referred to as Ebsworth
diagrams.
:>
Q)
4
0- NH:t>H
e" 3
<I
-1
-3 -2 -1 o 1 2 3 4 5
Oxidation State
Figure 1.8 Free energy - oxidation state diagram for the nitrogen oxidation
states in acid (solid line) and basic (dashed line) solutions.
ai = Ci/Cio
so that when Cia = 1 M, ai = Ci. But when Cia = 10- 7 M, ai = 10 7 Ci.
Consider the Nernst equation for the hydrogen electrode:
E = EO _RT In (aH2f
F aw
With H2(g) at unit activity, this expression can be written
E =EO +RT In aw
F
Using the chemical standard state and EO = 0.00 V, we have
E =RIln[H+]
F
With the biochemical standard state, the corresponding expression is
E = E' + RT In (107[H+]) (1.24)
F
For equal [H+], the two equations must give the same potential;
subtracting one equation from the other thus gives
E' = -RIln 107
F
or E' = -0.414 V at 25°C. Similarly, for a half-cell reaction
A + m H+ + n e- ~ B
E' and EO are related by
E' = EO -l11B.X..ln 107
nF
or, at 25°C,
E' = EO - 0.414 min
Similar relations can be derived between· f.G' and f.Go and between
1(' and K (see Problems). For an equilibrium reaction
A+mH+ ~ B
we find that
K = 1O- 7m K (1.25)
f.G' = f.Go + mRT In 10 7 (1.26)
or, at 25°C,
f.G'IkJ mol- 1 = f.Go + 40.0 m
Conversion back and forth between chemical and biochemical
standard states is straightforward as long as the reactions involve H + or
OH- in clearly defined roles and the necessary pKa data are available.
§1.3 Some Uses of Standard Potentials 23
Potential - pH Diagrams
or
N03- + H20 + 2 e- -7 N02- + 20H-
so that dEldpH = -59 mV pH-I. The N(V)/N(III) half-cell potential is
plotted us. pH in Figure 1.9. Plots of the half-cell potentials of the
N(III)/N(O) and N(O)/N(-III) couples are also shown in Figure 1.9,
together with the 02iH20 and H201H2 couples.
The significance of the 02/H20 and H20/H2 couples is that these
define the limits of thermodynamic stability of an aqueous solution. Any
couple with a half-cell potential greater than that of the 021H20 couple is
in principle capable of oxidizing water. Similarly, any couple with a
potential less than (more negative than) that of the H201H2 couple is in
principle capable of reducing water. As it happens, there are many
26 Electrode Potentials
We see in Figure 1.9 that nitrous acid and the nitrite ion are in
principle capable of oxidizing water and are thus thermodynamically
unstable in aqueous solution. Although this instability is not manifested
because the reactions are slow, it raises the related question: What is the
thermodynamically most stable form of nitrogen at a given pH and
electrode potential? Referring to the free energy - oxidation state
diagram, Figure 1.8, we see that for both acidic and basic solutions the
points for N(-l) and N(-II) lie above the line connecting N(O) and N(-III);
these species thus are unstable with respect to disproportionation.
Similarly, the points for N(I), N(Il), N(lII), and N(IV) all lie above the
N(O) - N(V) line, so that these species are also unstable. The existence of
any of these species in aqueous solution therefore reflects kinetic
stability rather than thermodynamic stability. In fact, only N03-, NH3,
NH4+, and N2 are thermodynamically stable in aqueous solution.
Furthermore, according to thermodynamics, N03- is unstable in the
presence of NH4 + or NH3. One way of expressing the conclusions
regarding thermodynamic stability is by means of a predominance area
diagram, developed by Pourbaix (3,H5) and thus commonly called a
Pourbaix diagram.
1.5 1.5
EIV
N03-(aq)
1.0 1.0
0.5 0.5
0.0 0.0
-0.5 -0.5
-1.0 -1.0
024 6 8 ill ~ M o 2 4 6 8 W ~ M
pH pH
Figure 1.9 Potential - pH diagram Figure 1.10 Predominance area
showing the variation in half-cell diagram (Pourbaix diagram) for
potentials with pH for the nitrogen in aqueous solution. The
N(III)/N(O), N(V)/N(III), and labeled regions correspond to
N(O)IN(-III) couples. Note changes areas of thermodynamic stability
in slope of the first two lines at pH for the indicated species. Water is
3.14 (the pKa of HN02) and of the stable between the dashed lines
third line at pH 9.24 (the pKa of showing the 02/H20 and H20/H2
NH4+). Also shown (dashed lines) couples.
are the potentials of the 02/H20
and H20/H2 couples.
Potentiometers
...------IIII---..-'C"""
E
Electrometers
With the development of vacuum tube circuitry in the 1930's, d.c.
amplifiers with very high input impedances became available. Such
devices, called electrometers, are particularly well suited to
measurement of the potentials of cells with high internal resistance.
A number of different designs have been used for electrometer
circuits, but one which is particularly common in electrochemical
instrumentation is the so-called voltage follower shown in Figure 1.13.
A voltage follower employs an operational amplifier, indicated by the
triangle in the figure. The output voltage of an operational amplifier is
proportional to the difference between the two input voltages with very
high gain (> 104 ). In a voltage follower, the output is connected to the
negative (inverting) input. Suppose that the output voltage is slightly
greater than that at the positive input. The difference between the inputs
will be amplified, driving the output voltage down. Conversely, if the
output voltage is low, it will be driven up. The output is stable, of course,
when the output and input are exactly equal. Since the input impedance
§ 1.4 Measurement of Cell Potentials 31
is high (> 10 10 0), a very small current is drawn from the cell. Since the
output impedance is low (ca. 10 0), a voltage-measuring device such as a
meter or digital voltmeter can be driven with negligible voltage drop
across the output impedance of the voltage follower. Direct-reading
meters typically are limited to an accuracy on the order of 0.1% of full
scale (±1 mV for a full-scale reading of 1 V), but the accuracy can be
improved considerably by using an electrometer in combination with a
potentiometer circuit.
Reference Electrodes
A practical reference electrode should be easily and reproducibly
prepared and maintained, relatively inexpensive, stable over time, and
usable under a wide variety of conditions. Two electrodes-the calomel
and silver-silver chloride electrodes-are particularly common, meeting
these requirements quite well.
The calomeP electrode is shown in Figure 1.14 and represented in
shorthand notation as follows:
filling hole
KCI solution
silver and there is no liquid mercury or Hg2Cl2 paste to deal with. The
Ag/AgCI electrode is easily miniaturized and is thus convenient for
many biological applications.
Since K+ and CI- have nearly the same ionic conductivities, liquid
junction potentials are minimized by the use of KCI as a salt bridge
electrolyte (see Sect. 3.4). Both the calomel and Ag/AgCI electrodes are
normally filled with KCI solution, but there are occasions when other
electrolytes are used. This matter is discussed further in Sect. 4.2.
Indicator Electrodes
If one of the electrodes of an electrochemical cell is a reference
electrode, the other is called a working or indicator electrode. The latter
designation implies that this electrode responds to (indicates) some
specific electrode half-reaction. At this point, it is appropriate to ask
what determines the potential of an electrode. Consider a solution
containing FeS04 and H2S04 in contact with an iron wire electrode. If
we make a cell by adding a reference electrode, there are at least three
electrode processes which might occur at the iron wire:
Fe 2+ + 2 e- --7 Fe(s) EO = -0.44 V
S042- + 4 H+ + 2 e- --7 S02(aq) + 2 H20 EO = 0.16 V
1
H+ + e- --7 2H2(g) EO = 0.00 V
What then is the actual potential of the iron wire electrode? We first note
that in order to have a finite reversible potential at an electrode, all the
participating species must be present in finite concentration so that
appreciable current can be drawn in either direction. Thus, as we have
defined the system with no sulfur dioxide or gaseous hydrogen present,
only the first electrode reaction qualifies and we might guess that the
potential should be determined by the Fe 2 +lFe couple. This is the right
answer but for the wrong reason. Suppose that we bubbled some
hydrogen gas over the iron wire electrode or added a little sodium
sulfite-what then? There is another way to look at this system:
according to the half-cell potentials, the iron wire should be oxidized
spontaneously by either H+ (t1Go = -85 kJ mol-I) or by S042- (t1Go = -116 kJ
mol-I). That neither reaction occurs to any appreciable extent is because
the reactions are very slow. Just as these homogeneous reactions are
very slow, the electrochemical reduction of S042- or H+ at an iron
electrode is slow. If one electrode reaction is much faster than other
possible reactions, only the fast reaction will contribute to the potential-
the slower electrode processes will appear irreversible under the
conditions where we can measure the potential due to the faster process.
Thus in the example, we expect the potential of the iron wire electrode
will be determined by the Fe2+lFe couple with
34 Electrode Potentials
or, at 25°C,
EN = -0.440 - 0.0296 pFe
In other words, the iron wire electrode acts in this case as an indicator
ofthe Fe2+ activity.
Although it is difficult to predict the rates of electrode processes
without additional information, there are a few useful generalizations.
First, electrode processes which involve gases are usually very slow
unless a surface is present which catalyzes the reaction. Thus a surface
consisting of finely divided platinum (platinum black) catalyzes the
reduction of H+ to H2, but this process is slow on most other surfaces.
Second, as a general rule, simple electron transfers which do not involve
chemical bond breaking, e.g., Fe 3 + + e- ~ Fe 2+, are usually fast
compared with reactions which involve substantial reorganization of
molecular structure, e.g., the reduction of sulfate to sulfite.
What then should we expect if the components of two reversible
couples are present in the cell? Consider, for example, a solution
containing Fe3+, Fe 2+, 13-, and 1-. Both half-cell reactions,
Fe3 + + e- ~ Fe2+ EO = 0.771 V
EO = 0.536 V
are reversible at a Pt indicator electrode. If both couples are reversible,
the electrode potential must be given by both Nernst equations
E = EOF 3+1F 2+ -l1:I.ln aFe 2+
e e F aFe3+
1 See books by Koryta (All), Vesely, Weiss, and Stulik (DIl), and by Koryta and
Stulik (D 14) for further details.
2 The first membrane potential was discovered by nature eons ago when animals
first developed nervous systems. A nerve cell wall can be activated to pass Na+ ions
and so develop a membrane potential which triggers a response in an adjacent cell.
Chemists were slower in appreciating the potential of such a device, but we are
catching up fast.
§1.6 Ion-Selective Electrodes 37
(a) (b)
lead wire
internal
solution
Figure 1.15 (a) pH-sensitive glass electrode; (b) schematic view of the glass
membrane.
where aH and aNa are the activities of H+(aq) and Na+(aq) and kH Na is
called the potentiometric selectivity coefficient. The selectivity coefficient
depends on the equilibrium constant for the reaction of eq (1.27) and on
the relative mobilities of the H+ and Na+ ions in the hydrated glass.
These properties depend on the composition and structure of the glass
and can be controlled to some extent. Thus some glass electrodes now
available make use of glasses where lithium and barium replace
sodium and calcium, giving a membrane which is much less sensitive
to sodium ions, allowing measurements up to pH 14. Corrections for
Na+ ions may still be required, however, above pH 12.
38 Electrode Potentials
1 For a good review, see Murray (5). The field is reviewed every two years in
Analytical Chemistry, e.g., by Janata in 1992 (6).
40 Electrode Potentials
Direct Methods
The measurement of pH using a glass electrode is by far the most
common of all electroanalytical techniques (9,D7,DI0) and is also
characteristic of direct determinations using cell potentials. A typical
electrode arrangement for pH measurements is as follows:
Ag(s)IAgCI(s)IKCI( 1 M)IIH +(aq)lglassIHCl(1 M)IAgCl(s)IAg(s)
i.e., a solution in which a Ag/AgCl reference electrode and a glass
electrode are immersed. The half-cell reactions,
Ag(s) + Cl-(1 M) ~ AgCI(s) + e-
AgCI(s) + e- ~ Ag(s) + CI-(1 M)
lead to no net change so that the overall cell reaction is the nominal
transfer of H+ from the test solution to the 1 M solution inside the glass
membrane
H+(aq) ~ H+(aq, 1 M)
The free energy change for this process is
I'1G=RTln-1-
aH+
so that the cell potential is
E=RI..ln aw
F
or, at 25°C,
EN = -0.0592 pH
Thus the cell potential is directly proportional to the pH of the test
solution. In practice, this may not be exactly true because of a liquid
junction potential at the salt bridge linking the reference electrode to the
test solution and a small potential intrinsic to the glass electrode. Thus
buffers of known pH are used to calibrate the pH meter so that the pH is
determined relative to a standard rather than absolutely.
Most of the commercial ion-selective electrodes can be used directly
to determine pX = -log ax, just as the glass electrode is used to measure
pH. Two general procedures are used commonly.
(1) Calibration curve. The potential of a given cell can be
measured with standard solutions and a potential us. concentration
curve constructed. The concentration of an unknown may then be read
from the calibration curve given the potential of the unknown. When
this method is used, it is important that the unknown solution have the
same ionic strength as that of the standards so that the activity
§ 1. 7 Chemical Analysis by Potentiometry 41
Notice that in the example above, the fluoride ion concentration was
determined to approximately 4% expected accuracy. Such accuracy is
characteristic of direct methods. For concentrations of chemical species
above 10- 4 M, there are usually methods (such as the titration methods
discussed below) which are capable of considerably greater accuracy.
But for very low concentrations, a direct electrochemical measurement
may well be the most accurate method and frequently is the only method
available.
Titration Methods
Both the titrant and the sample form reversible couples at the
platinum electrode with formal potentials:
Fe3 + + e- --7 Fe2+ EO' = 0.68 V
Ce 4 + + e- --7 Ce3 + EO' = 1.44 V
The titration reaction is
Fe2+ + Ce4 + --7 Fe3 + + Ce3 +
We first compute the equilibrium constant for the reaction.
Consider the hypothetical cell (not the one being measured in
the experiment) in which the two half-cells are the Fe 3 +lFe 2 +
couple and the Ce 4 +/Ce 3 + couple; the formal cell potential is
1.44 - 0.68 = 0.76 V. Thus we have !:lGo = -FEo = 73 kJ mol- 1
which corresponds to an equilibrium constant K = exp(-
!:lGoIRT) = 7.0 x 10 12 . The reaction will go to completion if the
rate is fast enough.
Up to the endpoint, the potential of the Pt electrode will be
determined by the Fe 3 +lFe2 + couple. The potential of the cell is
[FeZtJ
E = 0.68 - 0.0592 log [--] - 0.244
Fe3+-
Consider the situation when 10 mL of titrant has been added.
Since we have added (0.01 L)(O.OI mol L-1) = 10- 4 mol ofCe 4 +,
there then must be 1 x 10- 4 mol Fe3 + and 1.5 x 10- 4 mol Fe2 +
must remain. The concentration ratio then is [Fe 2+]![Fe 3 +] =
1/1.5 and the cell potential is E = 0.45 V.
At the equivalence point, [Fe3 +] = [Ce3 +] '" 0.005 M and the
equilibrium constant expression gives [Fe 2+] = [Ce 4 +] = 1.9 x
10- 9 M. The Fe 2 +lFe 3 + concentration ratio is 3.8 x 10- 7 and the
cell potential is E = 0.82 V.
Beyond the endpoint, virtually all the Fe 2 + has been
oxidized; 2.5 x 10- 4 mol of Ce3 + has been produced by the
titration reaction and excess Ce 4 + is being added. The cell
potential is given by
E = EOee4+/c 3+ _ RT In [Ce3+-] - E
e F [Ce4+] see
44 Electrode Potentials
1.25
EIV
1.00
0.75
0.50
[Ce3+]
E = 1.44 - 0.0592 log [--] - 0.244
Ce4+
When 30 mL of titrant has been added, there will be 0.5 x 10- 4
mol of Ce 4 + present in excess. The Ce 3 +/Ce 4 + concentration
ratio is 2.5/0.5 and the cell potential is E = 1.15 V.
Carrying out several such calculations, we can sketch the
titration curve shown in Figure 1.16.
1 A battery is a collection of two or more cells connected in series so that the battery
potential is the sum of the individual cell potentials.
§ 1.8 Batteries and Fuel Cells 45
sources (e.g., cells used in flashlights, children's toys, etc.) or as portable
energy storage systems (e.g., automobile batteries).
Interest in electrochemical cells as power sources has revived
recently (10,11) as society has become more concerned for the
environment and as fossil fuels have become scarcer and more
expensive. Electric-powered automobiles offer the hope of substantially
reduced environmental pollution and direct conversion of fossil fuel
energy into electrical energy via an electrochemical cell could in
principle be achieved with much greater efficiency than is possible with
a heat engine. In this section, we will review briefly the operation of
some electrochemical cells which convert chemical energy into
electrical energy. For further information, see books by Mantell (G2),
Angrist (G4), Bagotzky and Skundin (G5), Pletcher (G6), or Ventatasetty
(Gg).
Electrochemical cells can be divided into two classes: primary cells,
which are intended to convert chemical energy into electrical energy,
and secondary cells, which are intended to store electrical energy as
chemical energy and then resupply electricity on demand. A flashlight
cell is used until all the chemical energy has been converted to electricity
and is then discarded. A fuel cell in a spacecraft converts hydrogen and
oxygen to water, extracting the energy as electricity. We usually do not
attempt to recharge a flashlight cell and the fuel cell would not normally
be run backwards to regenerate H2 and 02 gases. A storage battery, on
the other hand, is intended to store electrical energy. Discharging and
recharging are equally important parts of the operational cycle. The
difference in function leads to differences in design.
We first consider three common primary cells which have a family
resemblance, all having a zinc anode (12).
contact cap
insulation
t;=:l5~r-_air space
carbon rod cathode
manganese dioxide
1+--- paper spacer
gelled electrolyte
zinc anode
Figure 1.17 Construction of the
Zn/Mn02 dry cell. insulating disk
One of the few commercially important cells which does not have
nineteenth century origins is the alkaline zinc/mercury cell, developed
for the U. S. Army during World War II. The cell,
§ 1.8 Batteries and Fuel Cells 47
Zn(Hg)IZn(OH)2(S)IKOH(aq),Zn(OH)42-(aq)IHgO(s)IHg(l)
has an open-circuit potential of 1.35 V and is very widely used to power
transistor radios, watches, hearing aids, etc. Unlike the ZnlMn02 dry
cell, the electrolyte is not consumed by the Zn/HgO cell reaction
Zn + HgO + H20 ~ Hg + Zn(OH)2
and as a result, the cell potential is much more constant during
discharge. Accordingly, mercury cells have often been used to provide a
voltage reference in electronic instrumentation.
The cell is constructed with a mercuric oxide cathode (mixed with a
little graphite to improve conductivity) pressed into the bottom of a
nickel-plated steel case. Upon discharge, this layer becomes largely
liquid mercury. The alkaline electrolyte is contained in a layer of
adsorbent material and the zinc amalgam anode is pressed into the cell
top. In addition to the relatively constant potential, mercury cells have a
long shelf life and a high energy-to-volume ratio. The major
disadvantages are that power output drops precipitously at
temperatures below about 10°C and that disposal presents
environmental problems.
Fuel Cells
ohmic heating of the electrolyte solution, most successful fuel cells use a
thin film of solution or in some cases a thin film of an ion exchanger
between the porous anode and cathode.
Storage Batteries
The cell voltage is about 2 V, depending on the state of charge, and the
theoretical maximum energy density is about 5000 J g-I (1.4 kWh kg-I).
The major problem with the cell is the membrane. Na atom diffusion
into grain boundaries in the p-alumina membrane causes short circuits
and structural failure. The sodium-sulfur cell is probably not practical
for electric vehicles (heating the cell to 300°C before operation is one
problem), but as a power plant load-leveling device it has some promise.
REFERENCES
PROBLEMS
0.1
End Point
1.272 mL
Fi~ure 1.18 Titration of 0.0
Hg + with NaI with potential
measured by an iodide-
sensitive electrode. Reprint-
ed with permission from R. -0.1 =--L-.---'-_'---L-.---L.---"'---'---L.____'---'
F. Overman, Anal. Chern. 1.0 1.1 1.2 1.3 1.4 1.5
1971,43, 616, copyright 1971
American Chemical Society. Vol. 0.0100 M NaIlmL
1.14 From data given in Table AA, compute the standard free energy of
disproportionation for the following reactions:
2 Cu+ ~ Cu2+ + Cu
3 Fe2+ ~ 2 Fe3 + + Fe
5 Mn042- + 8H+ ~ 4 Mn04- + Mn2+ + 4 H20
Can you generalize from these results to obtain a criterion for the
stability of a species toward disproportionation based on half-cell
potentials?
1.15 The potential of a glass electrode was 0.0595 V us. s.c.e. when the
electrodes were immersed in a pH 7.00 buffer solution and 0.2598 V
when immersed in a solution of unknown pH.
(a) Calculate the pH of the unknown solution.
(b) If a possible junction potential introduces an uncertainty of
±0.0005 V in comparing the potentials of the known and unknown
solutions, what is the uncertainty in the pH of the unknown?
59
60 The Electrified Interface
structure is called the electric double layer. For reviews of double layer
theory, see Grahame (1), Parsons (2), Mohilner (3), or Reeves (4).
Helmholtz considered the problem of a charged surface in contact
with an electrolyte solution in 1871 (5). He assumed that a layer of
counter ions would be immobilized on the surface by electrostatic
attraction such that the surface charge is exactly neutralized. In the
Helmholtz model, the electric potential falls from its surface value, <Po, to
zero in the bulk solution over the thickness of the layer of counter ions.
This behavior is shown schematically in Figure 2.1.
Later, Gouy (6) and Chapman (7) pointed out that ions are subject to
random thermal motion and thus would not be immobilized on the
surface. They suggested that the ions which neutralize the surface
charge are spread out into solution, forming what is called a diffuse
double layer. According to the Gouy-Chapman model, the potential falls
more slowly to the bulk solution value, as shown in Figure 2.1.
In 1924, Stern (8) observed that neither the Helmholtz model nor the
Gouy-Chapman model adequately accounts for the properties of the
double layer and suggested that the truth lies in a combination of the two
models. Thus some ions are indeed immobilized on the surface (the
Helmholtz layer), but usually not enough to exactly neutralize the
charge; the remainder of the charge is neutralized by a diffuse layer (the
Gouy layer) extending out into the solution. This model is also pictured
schematically in Figure 2.1.
In the following sections, we shall work through some of the details
of the Gouy-Chapman theory, as modified by Stern, and discuss some of
the implications.
§2.1 The Electric Double Layer 61
Gouy·Chapman Theory
The immediate goal of a theoretical description of the electrified
interface is the derivation of an equation giving the electric potential <l>
as a function of the distance from the surface.
The electric potential at any point in the solution is related to the net
charge density at that point by a fundamental relationship derived from
electrostatics and known as Poisson's equation:
V2<l> = _ ~ (2.1)
10100
10( 0)
V 2 = - - r2- + 1 d(.
sm ~ - 0) + 1 -i
r2 or or r2 sin ~ o~ o~ r2 sin2~ 0$ 2
The space charge density p is related to the concentrations and charges
of the ions in solution and can be expressed as
p= L ZiFCi (2.2)
62 The Electrified Interface
1 In this and subsequent chapters we will mostly use concentration units of mol m- 3
(mM). The major exceptions to this rule are those cases where the concentration is
an approximation to an activity referred to the 1 M standard state.
2 The electronic charge e is taken as a positive constant, e = 1.602 x 10- 19 C; zi then
includes both the magnitude (in units of e) and the sign of the charge on ion i.
§2.1 The Electric Double Layer 63
V =_ L
2<J>
££0
L Zi Ci °[1- ZiF<J> + 1 (ZiF<J»2 + ...J
i RT 2 RT
The first term in the sum vanishes since the bulk solution is electrically
neutral. Retaining only the second term, we have
V 2<J> = F 2<J>
EEoRT i
L z·2Ci o
I
(2.6)
XA = ~ eeoRT (2.8)
F2[
a-x
<l>(x) = <l>a exp-- x>a (2. lOb)
XA
where <Da is the potential at the outer surface of the Helmholtz layer (r =
a). Notice that this result differs from eq (2.10b) by the factor (air).
If we restrict our attention to a symmetrical electrolyte such as
NaCI or MgS04, eq (2.5) can be integrated without approximations to
obtain (see Problems)
[exp(zF<D(x)lZRT} -1][exp(zF<D a /2RT} + IJ = exP[( a - x )',XA ] (2 12)
IJ .
;---.:..--------";-';c-----=----'C----~
[exp(zF<D(x)IZRT} + l][exp{zF<DaIZRT} -
Equation (2.12) reduces to eq (2.10b) in the limit where zF<Da «RT.
Now that we have solutions to the Poisson-Boltzmann equation, let
us examine some of the consequences. First, we look at the shape of the
solutions represented by eq (2.12). In Figure 2.2 are plotted <D(x)/<Da us. (x
- a)lxA for <Da = 0.01 and 0.1 V, assuming that z = 1 and T = 298 K. The
curve for <Da = 0.01 V is indistinguishable from the simple exponential
dependence predicted by the approximate solution, eq (2.10b). Indeed,
even for <Da = 0.1 V, deviations from the approximate solution are not as
great as we might have expected.
1.00
<D/<Da
0.75
0.50
0.25
0.00
0.0 0.5 1.0 1.5 2.0 2.5 3.0 0.0 0.5 1.0 1.5 2.0 2.5 3.0
(x - a)lxA (x - a)lxA
xAlnm
C/mM 1:1 1:2 1:3 2:2 2:3
0.1 30.4 24.8 21.5 15.2 7.8
1.0 9.6 7.8 6.8 4.8 2.5
10.0 3.0 2.5 2.15 1.52 0.78
100.0 0.96 0.78 0.68 0.48 0.25
a=- L~ p dx
1 In the context of Debye-Huckel theory (Section 2.6), XA is called the Debye length.
§2.1 The Electric Double Layer 67
0=E£o1°O ~dx
dx 2
=EEo~t
dx
a
The total potential energy is the sum of the attractive and repulsive
contributions given by eqs (2.14) and (2.15), respectively.
To get a feel for the nature of these equations, the total potential
energy is plotted as a function of distance between the particles in Figure
2.4, where we have assumed particles with radii of 50 nm, A = 5 x 10-20 J,
and E = 78.5. The energy is given in units of kT (T = 298 K). Since kT is a
measure of the amount of thermal energy available, such a scale
permits a rough judgment on the significance of a barrier to
coagulation.
Figure 2.4a suggests that particles with a surface charge density of
0.01 e nm- 2 should coagulate quickly when the electrolyte concentration
is sufficiently high that the ion atmosphere is only 5 nm thick. When the
thickness is 10 nm, there is a small barrier to coagulation (about 3.5 kT),
and coagulation would be expected to occur fairly rapidly (e- 3 .5 = 0.03).
When the ion atmosphere is 15 nm thick, the barrier is about 36 kT and
coagulation should be much slower (e-35 = 2 x 10- 16 ) and when the ion
atmosphere is 20 nm thick, the barrier is even greater. Figure 2.4b
shows that coagulation is very sensitive to surface charge density with
no barrier at all for ()" = 0, a small barrier for 0.01 e nm- 2 , and barriers of
33 kT and 96 kT for ()" = 0.015 and 0.02 e nm- 2 , respectively.
According to eqs (2.8), (2.14), and (2.15), the potential energy also
depends on the temperature and the dielectric constant of the medium
as well as the size of the particle. We expect that salting-out should
occur at lower ionic strength when either the dielectric constant or the
temperature is reduced. Since the surface charge of a protein, for
example, depends on the number of free carboxyl or amino groups
present and thus on the pH of the medium, protein solubility is strongly
pH dependent, going through a minimum at the isoelectric pH. Since
protein solubilities are such intricate functions of ionic strength,
temperature, dielectric constant, and pH, it is not surprising that protein
mixtures can frequently be separated by selective precipitation
techniques.
§2.2 Some Properties of Colloids 71
100
"
,'\ (a)
, ,, 5
'..,, (b)
o
'. ..•.••.••• 10 ,, 0.01
,
", _................. 15 " 0.015
50
, ,
'"'''''''' -------!2() ". . 0.02
'........... ,..... "',
........ " ...., ' ......
........ ! ......... . .
---...............
........ .. .................. ..
o ...............::::-..... _..... _..... _.- !
~~.:.:.:;:.:;.:==.........................'1!": t'· ................
---.. ..............."':::=:~,-
......... . . "-T_
:l
...
-50
o 10 20 30 o 10 20 40
Rlnm Rlnm
Figure 2.4 Potential energy of interaction (in units of kT) as a function
of separation for two spherical particles of radius 50 nm; (a) cr = 0.01 e
nm-2~z4 = 5,10,15, and 20 nm; (b) XA = 10 nm, cr = 0, 0.01, 0.015, and 0.02
e nm .
Surface pH
We have seen that ions are either concentrated in or excluded from
the double layer, depending on the relative charges. This conclusion
applies to hydrogen and hydroxide ions as well, so that the pH at the
surface of a charged particle will be different from that of the bulk
solution. According to eq (2.3), the concentration of H + ions at the
surface is given by
The Phenomena
(a)
pressure
head
~
capillary
current or
voltage meter
~------~~~------~
Figure 2.5 (a) Apparatus for measuring the streaming current or
streaming potential. (b) Ions moving in the capillary; with adsorbed
anions, positive space charge is swept along by solution flow.
center of the capillary along as well. The effect of liquid flow under the
action of an electric potential difference is called electroosmosis.
If we carry out an electroosmosis experiment and allow flow to
continue such that a pressure difference is built up, the flow will
eventually slow and come to a halt as the pressure opposes and finally
cancels the effect of the field. The relation of this steady-state
phenomenon, called the electroosmotic pressure, to electroosmosis is
analogous to the relation of streaming potential to streaming current.
Electroosmosis and electroosmotic pressure may be expressed by the
phenomenological equation
} = a3MJ + a4~<I> (2.19)
where} is the flux ofliquid through the capillary (in m 3 s- I ). Thus, in the
absence of a pressure difference, the flux is proportional to the electric
potential difference ~<I>. When the flux goes to zero, the pressure
difference is proportional but opposed to the electric potential difference.
If we assume that a, the radius of the capillary, is much larger than the
diffuse layer thickness XA, then we need not rederive an equation for the
potential but can use the result obtained for a planar surface,
<l>(r) = ~ exp[(r - a)/xA] (2.21)
where the coordinate r is measured from the center of the capillary. We
have assumed that a / XA is large enough that the potential in the center
of the capillary is nearly zero.
76 The Electrified Interface
i = - 1teEot;t:J>
211L
[2a 2 - 4xl (-'L-1) + (a 2_ 4xl) exp (_-'L)~
XA XA ~
But since a »
XA, the last two terms are negligible and
. 1ta 2eEot;
<Xl = _l_ = - -----=- (2.25)
t:J> 11L
The coefficient <X2 is the current per unit potential difference when there
is no pressure difference across the capillary. Under these conditions
the current and potential difference are simply related by Ohm's law, i =
t,.cl>/R. The resistance R of a cylinder of radius a, length L, and
resistivity p is R = pL/1ta 2 , so that the current is
i = 1ta2 t,.cl>/pL
and the coefficient is
. 1ta 2
<X2=_l_=_- (2.26)
t,.cl> PL
Similar calculations (see Problems) give
§2.3 Electrokinetic Phenomena 77
1ta 4
<Xa=-- (2.27)
811 L
and
(2.28)
This last result turns out to be more general than we might have
expected from our rather simple calculation. Indeed, the electrokinetic
effects can be treated by the methods of non equilibrium thermodynamics
with the completely general result that <Xl = <X4 independent of the size or
shape of the holes through which the solution flows. 1
Let us restate the four electrokinetic effects in terms of the
phenomenological equations
i = <X1.1P + <X2.1<1> (2.18)
j = <X3.1P + <X4.1<1> (2.19)
The streaming potential is the potential difference per unit pressure
difference at zero current, and thus may be expressed quantitatively as
1 These results are the more surprising since it would appear that the four quantities
should have different units. However, the apparent units of the electroosmotic flow,
(m 3s- 1)/(C s-l), reduce to m 3C-1 and the apparent units of the streaming potential, Y
Pa-\ also reduce to m 3C·I. Simila;ly, the apparent units of the streamin! current,
(C s- )!(m 3s- I ), and the electroosmobc pressure, Pa y-l, both reduce to C m- .
§2.3 Electrokinetic Phenomena 79
0.0
KOH
Figure 2.6 The zeta potential at SlV
a glass-aqueous solution inter- KN03
face as a function of electrolyte
concentration for (a) KOH, (b)
KN03, and (c) HN03 deter- .'
mined by measurement of the
electroosmotic flow through a
-0.1 ........... .,.,.,
....::::::::>..//
capillary tube. The lines
represent smooth curves drawn
through data points rather than
fits of data to a theoretical
equation. Reprinted with per- ._ .... -.- ." .....
..'
mission from A. J. Rutgers and -0.2
M. de Smet, Trans. Faraday -Q -4 -2
Soc. 1945, 41, 758, copyright 1945
Royal Society of Chemistry. log C
1.0
~/V ........ .............................
<1>/<1>0
/.~···········:rn03
0.1
0.5
0.0 0.0 l
!.: -----.....
##
Ca(NOs)2
.l /' .................... Al(NOn)
II
...
~S
,
-0.5
..
-0.1 ":/
-1.0
0 1 2 3 -7 -6 -5 -4
log C
Figure 2.7 Reversal of zeta potential Figure 2.8 Zeta potential at a glass-
by adsorption of ions. The upper aqueous solution interface v s.
curve shows the normal behavior concentration for (a) KNOs, (b)
with <l> falling exponentially through Ca(NOS)2, and (c) Al(NOS)S,
the ion atmosphere. The lower curve determined by electroosmotic flow
shows the behavior expected when through a glass capillary. Reprinted
oppositely charged ions are adsorbed with per-mission from A. J. Rutgers
on the surface forming a "triple and M. de Smet, Trans. Faraday
layer" and reversing the sign of the Soc. 1945,41, 758, copyright 1945
zeta potential. Royal Society of Chemistry.
§2.4 Electrophoresis and Related Phenomena 81
1.0 r-----r---r----,r-----r----,
.$
.~ 0.9
'-'
en
:>
<l)
Figure 2.9 Relative viscos- >
ity of starch solutions as a ~
function of salt concentra- ~ 0.8
<l)
tion for 1: 1, 1:2, and 1:3 ~
electrolytes. Reprinted with
per-mission from H. G.
Bungenberg de Jong, Rec.
Trav. Chim. 1924,43 189, 0.7 1--_-'-_ _.L.-_-..L._ _.L.-_---J
copyright 1924 Royal Neth- o 1 2 3 4 5
erlands Chemical Society. Electrolyte ConcentrationlmM
Sedimentation Potential
When colloidal particles move through an electrolyte solution under
the influence of gravity (as, for example, in an ultracentrifuge), the
particles may leave some of their ion atmosphere behind, especially if
the diffuse layer is very thick (low ionic strength). Thus charge
separation may accompany sedimentation, leading to the development of
an electric potential gradient along the length of the sedimentation tube.
The phenomenon is thus analogous to the streaming potential discussed
above. A sedimentation potential is generally an unwanted complication
in an ultracentrifugation experiment. To reduce the magnitude of the
effect, ultracentrifugation is usually carried out in high ionic strength
media so that the diffuse layer is relatively compact and charge
separation is minimized.
Electrophoresis
cr
= eeo~
a (1 + .iL)
XA
For 1 mM ionic strength, XA = 9.6 nm, so that a/xA = 0.31 and f '"
2/3; eq (2.38) gives u = 9.3 x 10-9 m 2 s- 1V-l. This mobility is typical
for charged macromolecules (see §3.3).
Practical Electrophoresis
Although eq (2.38) does not explicitly include the thickness of the
diffuse layer, XA, the zeta potential depends on XA (or C) as shown by eq
(2.33) and the geometric factor f depends on alxA. Thus the
electrophoretic mobility is expected to decrease with increasing ionic
strength of the electrolyte solution. The mobility is also proportional, of
course, to the surface charge density on the moving colloidal particle
and, for a protein, is expected to be pH dependent, even at constant ionic
strength; indeed, the mobility is expected to go to zero at the isoelectric
pH. Because of the dependence of the electrophoretic mobility on ionic
strength and pH, a mechanism for control of protein mobilities is
available which is used in practical separation techniques.
In our discussion thus far, we have tacitly assumed that
electrophoresis experiments are done on a suspension of colloidal
particles in liquid solution, and, indeed, the early experiments were.
Thus, in classic experiments done in the 1930's, Tiselius found that
electrophoretic mobilities of protein molecules in solution could be
measured by a moving boundary method (16). Tiselius' apparatus was
essentially a U-tube, shown in very simplified form in Figure 2.10. The
U-tube was constructed so that the solution of the colloid could be placed
in the bottom of the tube and solvent (containing electrolyte) layered on
top on both sides. Electrodes were then mounted in the two ends of the
U-tube and a high voltage applied. When the protein molecules
migrated in the solution, the boundary between the two regions of the
solution remained intact and was seen to move with a velocity which
was easily related to the electrophoretic mobility of the protein. To
minimize convective mixing of the solutions and blurring of the
boundary, it was necessary to very carefully thermostat the U-tube, and,
since the rate of change of the density of water with temperature is
smallest near the density maximum (4°C), best results were obtained
when the system was thermostated at about 2-3°C. In practice, the
§2.4 Electrophoresis and Related Phenomena 85
Unlike a protein molecule where the surface potential and the ion
atmosphere are a response to the solution pH and ionic strength, the
potential of an electrode, and thus the nature of the double layer, can be
controlled by external circuitry. This control permits quantitative
experimental studies of double-layer effects. In this section, we will
discuss two effects which have important consequences In
electrochemical experiments to be described in subsequent chapters.
86 The Electrified Interface
Electrocapillarity
The interfacial tension (surface tension) at an electrode-solution
interface turns out to be a function of the electrode-solution potential
difference. When the electrode is a rigid solid, the effects of interfacial
tension are difficult to discern, but when the electrode is nonrigid, a
mercury drop for example, the effects are more obvious. In a dropping
mercury electrode, mercury is allowed to flow through a fine-bore
capillary immersed in the electrolyte solution. A mercury drop forms at
the end of the capillary and grows in size until it becomes too heavy to be
supported by the interfacial tension. l Thus changes in interfacial
tension are reflected in changes in the maximum mercury drop size, or
if the flow through the capillary is constant, in the lifetime of a mercury
drop.
Consider a mercury drop attached to a mercury-filled capillary and
suspended in an electrolyte solution. If a second electrode is provided,
we can change the mercury-solution potential difference and thus
induce a positive or negative charge on the mercury drop by passing a
small current through the external circuit. Similarly, we can change
the drop size by allowing mercury to flow through the capillary.
Expansion of the drop involves an increase in surface area and thus
work is done against the interfacial tension, dw = ydA. Similarly,
electrical work is done when the drop is charged, dw = <l>dQ. Thus at
constant temperature, pressure, and composition, the change in Gibbs
free energy is
dG = y dA + <I> dQ
This expression can be integrated
G=yA+<I>Q
and redifferentiated
dG = ydA +A dy+ <I> dQ + Q d<l>
Subtracting the two expressions for dG leaves a form of the Gibbs-
Duhem equation,
o=A dy+ Q d<l>
Introducing the surface charge density, cr = QIA, we have
(:;tR =-0
Some typical data from the work of Grahame (1) are shown in Figure
2.11 for several aqueous electrolyte solutions of equal ionic strength.
The curves of Figure 2.11 show the qualitative behavior predicted by
eq (2.41). However, the simple theory predicts identical curves for all
electrolytes of equal ionic strength. The curves of Figure 2.11 coincide at
negative potentials but show rather different behavior near zero or at
positive potentials. When the mercury drop is positively charged, anions
are expected to be concentrated near the electrode surface, and the
different behavior suggests specific interactions of the various anions
with the surface. The "soft" bases like I- and SCN-- apparently interact
more strongly than the "hard" bases like OH-, CI-, and N03-. Since
mercury can be thought of as a soft acid, the interaction most likely
involves some degree of chemical bond formation and we say that the
anions are chemically adsorbed on the mercury surface. When the
potential is made sufficiently negative, anions are desorbed and all
0.45
yIN m- 1
0.40
,.,.,........................... .
,. /
0.35 I
./ ---
,.i ................... .
,
; NaCI
0.30 ! NaBr
------- KSCN
_..... _.... KI
0.25
0.50 0.00 -0.50 -1.00
EIV
Figure 2.11 Interfacial tension of mercury in contact with various
aqueous electrolyte solutions. The potential scale is adjusted so that zero
potential occurs at the electrocapillary maximum for a capillary-inactive
electrolyte. Reprinted with permission from D. C. Grahame, Chern. Rev.
1947,41,441, copyright 1947 American Chemical Society.
curves merge together, suggesting that cations (at least Na+, K+, and
Ca2 +) do not specifically interact with the mercury surface, even when it
is highly negatively charged.
It is commonly observed that the drop time of a dropping mercury
electrode depends on potential, usually varying by a factor of 2 or more
over the experimentally accessible range. More complex behavior is
sometimes observed when surfactants are adsorbed on the electrode
surface. Since adsorption is often potential dependent, very rapid
changes in interfacial tension (and thus drop time) with potential are
sometimes observed.
Double-Layer Capacitance
differential capacitance)
C=dQ (2.42)
d<l>
If we take C to be the capacitance per unit area (the capacity), then we
can replace Q by the surface charge density cr. Consider first the
capacity of the Gouy layer, Co. Differentiating eq (2.13), we obtain
C - ££0 (2.43)
G - XA
The curve for 0.001 M NaF solution in Figure 2.12 goes through a
minimum of about 0.06 F m- 2 , in excellent agreement with the value
computed in Example 2.6 considering the simplicity of the theory.
However, the minimum capacity of the 0.01 M NaF solution is about 0.1 F
m- 2 and for 0.1 M NaF the observed capacity at the electrocapillary
maximum is only 0.2 F m- 2 , rather less than the computed values, 0.23
and 0.72 F m- 2 . Furthermore, the capacity minimum at <I> = 0 nearly
disappears for concentrated solutions. The reason for these apparent
discrepancies is not hard to find: we have neglected the effect of the
immobilized ions and solvent dipoles on the surface which comprise the
Helmholtz layer. These also contribute to the capacity. From the point of
view of the external circuit, these two contributions-C H for the
Helmholtz layer and C G for the Gouy layer-behave like capacities in
series. Thus the total observed capacity is given by
1/C = 1/CH + 1/CG
or
C = CGCH (2.44)
CG+CH
For dilute solutions, the capacity of the diffuse layer C G is small
compared with the more or less constant Helmholtz layer contribution,
so that C'" CG. When the diffuse layer becomes more compact and its
capacity becomes large compared with that of the Helmholtz layer, then
the latter contribution should dominate. Indeed, it is found that at zero
potential the double-layer capacity initially increases with concentration
and then levels off. In the case of NaF solutions in contact with
mercury, the Helmholtz contribution to the double-layer capacity at <I> = 0
appears to be approximately 0.29 F m- 2 .
In 1923 Debye and Huckel (21) found a way to calculate ionic activity
coefficients for dilute electrolyte solutions. The basic idea of the theory is
that, just as in the case of the charged surfaces we have been
considering, individual ions possess an atmosphere of ions of opposite
charge. By calculating the electrostatic free energy of interaction of an
ion with its atmosphere, an estimate of the activity coefficient can be
obtained. Debye-Huckel theory starts with the Gouy-Chapman
description of the ion atmosphere derived from the linearized Poisson-
Boltzmann equation.
§2.6 Debye-Huckel Theory 91
Ili(elec) ::: RT In ai - RT In Ci
and, since the ratio of the activity to the concentration is the activity
coefficient,
(2.46)
we have
Ili(elec) ::: RT In Yi (2.47)
Thus if we can calculate the electrostatic contribution to the chemical
potential of an ion, we win have the activity coefficient.
Consider an ion of charge zie surrounded by its ion atmosphere and
suppose that the electric potential at the surface of the ion is <D. We now
1 The standard state in most treatments of activity coefficients is the 1 molal ideal
solution. Since most practical work employs the molar concentration scale, we shall
use the 1 M standard state here. The penalty for this choice is that most data
tabulations must be converted to molar concentration units.
92 The Electrified Interface
f'"
case is the electrical contribution to the chemical potential:
where <l> is the contribution to the potential from the ion atmosphere.
Notice that this chemical potential has units of energy per ion; we will
convert to a molar basis later.
For a spherical particle of radius a, the surface potential <l>a is
related to the surface charge density () by (see Problems)
<l> - aa
a - ££0(1 + a/xAl
where £ is the dielectric constant of the medium and XA is the thickness
of the ion atmosphere (called the Debye length in this context). If Q is the
total charge on the ion, the surface charge density is
a =--.SL
4n:a 2
We then have
<l>a =~ 1 (2.49)
4n:££0 a (1 + a/xAl
For the thermodynamic calculation, we need the potential at the ion site
due to the ion atmosphere. Equation (2.49) contains contributions from
both the ion atmosphere and the central or test ion. However, the latter
contribution is easily computed:
<l>central(a) = _Q_-
4n:££(j1
Subtracting this from eq (2.49), we have the ion atmosphere contribution
to the potential at the central ion:
<l>atm(a) =_~_1_
4n:££o a + XA
Substituting this expression in eq (2.48) and integrating, we have
Ili(elec) = _ _ 1 _ +1
_
4n:££0 a XA
r
Jo
e
Q dQ
Z·~2 1
Ili(elec) :;:; - - '- ---
8n:££0 a + XA
§2.6 Debye-Huckel Theory 93
where V = VI + V2. Comparing eqs (2.52) and (2.53), we see that the mean
ionic activity is related to the single ion activities by
(2.54)
(2.55c)
But, since CI =VIC and C2 =V2C, C± is not the ordinary molar electrolyte
concentration C, but is
(2.56)
1 Al z lZ21 fl (2.57)
og y± = 1 + Ba fi
In eq (2.57), A and B are constants, I is the ionic strength in
concentration units of mol L-l, and we have converted to common logs.
Inserting the values of the constants and assuming a temperature of
25°C and the dielectric constant of water (78.54), we can write
1 0.5092I z 1Z21 fl (2.58)
ogy± =- 1 + 3.29afl
In eq (2.58), the ionic strength and IZ1Z21 are known for any given
electrolyte solution. The parameter a (in nm) entered the theory as the
ionic radius. However, in going from single ion activity coefficients to
mean ionic activity coefficients, it has lost its simple meaning, and
should to be regarded as a parameter adjustable to fit the theoretical
activity coefficients to experiment.
At low ionic strength, the second term in the denominator of eq
(2.58) can be neglected compared with unity and we have the limiting
form of Debye-Huckel theory:
log y± = - 0.50921 ZlZ21 fl (2.59)
Equation (2.59) contains no adjustable parameters and thus is an
absolute prediction of the activity coefficients.
1.0
F(I)N (a)
y±
0.230
0.9
0.225
1oC..L..................................&....L.............................................I 0.8
0.0 0.1 0.2 0.0 0.1 0.2
[112 [1/2
Figure 2.13 (a) Corrected cell potential F(I) for the cell of Example 2.8 as
a function of the square root of ionic strength. The solid line corresponds
to a least-squares fit of the data to a parabola. (b) Mean activity
coefficients, y±, computed from the cell potentials. The solid line
corresponds to activity coefficients computed using eq (2.59). Data from
Harned and Ehlers (22).
Ksp = (C±y±)V
where v = m + n. If the solubility is measured at several ionic strengths,
adjusted with an electrolyte which has no ion in common with the
slightly soluble salt, Ksp can be determined by extrapolation to zero ionic
strength and, knowing K sp , the mean ionic activity coefficients y± can be
computed for each solution.
Some data from such a study (23) are shown in Figure 2.14 for the
1:1 electrolyte,
[Co(NH3M C204)] [CO(NH3)2(N02)2( C20 4)]
the 1:2 electrolyte,
-0.00 2.0
log y± y+ KCI
-0.03 1.5 CaCl2
LaCIs
-0.05 HCI
1.0
-0.08
-0.10
0.0
0.00 0.04 0.08 0.12 0.0 0.5 1.0 1.5 2.0
[112 [112
-0.00
logy±
-0.05
m= C (2.61)
d-CMw1000
where MB is the solute molecular weight and d is the solution density.
If an electrolyte of molality m ionizes to VI cations and V2 anions,
then the mole fractions Xi are
x. - vim (2.62a)
• - vm + 1000lMA
where v = VI + V2, and MA is the molecular weight of the solvent. In
terms of molar concentrations, the mole fractions are given by
100 The Electrified Interface
x. - ViC (2.62b)
, - vC + 1000 dlMA - CMBIMA
If the molal concentration scale is used, then the mean ionic
activity is defined by eq (2.51) with
a± = m±y±m
The mean ionic activity coefficient y±m is related to the activity coefficient
for the molar concentration scale by
y±m = Cd y± (2.63)
m 0
where do is the density of the solvent. If the mole fraction scale is used,
the corresponding activity coefficients are
X d + 0.001 C(VMA - MB
(2.64)
~ = do ~
Solvent Activity
The standard state of the solvent in an electrolyte solution is usually
taken to be the pure solvent at 1 bar total pressure and 25°C. In simple
treatments of solution equilibrium, it is usually assumed that the solvent
activity is I, but a more exact value is sometimes required. If the
solution were ideal, the solvent activity would be equal to the mole
fraction, but in solutions where the mole fraction of solvent is
significantly less than unity, nonideality is usually sufficiently
important that the mole fraction is a poor estimate of the activity.
The most straightforward way to determine solvent activity is to
measure the partial pressure of solvent vapor. We define the activity of
the solvent in solution in terms of the chemical potential
IlA = IlA ° + RT In aA (2.65)
where IlA ° is the chemical potential of the pure solvent. When the pure
solvent is in equilibrium with its vapor,
IlAo = IlA(vapor) = IlAO(vapor) + RT In PAc
where PA ° is the equilibrium partial pressure. 1 When a solution is in
equilibrium with vapor, we have
IlA = IlA O(vapor) + RT In PA
Subtracting these expressions gives
1 In very careful work, account must be taken of nonideality of the vapor phase as
well. We ignore that complication here.
§2.6 Debye-Hiickel Theory 101
IlA -IlA °= RT In PA
PAo
so that
(2.66)
While the vapor pressure method is theoretically straightforward, it
is usually rather inconvenient experimentally. Another approach to
solvent activities is through solute activity coefficients. Consider the
fundamental equation for the Gibbs free energy:
dG = P dV - T dS + L Ili dni
i
Integrating and redifferentiating, we have
G = PV - TS + L J.li ni
i
and
we have
In aA =-f-
A 0
i y+c
CV d In(y±C) (2.67)
X I' = __V-'I'---_
nA -ns +v
where v = VI + V2. Thus we have
Xl _ nA -ns+v _ X 2
Xl' - nA + v - X2'
When the solute concentration is very small, nA is very large (we have
assumed one mole of solute). Thus as the solute concentration
approaches zero, XliX I' ~ 1, aA ~ 1, all the activity coefficients
approach 1, and all the log terms on the right-hand side of eq (2.69) go to
zero; thus the sum of standard state terms on the left-hand side of the
equation must be zero, independent of the solute concentration.
Rewriting eq (2.69) with the new information included, and using eq
(2.55c) to convert the activity coefficient terms to mean ionic activity
coefficients, we have
nA -ns + v X X
ns In aA + v In + v In Y± = v In Y±
nA +v
104 The Electrified Interface
Electrolyte ns alnm
HCI 8.0 0.447
HBr 8.6 0.518
NaCI 3.5 0.397
NaBr 4.2 0.424
KCI 1.9 0.363
MgCl2 13.7 0.502
MgBr2 17.0 0.546
From Robinson and Stokes (B9).
References 105
REFERENCES
PROBLEMS
2.2 Derive eq (2.12). Hint: With u = zFc'PIRT, show that eq (2.5) can be
written
V 2 u = XA- 2 sinh(u)
This expression is then integrated to obtain
duldx = --{21xA) sinh(uI2)
A second integration, followed by rearrangement, yields eq (2.12).
2.4 Use eq (2.11) to derive an expression for the surface charge density
cr at a spherical surface as a function of the surface potential <P a ,
the radius of the sphere a, and the ion atmosphere thickness XA.
2.7 Use the result of Problem 2.3 to show that eq (2.43) is actually the
minimum on a C us. <Pa curve.
2.8 Consider an electrochemical cell with a planar electrode of area 1
cm 2 and a double-layer capacity of Cd = 0.20 F m- 2 in contact with
an electrolyte solution. A reference electrode is placed in the
Problems 107
2.9 Butler and Huston (Anal. Chern., 1970, 42, 1308) used
electrochemical measurements to obtain activity coefficients in
mixed electrolyte solutions. Using solutions containing known
concentrations of NaCI and NaF, they measured potentials of the
following cells:
AgIAgCI(s)IN a+(aq),CI-(aq),F-(aq)IN a+ -selective glass electrode
AgIAgCI(s)INa+(aq),CI-(aq),F-(aq)IF--selective electrode
Na+-electrodeINa+(aq),CI-(aq),F-(aq)IF--selective electrode
What combinations of activity coefficients were obtained from each
of the three cells?
2.11 Harned and Geary (J. Am. Chern. Soc. 1937,59,2032) measured the
potential of the cell
PtIH2(g)IHCl(O.01 M), BaCI2IAgCI(s)IAg(s)
at 25°C as a function of the BaCl2 concentration. The potential of
the cell is given by the Nernst equation, corrected for activity
coefficients:
E = EO _RT In [H+] [CI-] RT In it
F PH 2 F
Potentials, corrected to the standard state pressure of hydrogen,
are as follows:
2.12 The mean ionic activity coefficient of 1.00 molal KOH is 0.735. The
solution density is 1.0435 g cm· 3 and the density of pure water is
0.9971 g cm-3. What is the molar concentration of KOH? What is the
mean ionic activity coefficient on the molar scale?
3.1 CONDUCTIVITY
109
110 Electrolytic Conductance
The unit of resistance is the ohm 0, so that resistivity has units of a-m.
The unit of conductance is the siemens (S), 1 S = 1 0-1, so that the units of
conductivity are S m- 1.
(a) (b)
v
ose
Molar Conductivity
Pt electrodes
(b)
According to eq (3.1),
L/A = KR = (1.2896 S m- I )(89.3 n) = 115.2 m- I
so that for the NaCI solution,
1C =(115.2 m- I )/(56.6 Q) =2.035 S m- I
A = (2.035 S m- I )/(200 mol m- 3 ) = 0.0102 S m 2mol- I
I In the older literature K has units of Q-Icm- l (or the whimsical units, mho cml),
and A, with units of Q- 1cm 2mol- 1 , is called the equivalent conductivity and defined
as IOOOKlzC, where C is the molar concentration and z = n+z+ =n_1 z_1 is the number
of equivalents per mole. The new definition is rather more straightforward, but of
course leads to different numerical values; thus beware when using older
tabulations of data.
§3.1 Conductivity 115
Ionic Conductivities
If, at infinite dilution, the behavior of electrolytes is indeed ideal (i.e.,
we can neglect interionic interactions), and if electrolytes are completely
ionized, then the molar conductivity of an electrolyte should be the sum
of the molar conductivities of the individual ions. 1 If one mole of the salt
produces v+ moles of cations with molar conductivity A+ °and v_ moles of
anions with molar conductivity fLo, then
AO=v+~o+v~o (3.4)
The molar conductivity of an individual ion cannot be measured directly,
but if eq (3.4) is correct, then subtracting the molar conductivities of NaCI
0.05
6 NaCI
0.04
.....
,.:....
0
S
C'I
0.03
S
~ 0.02
1 This idea, known as Kohlrausch's law of independent ionic migration, was first
proposed in 1876.
116 Electrolytic Conductance
and KCI, for example, should give the difference between the molar
conductivities of the sodium and potassium ions.
AO(KCl) - N(NaCl) = (149.8 - 125.4) x 10-4 S m 2mol- 1
N(K+) - N(Na+) = 23.2 x 10-4 S m 2 mol- 1
Comparing KI and NaI, the difference is 23.8 x 10- 4 , and for KBr03 and
NaBr03, the difference is 23.3 x 10-4 . Although there is a little scatter, the
agreement is quite good.
Transference Numbers
For KCI,
t+O = N(K+)IN(KCl)=73.51149.8 =0.491
LO =1-t+0 =0.509
for HCI,
t+O = AO(H+)IN(HCl) = 349.81426.1 = 0.821
LO =1-t+0 =0.179
and for BaCI2,
t+o = N(Ba2+)/ACBaCI2) = 127.7/280.3 = 0.456
LO =1-t+0 =0.554
Ionic Mobilities
It is sometimes useful to think of the transport of an ion in terms of
its mobility. (We encountered the electric mobility in our discussion of
§3.1 Conductivity 119
Ui =-~ (3.9)
IZil F
Thus ionic mobilities at infinite dilution can be obtained from molar
ionic conductivities such as those given in Table A.8.
Frictional Coefficients
When an ion moves through solution, it is subject to a viscous drag
force proportional to the ion's velocity,
Fi =-fiVi
where the proportionality constant fi is called a frictional coefficient.
Under steady-state conditions, the ion moves at constant velocity so that
the viscous drag force exactly cancels the electrical driving force, zieE,
and the net force is zero:
or
(3.10)
Example 3.6 Compute the Stokes' law radii ofNa+ and Ba2 +.
The viscosity of water at 25°C is 0.890 x 10-3 kg m-Is- I ; inserting
the mobilities from Example 3.5, eq (3.13) gives
rCNa+) = 1.84 x 10- 10 m (184 pm)
r(Ba2+) = 2.89 x 10-10 m (289 pm)
§3.1 Conductivity 121.
Faster moving ions, of course, have smaller Stokes' law radii, and
slower ones have larger radii. While Stokes' law is not really valid for
small ions in ordinary solvents,! Stokes' law radii do give a rough
measure of the effective size of ionic species as they move through a
solution. Some Stokes' law radii, calculated as in Example 3.6, are
compared with ionic radii from crystal structure data in Table 3.I.
Several insights can be obtained from examination of these data. In the
first place, it is hard to imagine a hydrodynamic radius being smaller
than the crystal radius of the same ion. Comparison of the radii ofI-, for
example, suggests that the Stokes' law radii are probably systematically
underestimated, perhaps by about a factor of 2. However, within a
related series of ions, e.g., the alkali metal cations, the trends are
interesting. The crystal radii increase, as expected, in the series Li+ <
Na+ < K+, but the Stokes' law radii go in the opposite direction. This
must mean that the small, highly polarizing Li+ ion tightly binds a lot of
solvent molecules which must move with the ion as a unit. The same
effect is seen in comparison of the alkaline earth cations Mg2+, Ca2+, and
Ba2+.
The three halide ions, CI-, Br-, and I- all have about the same Stokes'
law radii despite an increase in the crystal radii; this again is consistent
with larger, less polarizing ions binding fewer solvent molecules or
binding them less tightly so that fewer move as a unit. Comparing the
1 The law was derived in 1845 by Sir George Stokes assuming a rigid sphere moving
through a continuous medium; using Stokes' law for a charged deformable ion
moving through a solvent of discrete dipolar molecules of size comparable to the ion
must introduce some serious errors.
122 Electrolytic Conductance
isoelectronic ions Na+, Mg2+, and A13+ we see an even more dramatic
effect which has the same qualitative explanation. Extending the
comparison to F-, which is isoelectronic with Na+ and has about the
same crystal radius, we see that the solvation sheath is apparently
greater for Na+ than for F-, consistent with the idea that cations are
rather more specifically solvated (formation of coordinate covalent
bonds) than are anions (orientation of water dipoles and/or hydrogen
bonding).
From crystal structure data, OH- and F- would be expected to be
nearly the same size, but the Stokes' law radius of OH- is 47 pm, much
less than the radius of any other anion. A similar discrepancy is found
for H+, which has a Stokes' law radius of 26 pm. Structural evidence
suggests that H+(aq) is better represented by H30+, which might be
expected to have a Stokes' law radius comparable to Li+ or Na+, say about
200 pm. The anomalously high conductivities of H+ and OH- in water
can be understood if we recall that these water-related ions are part of
an extensive hydrogen-bonded network involving solvent water
molecules. A small shift in hydrogen bonds thus can move the charge
from one oxygen center to another and several such shifts can move the
charge much faster than any of the nuclei travel. Thus the conductivity
is much higher than might have been expected from the size of the HaO+
or OH- ions.
1 See Robinson and Stokes (B9) for a review of the various approaches.
§3.1 Conductivity 123
The first term in the Onsager limiting law arises from the so-called
ion atmosphere relaxation effect. When an ion is attracted by an electric
field, it is also subject to an opposite force exerted by its ion atmosphere,
which tends to restrain the ion and thus lowers its contribution to the
solution conductance. The effect increases with the density of ions in the
atmosphere, i.e., inversely proportion to XA. The second correction term
results from the electrophoretic effect. When the ion in question moves
through the solution, it tends to take its ion atmosphere with it, resulting
in a viscous drag force opposing the motion, thus the dependence on
viscosity.
Like Debye-Hiickel theory from which it is derived, eq (3.14) is a
limiting law valid only at very low concentrations, usually less than
O.OOlM.
Example 3.7 Compute the limiting law slope for a plot of the
molar conductivity ofKCl (AO = 149.8 x 10- 4 S m 2 mol- 1 ) us. -rc
assuming the dielectric constant and viscosity of pure water (10
=78.54, 'l1 =0.890 x 10-3 kg m- I s- l ), and compare with the slope of
the KClline in Figure 3.5.
Conductometric Titrations
Solution
------- H+
,, CI-
,, Na+
,, OH-
,,
,,
,,
,, ,
" ",'
'\, ;",.;
since the OR- molar ionic conductivity is very large, the conductance of
the solution rises again. The conductance of the solution is due, of
course, to all the ions present. This sum is demonstrated graphically in
Figure 3.7. In practice, a conductometric titration is carried out by
adding aliquots of titrant, measuring the conductance and plotting the
measured conductance vs. the volume of titrant added. The points can
usually be fitted to two straight-line segments which intersect at the
equivalence point.
This technique can be extended to any titration reaction which leads
to a change in the slope of the conductance vs. concentration plot at the
equivalence point. Any ionic reaction which produces a precipitate, a
gas, or any other nonelectrolyte is well suited for use in a
conductometric titration. Although the technique is very general, there
are practical limitations: (1) the change in slope at the equivalence point
is often small; (2) the analyte and titrant concentrations usually must be
greater than 0.001 M; and (3) nonparticipating electrolytes should be
absent.
3.3 DIFFUSION
N 1 -N2=--Ox
aN
ax
Since N is apparently a function of both x and t, we have written the
gradient as a partial derivative. Substituting this expression into eq
(3.17), we have
J=_D ac (3.19)
ax
where the diffusion coefficient D is given by
(Bx)2
D=- (3.20)
2&
with 81 units of m 2 s- 1 . This relationship was first established
empirically by Fick in 1855, and is known as Fick's first law of diffusion.
Notice that eq (3.19) tells us that the flux is positive (a net flow of
particles from left to right) only if the concentration gradient is negative
(there are fewer particles on the right to begin with). Thus this simple
equation contains the elements of the phenomenon of diffusion: flow of
particles into regions of lower concentration.
Fick also discovered a second empirical law, which was later found
to be derivable from the first law. Consider an infinitesimal volume
element of unit cross-sectional area and thickness dx. The flux into this
volume element is J(x), and the flux out of the volume element is J(x +
dx) = J(x) + (aJ/ax)dx. The difference between the flux in and the flux out
is the net increase in number of moles of particles in the volume element
per unit time. Dividing this by the volume of the element dx, we have the
time rate of change of the concentration ac/at.
ac = J(x) - J(x + dx) aJ
=--
at dx ax
Differentiating eq (3.19), we get
aJ =_Dic
ax ax 2
so we are left with
ac =_Dic (3.21)
at ax 2
which is Fick's second law. This result, of course, is a second-order
partial differential equation, which might not seem to be a particularly
simple way of expressing what we know about diffusion. However, it
turns out that the equation can be solved relatively easily for a wide
variety of boundary conditions.
Before we go on, it is instructive to write down the solution to eq
(3.21) for a particularly interesting case. Suppose that we bring two
solutions into contact at time t = O. In one solution the concentration of
the diffusing species is Co and in the other solution, the concentration is
zero; the initial interface is at x = O. We suppose that for t > 0, C ~ 0 as x
~ 00 and C ~ Co as x ~ - 0 0 . With these initial and boundary conditions,
the solution to eq (3.21) is (see Appendix 4)
C(x,t) = tco [1- erf{2illt}] (3.22)
The function erf(\jI) is called the error function and defined by the
integral
(3.23)
1.0
erfC",)
0.8
0.6
0.4
0.2
0.0
Figure 3.9 The error 0.0 0.5 1.0
function. 1.5 '" 2.0
(a) C=C o
(b)
-x 0 +x 0------.1.---=----
-x o +x
(c)
Figure 3.10 (a) Initial distrib-
ution of solute in a diffusion
experiment. (b) Concentration
profiles at successive times after
beginning of diffusion. (c) Con-
centration gradient at successive
times after beginning of
diffusion. -x o +x
are ions so that the chemical potential depends not only on their activity,
but also on the electric potential. Thus we have l
Ili = Ilio + RT In ai + ZiF(J) (3.26)
Differentiating eq (3.26) and substituting into (3.25), we have the Nernst-
Planck equation
v. =-~[kT d In ai + z.e~] (3.27)
'Ii dx 'dx
We now consider two limiting cases. First we assume that the
activity is uniform, but the electric potential nonuniform. Equation (3.27)
then becomes
Vi = + Zi e E
Ii
where we have recalled that the negative gradient of the electric
potential is the electric field strength, E = -d(J)/dx. Dividing through by E,
we have the mobility:
(3.28)
At 25°C, the factor RTIF has the value 0.0257 V. Thus, using
the mobilities computed in Example 3.6, we have
DOCNa+) = (0.0257 V)(5.2 x 10-8 m2V- 1s- 1)
DO(Na+) = 1.34 x 10-9 m 2 s- 1
DO(Ba2+) = (0.0257 V)(6.6 x 10-8 m2V- 1s- 1)/2
DO(Ba2+) = 0.85 x 10-9 m 2s- 1
,, ,,
,
o ___________ L - -_ _
d
I
+
\:.)
--------+
Figure 3.11 Diffusion of a 1:1 electrolyte solution in a tube. (a)
Concentration profiles at several times; the solid lines represent the cations
and the dashed lines represent the anions. (b) Corresponding concentration
differences, C+ - C_.
oni =~oQ
Zi F
In moving from x to x + dx, ion i moves through a chemical potential
difference, given by eq (3.26):
dlli = RT d In ai + FZi d<l>
where d<l> is the electric potential difference across the distance dx. The
total change in free energy then is
oG = I. dlli oni
1
138 Electrolytic Conductance
or
~<D=_RT"
F k."
i~ ELdlna· Zi '
(3.34)
, CI.
~<D = - R1
-
F
'l afa~ a~]
t1 In _1 - t2 In --..2...
afJ
or if we further assume that af =afJ =aCi. and af =ag = a~,
~<D = RT (t1 - til In a CI. (3.35)
F a~
Thus we predict that when two solutions, differing only in the
concentration of a 1:1 electrolyte, are in contact the liquid junction
potential is proportional to the difference in transference numbers of the
positive and negative ions and to the log of the activity ratio.
Thus we have
d In ai =d In ai dx =l dCi dx
ax Ci ax
ca-c~
dIn ai = ' I dx
Ci
The transference number of ion i can be written in terms of the
concentrations and mobilities:
t. _ IZi ICiUi
, - L IZjlCjuj
j
where we assume that the mobilities are independent of concentration
and thus of x. Substituting the expressions for ti and d In(ai) into eq
(3.34), we obtain
140 Electrolytic Conductance
where
a= LIZjlUPl
j
and
b=L IZjlUj (eJ - ell
j
t ~=ln(a
Jo a + bx
+ bx)1 1 =~lnCL.±..b..
bob a
Thus
(3.36)
L uJzil ep 1
~<l> = ± RT In
F
rL u,jz;i er
..":i:,,,-_ __ (3.37)
Comparing the results of Examples 3.13 and 3.14, we see that the
use of a saturated KCl salt bridge is expected to reduce the junction
potential for 0.1 M KCl and HCl solutions by nearly a factor of 10. The
secret, of course, is that the K+ and Cl- ions, which have very nearly
equal mobilities, carry most of the current at the salt bridge/solution
junctions. The result is generally applicable, however, and explains in
part the popularity of KCI as an electrolyte, particularly when liquid
junctions are unavoidable.
chemical potentials of the small ions and solvent! will be equal in the
two phases. However, since the condition of electrical neutrality
requires that the ionic concentrations be unequal in the two solutions, an
electric potential difference must develop across the membrane. Unlike
a liquid junction potential, however, the potential in this case does not
depend on the relative mobilities of the ions and the system is in a true
equilibrium state. The development of potentials across semipermeable
membranes was first studied by Donnan in 1911 (17) and the
phenomenon is usually referred to with his name.
Ion-Selective Membranes
a
outer electrolyte solution
~outer surface of the membrane
y interior of the membrane
~' inner surface of the membrane
a' inner electrolyte solution
The total membrane potential then has four contributions:
~<l> = ~<l>a~ + ~<l>~y + ~<l>~' + ~<l>Wa' (3.40)
We will assume that the membrane surfaces are in equilibrium
with the adjacent electrolyte solutions so that the arguments used in
discussing Donnan membrane equilibria can be used for the a-~ and W-
a'interfaces. If H+ is in equilibrium across the a-~ interface, then the
H + chemical potentials must be equal for the two phases:
!lHO(a) + RT In aHa + F<l>a = !lHO(~) + RT In aH~ + F<l>~ (3.41)
or
t
result (Dll,DI4), which can be expressed by
.1<l>memlrane = constant + Rll n [ai + kijaj ] (3.49)
where species i is the principal diffusing species of interest and kij is the
selectivity coefficient for j relative to i.
REFERENCES
PROBLEMS
3.1 (a) A conductance cell filled with 0.1 ~ KCl solution had a
resistance of 24.96 n at 25 C. Calculate the cell constant (LIA)
Q
3.3 Compute the Onsager limiting law slope for AgN03 and compare
it with the slope of the plot obLained in Problem 3.2.
3.8 Compute the Stokes law radii, the ionic mobilities, frictional
coefficients, and diffusion coefficients of Fe(CN)63- and Fe(CN)6 4 -.
(a) Compute the "equivalent weight" (MWln) and use this value to
convert the concentrations to equivalents per cubic meter (nClmol
m -3) and the conductances to "equivalent conductivities" (A/n)
with units ofS m 2equiv- l .
(b) Plot Aln us. 'InC, determine the intercept and the slope. How
does this slope differ from the slope of a plot of A us. VC?
(c) Given the molar ionic conductivity of K+, 73.5 x 10-4 S m 2moI-l,
determine the transference numbers, t+ and L.
(d) Suppose that the salt is a 1:1 electrolyte (n = 1). Does the slope
obtained in part (b) agree with theory? Suppose it is a 2:1
electrolyte? Is the fit better? A 3:1 or 4:1 electrolyte?
Problems 149
3.10 Verify that eq (3.22) is a solution to Fick's second law, eq (3.21).
3.16 150 m.L of a saturated solution of Ba(OH)2 was titrated with 1.00 M
H2S04 and the resistance of the solution measured with a dip-type
cell. The following data were obtained:
lOur sign convention for indicator electrode current is the usual one among
electroanalytical chemists, but many other electrochemists employ (and the
International Union of Pure and Applied Chemistry recommends) the opposite
convention with positive anodic current.
151
152 Voltammetry of Reversible Systems
mechanisms. Thus the solution should not be stirred and the cell
should be free of vibration. In addition, we will require that the ratio of
electrode area to solution volume be small and that experiments be short
in duration so that relatively little electrolysis takes place. This
precaution will minimize local changes in concentration or local
heating which could lead to density gradients and convective mixing.
The solution should contain a large excess of an inert electrolyte (called
a supporting electrolyte) to ensure that the double layer is very compact
and that the electric potential is nearly constant throughout the solution,
thus minimizing electric migration effects.
1 Strictly speaking, eq (4.1) should include the activity coefficient ratio, YO/YR, but
we will ignore this complication; thus the standard potentials used here are really
formal potentials-see §1.3.
2 The use of Laplace transform methods for the solution of differential equations is
discussf'd, and eqs (4.2) are derived, in Appendix 4. See also Delahay (Bl), Bard
and Faulkner (B12), and MacDonald (F5).
§4.1 Diffusion-Limited Current 155
1.0
CIC O*
0.8
0.6
0.4
Current-Potential Curves
Jo(x,t) = - Do ~CoCx,t)
ax
Using eq (3.24) to differentiate eq (4.2a), we obtain
VDohr.t -x2
JoCx,t) = - Co exp - -
I + ~e 4Dot
Setting x = 0, we have the flux at the electrode surface
Equation (4.11) was derived by Heyrovsky and Ilkovic in 1935 (2) and is
called the HeyrovskY-Ilkouic equation.
Because of the term involving the diffusion coefficients, half-wave
potentials are not standard potentials (or formal potentials), even for the
electrochemically reversible processes we will be discussing. However,
even if the diffusion coefficients differ by a factor of 2, E 112 will differ
from EO by less than 10 mV (for n = 1, 25°C) and for many practical
purposes, half-wave potentials measured from current-potential curves
can be used as approximations to standard potentials. Some examples of
such applications are discussed in §4.9.
1.0
iliD
0.8
0.6
0.4
0.2
When i = 112 iD ,
(iD -i)/i = 1
so that E = E1J2, hence the name half-wave potential. When i =
1/4iD and 3/4 iD ,
(iD - i)/i =3 and (iD - i)/i = 113
Thus
or, at 25°C,
E 114 - E3/4 = 56.51n m V
This property of a reversible wave was proposed in 1937 by
Tomes (3) as a quick indication of nernstian behavior and is
called the Tomes criterion of reversibility.
JO(X,t) = -- Do aCo(x,t)
ax
O.B
2.0
0.6 if iD
1.5
~ 0.4
.8
I
<.> 0.2 1.0
.8
~ 0.0
0.5
iDa
-0.2
0.0
-0.4
0.2 0.1 0.0 -0.1 -0.2 0.2 0.1 0.0 -0.1 -0.2
(E -E 1/2 )N (E - E 1/2 )N
1.0
iliD
0.5
0.0
-0.5
Figure 4.5 Current-potential
curve for reversible electron
transfer where transport of
o is diffusion limited but R -1.0
is a solid on the electrode
surface, n = 1, Co* = 1 mM. -0.1 -0.2 -0.3 -0.4
(E -EO)N
100
jlA m- 2
10
1
Figure 4.6 Diffusion-
limited current density
at spherical electrodes
with ro = 1 ~m - 10 mm; 0.1
note the decrease of the
steady-state contribution 0.001 0.01 0.1 1 10 100
with increasing radius. tis
r Current
Output
of the potentiostat along with a constant potential source which sets the
initial reference electrode potential.
Still other control potentials can he added to program the reference
electrode potential as required. Thus a triangular wave potential is used
for cyclic voltammetry (§4.4), a sinusoidal a.c. potential for a.c.
polarography (§6.4), and various kinds of voltage pulses for pulse
polarography (§4.6).
The three-electrode configuration avoids the problem of reference
electrode polarization and somewhat reduces the measured iR drop in
the solution. One way of further reducing solution iR drop is to locate
the reference electrode junction as close as possible to the indicator
electrode, often through use of a salt bridge with a long curved capillary
tip as shown in Figure 4.12b (a Luggin probe).l Even then, iR drop often
remains a problem and most commercial electrochemical instruments
contain a circuit which partially compensates for iR drop by electrically
subtracting an adjustable voltage, proportional to to the current, from
the measured cell potential.
Control of Current
In some experiments (e.g., chronopotentiometry; see §4.3), a
constant current is pa::;::;ed through the cell, and the potential of the
indicator electrode is monitored using a reference electrode. An
operational amplifier circuit which performs this function (a
galuanostat circuit) is shown in Figure 4.9. Since the operational
amplifier input impedance is high, the cell current must flow through
the resistor R. Feedback through the cell ensures that the potential at
the inverting input is at virtual ground. Thus the potential drop across
the resistor is equal to the battery potential and the cell current is
i = f,,¢IR
With the battery polarity as ::;hown in Figure 4.9, the indicator electrode
is the cathode; reversal of polarity would change the direction of current
flow. In this circuit, the ultimate source of current is the battery; the
function of the operational amplifier is to adjust the potential of the
auxiliary electrode to keep the current constant. In experiments where
the current is to be controlled as a function of time, the battery can be
replaced by a potential program circuit as described above.
Choice of Solvent
>-.......--0 Reference
Potential
Monitor
Supporting Electrolytes
Voltammetry and related electrochemical techniques require an
excess of inert electrolyte to make the solutions conducting and to reduce
the electrode double-layer thickness. In principle, any strong electrolyte
will satisfy these basic requirements, but there are other considerations
(6).
In aqueous solutions, KCI and HCI are common choices since the
liquid junction potentials with s.c.e. or Ag/AgCI reference electrodes can
be eliminated. Aqueous solutions are usually pH-buffered and the
components of the buffer system often act as the supporting electrolyte.
Clearly, the salt should not be easily oxidizable or reducible. The
accessible potential range at Pt and Hg electrodes is given in Table A.9
for several electrolyte/solvent combinations.
In low-dielectric-constant solvents, tetraalkylammonium salts are
more soluble and are less easily reduced than alkali metal salts.
Quaternary ammonium ions do not form tight ion pairs with anions.
Thus R4N+ salts are by far the most common choice, although
tetraphenylphosphonium salts and lithium salts are sometimes used in
organic solvents. When tetraalkylammonium ions are reduced,
however, they form surface-active polymers which coat electrodes, foul
dropping mercury electrode capillaries, and generally raise havoc with
experiments.
The choice of anion is less obvious. The simple halides, CI-, Br-, and
1-, are relatively easily oxidized and often form tight ion pairs, so that
they are frequently avoided. The most common choices-CI04-, BF4-,
PFs-, and BPh4--have delocalized charge, so that their salts are often
soluble in organic solvents. Ion pairing usually is not severe and these
anions are not easily oxidized or reduced at electrodes.
Reference Electrodes
In aqueous solutions, the calomel and Ag/AgCI electrodes are well
characterized and give reproducible potential readings with a minimum
of experimental difficulty. Unfortunately, there is no universally
accepted reference electrode for use with nonaqueous solvents. There
are two commonly used approaches.
One school of thought is to use an aqueous s.c.e. with a salt bridge.
The advantage is that the reference electrode is well understood and
generally reproducible. There are some major disadvantages: an
unknown liquid junction potential is introduced and water
contamination through the salt bridge is difficult to avoid completely.
Since the s.c.e. electrolyte is normally KCI, contamination by K+ and CI-
ions can sometimes lead to problems. For example, a KCI-filled salt
bridge in contact with a solution containing CI04- ions frequently leads
to precipitation of KCI04, which is quite insoluble in organic solvents
§4.2 Experimental Techniques 171
and not very soluble in water. The precipitation usually takes place at
the point of solution contact, often clogging the salt bridge and blocking
current flow. One way of avoiding this particular problem is to replace
the s.c.e. electrolyte with NaCl; NaCl04 is much more soluble.
A second common approach is to use the Ag/Ag+ couple as a
reference, dissolving a suitable silver salt in the same
solvent/supporting electrolyte system used in the experiment and
placing it in contact with a silver wire. The advantages of this approach
are that the liquid junction problems and solvent and electrolyte cross-
contamination problems are minimized. The disadvantage is that
solvent evaporation from the reference electrode or change in the silver
surface with time may lead Lo nonreproduciblc potentials.
With many different reference electrodes in use, potential
measurements from different laboratories are often difficult to compare.
It is becoming standard practice in electrochemical studies of
nonaqueous systems to use a standard reference couple of known (or at
least commonly accepted) potential against which the reference
electrode can be checked from time to time; potentials can then be
reported relative to the standard couple or corrected to a common scale.
Ferrocene, (C5H5)2Fe, is reversibly oxidized to the ferrocenium ion,
(C5H5hFe+, at +0.08 V us. Ag/Ag+ in acetonitrile (7) and is the most
commonly used potential standard. Ferrocene and ferrocenium salts
are soluble in most nonaqueous solvents; in addition, the
ferrocene/ferrocenium couple is relatively insensitive to solvation or ion-
pairing effects and so provides an approximation to an absolute
reference.
Indicator Electrodes
1 For reviews on electrodes, see Adams (8,C3), Galus (9), Dryhurst and McAllister
(10), or Winograd (11).
172 Voltammetry of Reversible Systems
contact
bushing
Teflon
electricalvA-...---_ insulation
contact
(1"electrode
capillary
':::::?'
Figure 4.10 Dropping Figure 4.11 Rotating-
mercury electrode. disk electrode.
through which mercury flows, forming a drop at the end of the capillary
which grows until its weight exceeds the force of surface tension holding
it to the capillary. Depending on the length and bore of the capillary, the
pressure of mercury above the capillary, and the Hg-solution interfacial
tension, the lifetime of a mercury drop can be anywhere from 1 to 10 s.
The factors governing the drop time were discussed in §2.S. The current
through a d.m.e. is time dependent, but because the solution is stirred
when a drop falls, each new drop starts the experiment anew. If the
current is measured just before the drop falls, experiments using a
d.m.e. are essentially at constant time. We will discuss the operation of
the d.m.e. in greater detail in §4.S.
The rotating-disk electrode (r.d.e.), shown in Figure 4.11, consists of
a flat disk, usually 1-3 mm in diameter, mounted at the end of an
insulating rod which is rotated rapidl.y in the solution. The rotational
motion stirs the solution so that the diffusion problem is reduced to
transport across a stagnant layer at the electrode surface. Since the
rotation speed is constant and the stirring effect is reproducible, an
experiment using the r.d.e. is carried out under steady-state (i.e., time
independent) conditions. We will discuss the operation of a r.d.e. in
greater detail in §4. 7.
Cell Design
(a)
+--N2
d.m.e.
waste Luggin
Hg probe auxiliary
electrode
like this are quite well suited to most voltammetric techniques and work
well, provided that the solvent is not too volatile. The cell body can be
fitted with a flow-through thermostat jacket for work away from room
temperature and the entire experiment can be mounted in a glove box in
the event that the analyte is very sensitive to air or moisture.
When solvent volatility is a problem or when the sample is
exceedingly sensitive to air or moisture, cells are often designed to be
filled on a vacuum line so that the purge gas is unnecessary.
iDe = OoCO*
iDa = - ORCR*
where
00 = nFAkvo
OR = nFAkvR
It is also convenient to let X be the mole fraction of the electroactive
material in the oxidized form and C* be to the total concentration so that
Co* =XC*
CR* = (l-X)C*
With these changes in parameters, eq (4.17) becomes
E =Ell'!. + RT ooXC* - i
~ In ---=----- (4.22)
nF i + oR(l-X)C*
Equation (4.22) describes the interrelationships of three variables:
potential, current, and composition. These interrelationships are most
easily visualized by thinking of a surface in three-dimensional space as
shown in Figure 4.13. 1 We can describe a number of electroanalytical
techniques as excursions on this surface.
Potentiometric Titrations
When the current is equal to zero, eq (4.22) reduces to the Nernst
equation, which describes a potentiometric titration (see §1.7) in which
the cell potential is measured as a function of added titrant, i.e., as a
function of solution composition. The potentiometric titration curve
corresponds to the zero current path across the surface of Figure 4.13 as
shown in Figure 4.14. Of course, in a real case, the titrant and its
reduced form (assuming that the titrant is an oxidizing agent) form
another electrode couple, and beyond the endpoint it is this couple which
determines the cell potential. Thus, to complete the titration curve, we
would have to graft another similar surface onto Figure 4.13, so that
when X approaches 1, the potential goes to some finite value rather than
to infinity as implied by the single surface.
Amperometric Titrations
1.0 Ita)
Cr<£)i2-
0.5 r-
I I
0.0
0.0 -0.2 -0.4 -0.6 o 4 8 ~ ffi
EIV Vol. K2Cr207 solutionlmL
Figure 4.15 (a) Polarograms of Cr2072- and Pb 2 + and (b) amperometric
titration curve for the titration of 25 mL of 0.001 M Pb 2+ with 0.0015 M Cr2072-
atE =-0.7 V.
or
RTI Kc-i'ft
E = E 112 + - - n ------==----- (4.23)
nF i'ft +Ka
The interrelationship of potential, current, and time (for fixed solution
composition) given by eq (4.23) can also be represented by three-
dimensional surface shown in Figure 4.16. A number of additional
techniques can be understood as excursions upon this surface. l
Chronoamperometry
Figure 4.17 shows the potentials and currents for such an experiment as
functions of time. Suppose that the current is measured at time t1 after
§4.3 A Survey of Electroanalytical Methods 181
the first potential step and again at time t2 after the second step. The
ratio of the two currents will be
~= {f; - ~/ t2t~t
or iftl = t, t2 = 2t,
!1
il
= '1fT2 -1 =-0.293
V
The current after the second step is smaller in magnitude than that
after the first step since some R escapes into the bulk solution. This
measurement provides a good indication of an uncomplicated reversible
electron transfer. If the product R was consumed by a chemical
reaction, for example, the current after the second potential step would
be less than expected from eq (4.26), I i2/ill < 0.293 (see Example 5.7).
-----------
~ ---------.:..:.
if
-----------
Chronopotentiometry
the required current and the potential heads toward - 0 0 . The time 't
required for this potential swing is called the transition time,
-Fe = Kc1i
In reality of course the potential doesn't go to since there is always
-00
-0.3
-0.2
.
~
-.j'
~
I
-0.1
ES 0.0
0.1
1 _
1+~8(t) -
iat
0
X(z)
Ycrt-z
dz (4.30)
Comparing the results of Examples 4.1 and 4.5, we see that linear
potential scan peak currents can be quite large. Because of this, ohmic
potential drop in the solution is often a problem. Since the cell potential
includes the solution iR drop and the iR drop varies as the peak is
traversed, the indicator electrode potential is not linear in the applied
potential. Thus the observed peak potential becomes a function of scan
rate. In practice, the revcrsible peak potential usually can be extracted
from experimental data by measuring the peak potential at several scan
rates and extrapolating to v == 0, but the absolute peak current is more
difficult to correct.
186 Voltammetry of Reversible Systems
Cyclic Voltammetry
time potential
Figure 4.20 Cyclic voltamrnetry: (a) potential as a function of time; (b)
current as a function of time; (c) current as a function of potential.
§4.4 Cyclic Voltammetry 187
When the switching potential is less negative such that the cathodic
current is still significant at the switching point, the diffusion layer is
thinner, the R concentration falls to zero more rapidly with distance
from the electrode, and the resulting anodic peak is smaller. However, it
turns out that, if the anodic peak current is measured from a baseline
equal to the cathodic current which would have flowed at the time of the
anodic peak had the potential scan continued in the negative direction
(rather than from the zero current baseline), then the anodic-to-cathodic
peak current ratio is exactly 1 (for a reversible process uncomplicated by
capacitive current and iR drop). This result is shown schematically in
Figure 4.21 with cyclic voltammograms computed using the numerical
methods of Nicholson and Shain (19).
Provided that the process is reversible and is uncomplicated by
solution iR drop, Epc -El/2 = -28.5/n mV and Epa -E1I2 = 28.5/n mV (at
188 Voltammetry of Reversible Systems
25
o 5 10 15 25 0 5 10 15 25
xillm xillm
Figure 4.23 Concentration profiles for 0 (solid lines) and R (dashed lines) at
various points on a cyclic voltammogram. Concentrations computed using a
dig!tal simulation technique (see Appendix 5) for v = 1 V 5- 1 , DO = DR = 5 x
10- 10 m 2 s- 1.
25°C). Thus E1/2 = (Epc + Epa)/2 and Epa - Epc = 57.Dln mY. This peak
separation is often used as a criterion for a reversible process. l
_1
fit
it
0
met) dt
vt-t
=
it
0 ret) dt
In other words, m (t) is mid-way between the function f(t) and its
integral. Accordingly, m (t) is called the semi-integral of f, written
formally as
(4.36)
....
"'"' (a)
1...,
Q)
~
.';'
's
Q)
rn
4.5 POLAROGRAPHY
r
Substituting eq (4.38) in eq (4.7) and replacing Do by (7/3)Do gives
iD = nF [41t (!:~ 3
J Co* C~: (2
1 The convention that cathodic current is taken as positive results from this feature;
polarographic currents are usually cathodic and much of the theory of voltammetry
was developed with polarography in mind.
196 Voltammetry of Reversible Systems
1 The original work on adsorption effects in polarography was done by Brdi&a (27)
in the 1940's.
Sec. 4.5 Polarography 199
(4.41)
Thus, if AGado < 0, the two reactions are expected to be distinguished by
different half-wave potentials. If the bulk concentration of 0 is small
enough that less than monolayer coverage by R is achieved during one
drop life, we expect a wave at E10 with a limiting current given by eq
(4.39). If the drop is completely covered before it falls, then we expect the
current to be limited by the available adsorption sites. Excess 0 then
remains unreduced at the electrode surface when the potential is in the
vicinity of Elo and a second wave is observed at E2°. The total diffusion
current of the two waves is limited by diffusion and is given by eq (4.39).
The second wave, of course, is the expected one for a process
uncomplicated by adsorption. The first wave is called an adsorption
prewave.
A very similar phenomenon occurs when the substrate 0 is
adsorbed on the mercury drop (we assume that R is not then adsorbed).
For low bulk concentrations of 0, the electrode will be partially covered
and the electrode process will be
Oad + n e- f:! R
When Co* is big enough that complete coverage IS achieved, direct
electron transfer may be possible:
Osoln + n e- f:! R
The potential of the first process will be more negative (if AGado < 0), but
the main wave will occur only at higher concentrations. The more
negative wave is called an adsorption post-wave. The appearance of
adsorption pre-waves and post-waves are shown in Figure 4.28.
Adsorption pre-waves and post-waves can be distinguished quite
easily from ordinary electron-transfer processes. In addition to seeing
the normal wave at higher substrate concentrations, heights of
adsorption waves have a unique dependence on drop size and the
variation of current during the drop life is different from the normal
behavior. For further details, a specialized book on polarography should
be consulted (D1,D4,D5).
Polarographic Maxima
2.0
(a)
iliD2
1.5
1.0
1.0
0.5
0.5
0.0
-0.3 -0.4 -0.5 -0.6 -0.7 -0.3 -0.4 -0.5 -0.6 -0.7
EIV EIV
Figure 4.28 Polarograms for various substrate concentrations, CO*, where
Ca) the product Rand Cb) the reactant 0 are adsorbed on the mercury drop.
The curves correspond to CO*/CO' = 0.5, 1.0, 1.5, and 2.0, where CO' is the
concentration leading to monolayer coverage at the time of drop fall.
tend to drag the surface layer along, thus leading to convective mixing of
the solution. Type II maxima are less dramatic in their effect; the
convective contribution to mass transport is only weakly potential
dependent, and the effect may pass unrecognized in many cases. Type
II maxima depend on the mercury mass flow rate u and are most easily
recognized by a departure from the dependence of current on flow rate
predicted by the llkovic equation.
Type III maxima arise from nonuniform adsorption of surfactants
which lead to variations in interfacial tension and thus to convective
flow of solution in the vicinity of the mercury drop. Like Type I maxima,
these effects can be reduced or eliminated by adding a competing
surfactant.
Polarographic maxima are more common in aqueous solution
polarography than when organic solvents are used, presumably because
the mercury surface is hydrophobic and therefore susceptible to
surfactants in aqueous media, whereas many organic solvents are
themselves somewhat adsorbed on the mercury surface.
Pulse Polarography
A weakness of both ordinary d.c. and sampled d.c. polarography is
that these techniques allow faradaic current to flow during a time in the
drop life when the capacitive current is large. A considerable
improvement is possible with a technique called pulse polarography,
introduced by Barker in 1960 (29). A series of voltage pulses is applied to
the cell as shown in Figure 4.30; the drop time is synchronized to the
pulses which increase in amplitude at a rate comparable to the ramp
voltage used in ordinary polarography. The total current response from
such a pulse train still has a large capacitive component, but this is
largely suppressed by sampling the current for a short time late in the
pulse life after the capacitive current has mostly decayed. The sampled
current is converted to a voltage which is held and applied to the
recorder until the next sample. The appearance of a pulse polarogram,
Figure 4.29c, thus is virtually indistinguishable from a sampled d.c.
polarogram.
As we saw in the derivation of the llkovic equation, the current
varies with time because of the time-dependent electrode area and
because of the growth of the diffusion layer with time. In pulse
polarography, essentially no current flows until late in the drop life, so
that the diffusion layer is much thinner at the time the current is
actually measured. In particular, we expect the measured current to
depend on the times t1 and t2 (see Figure 4.30):
i oc t1213t2-1/2
Time t1 is the age of the drop when the current is measured (t1 thus
determines the drop area) and t2 is the time between application of the
voltage pulse and current measurement (t2 thus determines the
diffusion layer thickness). Suppose that the drop time is 5 s, that the
pulse is applied in the last 50 ms of the drop life, and that the current is
measured during the last 10 ms of the pulse. Then t1 = 4.99 s, t2 = 0.04 s
and
t12/3 t2-1/2 = 14.6
In sampled d.c. polarography, t1 =t2 =4.99 s, so that
t12/3 t2-1/2 = 1.31
(a) (b)
1.0 -
illlA
0.5
0.0 f---------------
I
(c) (d)
o -----------
-0.4 -0.6 -0.8 -0.4 -0.6 -0.8
EIV EIV
Figure 4.29 (a) D.c., (b) sampled d.c., (c) pulse, and (d) differential pulse
polarograms of 0.1 mM Cd(N03)2 in 0.10 M aqueous KCI. Potentials vs.
s.c.e., drop time 2 s, scan rate 2 mV s·l, differential pulse height 10 mV.
pulse train
pulse train
1.00
i/i p
0.75
0.50
Die =Vx ae
ax 2 ax
210 Voltammetry of Reversible Systems
We take Vx'" -afiiN (;(,/XH)2, an acceptable approximation for X/XH < 0.2,
and use the boundary condition C -+ C* as x -+ 00. Defining the
dimensionless variable
u = (a/D)113 0)112 y-116 x
the differential equation can be written as
£c.=_u 2 dC.
du 2 du
Setting C' = dC/du, we can integrate
(c ~=_
leo' C
r)0 u 2 du
In(C'/Co') = - u 3/3
or
dC.:::(dC) exp(-u3/3)
du du 0
(f) i
Integrating a second time, we have
U(X)
We will need the derivative (dC/du)o in order to compute the flux at the
electrode surface and the current. To evaluate it, we extend the integral
limits to infinity and apply the boundary condition C(x) -+ C* as x -+ 00.
dx 0
(de)
= C* - C(O)
XD
§4.7 The Rotating Disk Electrode 211
where Xn is the diffusion layer thickness. Combining the last two
equations, we can solve for Xn:
xn = 1.288 (D fa )1130)-112 v1l6
or, inserting a = 0.510,
xn = 1.61 D1/3 0)-112 v 1/6 (4.45)
The ratio of xn to XH then is
Xn/XH = 1.61 (Dfv)1/3
For an aqueous solution (D", 10- 9 m 2 s- 1, v'" 10-6 m 2 s- 1) XnfXH '" 0.16,
consistent with the assumption that the linear form of eq (4.42c) could be
used in eq (4.43). This result affords a considerable simplification in
practical calculations on experiments using the r.d.e. The same result
could have been obtained had we neglected the forced convection term in
eq (4.43) and solved the simple equation
Dd~ =0
dx 2
We will have recourse to this simplification in Chapters 5 and 6.
Consider now the standard electrode process
o +n e- ~ R
The flux of 0 at the electrode surface is given by
Jo(O) =-Do(~)o
or
~ (0) =-D Co*-Co(O)
o 0 xn
As usual, the electrode current is
i = - nFAJo(O)
i == nFA(Do/xn)[Co* - CoCO)]
iL = nFADoCo*fxn (4.46a)
or, with the mass-transport rate constant, kn = Dofxn,
iL = nFAknCo* (4.46b)
Substituting eq (4.45), we have
iL ::: 0.62 nFACo*Do2/3 v- 1I6 0)112 (4.47)
212 Voltammetry of Reversible Systems
Equation (4.47) was first derived by Levich in 1942 (35) and is usually
called the Levich equation. Notice that eqs (4.46) are identical to eqs
(4.15). The analogy with the planar diffusion problem is still more
complete, however, since we recall that we used eqs (4.15a) and (4.16),
together with the Nernst equRtion, to derive the Heyrovsky-Ilkovic
equation, eq (4.17), and this result is also valid for a reversible process at
the rotating-disk electrode. Since the flux ofR can be written
JR(O) = hDR CR(O)
the current can be expressed in terms of CR(O):
i = nFAhDR CR(O)
Thus we can write
iL-i _ h DO CoCO)
- - - hDR CR(O)
or, taking logs, using eq (1.45) to evaluate XDO and XDR, and assuming
nernstian behavior, we obtain the Heyrovsky-Ilkovic equation:
RT
E =Elf2+-ln-.-
iL-i
nF l
where
El/2 =Eo - 2RT In Do (4.48)
3nF DR
Current-potential curves for the r.d.e. thus look exactly like those for a
planar electrode (Figure 4.2) at const8nt time, but half-wave potentials
may be shifted slightly.
Concentration profiles for a r.d.e. can be computed by numerical
integration of eq (4.44) and are shown in Figure 4.34. The concentration
gradient actually is quite linear over some distance from the electrode.
Comparison with Figure 4.1 shows that the assumption of a linear
concentration gradient is rather closer to the truth for the r.d.e. than for
a stationary planar electrode. This of course means that results derived
from a simplified linear concentration gradient model will be more
accurate for the r.d.e. than for the d.m.e. or a planar electrode.
§4.7 The Rotating Disk Electrode 213
1.0
CICo
0.8
0.6
bulk
solution
0.4
0.2
We see from Examples 4.6 and 4.8 that currents expected for a 1-
mm r.d.e. are about a factor of ten larger than those expected for a d.m.e.
This has advantages and disadvantages. Everything else being equal,
214 Voltammetry of Reversible Systems
contact
bushings
Teflon
insulation
. electrodes
Figure 4.35
electrode.
The rotating ring.disk @f
':::?'
§4.7 The Rotating Disk Electrode 215
4.8 MICROELECTRODES
Beginning in the late 1970's, a lot of effort has been devoted to the
development of microelectrodes, electrodes with dimensions on the order
of 0.1-25 /lm. This effort has been motivated by the need for very small
electrodes for in vivo biological studies, by the recurring problem of iR
drop in conventional voltammetric experiments, and by the need for new
methods capable of probing electrochemical processes on a very short
time scale. Progress has been made along these lines, but meanwhile
microelectrodes have been found to have other advantages, unlooked for
when they were first conceived, and they have become important tools in
the electrochemist's bag of tricks.
Microelectrodes have been constructed in a number of different
geometries, including inlaid microdisks, microcylinders, and
hemispheres, as well as inlaid bands and arrays. For any geometry, the
fabrication of a microelectrode is nontrivial, involving
micromanipulation techniques and/or some of the microlithographic
methods used in the fabrication of integrated circuits. Wightman and
Wipf (37) describe some of these methods and provide references to the
literature. A more recent paper by Baer, et al. (38), describes improved
methods for the construction of microdisk electrodes. By far the most
common geometry has been the inlaid disk, and we will confine our
attention to such electrodes. The construction of a typical microdisk
electrode is shown in Figure 4.36.
We begin with consideration of the diffusion problem in the context
of potential-step chronoamperometry and then discuss the use of
microelectrodes in cyclic voltammetry. Applications of microelectrodes
in studies of homogeneous and electron-transfer kinetics will be
discussed in §5.3 and §6.3. For further details, the reviews by Wightman
and Wipf(37) and by Bond, Oldham, and Zoski (39) are recommended.
216 Voltammetry of Reversible Systems
lead
wire
epoxy
glass
tube
conducting
epoxy
Figure 4.36 Cross-sectional view of a
microdisk electrode constructed using 4-
mm glass tubing.
ac
at
=D(ic
ar2
+ 1 ac +
r ar
iC)
az2
The boundary conditions imposed by the electrode geometry are
C -+ C* as r -+ 00, Z -+ 00
10
~ ~:-::~-=~-=~-:-~::-~~-~-~-_..J
The electrode areas are n:r02 = 3.14 x 10- 12 and 3.14 x 10-6 m 2;
substituting into eq (4.50), we have
in = 0.096 and 96 nA
The current densities are
in/A = 123 and 0.123 A m- 2
Substituting into eq (4.51), we get the mass-transport rate
constants are
kn = 1.27 x 10-3 and 1.27 x 10-6 m s-1
Notice that the current is 1000 times larger at the 1-mm
electrode, but the current density and mass-transport rate
constant are 1000 times larger at the l-I-lm electrode!
Comparing this result with Example 4.8, we see that kn is 10-
100 times bigger for steady-state diffusion to the 1-l-lm electrode
than under forced convection conditions at a rotating-disk
electrode. The current is close to steady-state for t > 10 ro2/D.
Thus we have
tsteady-state = 0.01 s and 10 4 s
For the 1 mm electrode, more than two hours is required to
reach steady-state conditions, by which time convective mixing
will long since have perturbed the experiment.
J:_2_n_A. . .". -_
-:1~I:::2==nA=~= ~
peaks develop, and, at the fastest scan, the shape approaches that
expected for cyclic voltammograms using macroelectrodes. Second, we
notice that, for slow scans, the current is independent of scan rate,
approaching the usual ..v-dependence only when the peaks are well
developed.
This behavior is to be expected, given our results for
chronoamperometry. For a microelectrode with ro = 6 11m, we would
expect essentially time-independent current about 0.4 s after application
of a potential step (assuming D '" 10. 9 m 2 s· 1). In an experiment where
0.4 s is a long time (cyclic voltammetry at 10 V s·l), the steady state is not
reached and the voltammogram has a "normal" shape and the current
increases as..v. When 0.4 s is a short time (cyclic voltammetry at 10 mV
s·l), the steady state is reached early in the experiment and is
maintained. Under steady-state conditions, the current depends only on
the potential; it is independent of time and thus of scan rate and scan
direction. Notice the absence of cathodic current on the reverse scan at
0.01 and 0.1 V s·l. Apparently no ferrocenium ions remain near the
electrode to be reduced; in the steady-state regime, the electrode product
diffuses away rapidly and irretrievably. Thus steady-state cyclic
voltammograms do not provide information about the chemical stability
of a reduction or oxidation product.
222 Voltammetry of Reversible Systems
4.9 APPLICATIONS
-2.2 1.6
Ell2N
-2.0 1.4
-1.8 1.2
-1.6 1.0
-1.4 0.8
-1.2 0.6
-1.0 0.4
0.3 0.4 0.5 0.6 0.7 -0.3 -0.4 -0.5 -0.6 -0.7
(E + a)/~ (E + a)/~
-4.0 -2.0
EV2N 1
-3.8 -1.8
-3.6 -1.6
-3.4 -1.4
-3.2 -1.2
-3.0 -1.0
-2.8
-0.8
1950 2000 2050 2100 2150 1970 1980 1990 2000
vCO/cm- 1 v CO /cm- 1
easily reduced species is nearly complete, then all CiO) = and the
current is diffusion limited,
°
When the potential is sufficiently negative that reduction of the most
m
in = nFA I knj C/
j=O
Subtracting i from in, we have
m
in - i = nFA I knj C;(0)
j=O
If the complexation steps are fast enough that they are at equilibrium
even at the electrode surface, and if the free ligand concentration is
§4.9 Applications 229
sufficiently high that CL is uniform throughout the solution, then we
have
Thus
m .
iD - i == nFACo(O) L kDj PjCr: (4.53)
j=O
Reduction of a metal complex at the d.m.e. produces metal atoms
which dissolve in the mercury and diffuse into the interior of the drop
with a flux
JM(O) == kDM [CM* - CM(O)]
where kDM is the rate constant for diffusion of M in the amalgam; we
assume that the concentration of metal in the interior of the drop, CM*,
is negligible. This flux is also related to the electrode current,
(4.54)
If we assume that the electron-transfer process is reversible, the
concentration ratio COCO)/CMCO) must be given by the Nernst equation
CoCO) nF(E - EO)
CMCO) == exp RT
Using eqs (4.53) and (4.54), this ratio can be written
CoCO) - - iD -=- i- - - xkDM
--
CMCO) L kDjpjCr:
Combining the two concentration ratio expressions, we have
E==Eo+RTln kDM +RTlniD-i (4.55)
nF L kDjPjCr! nF i
or
RT
Lilll/2 =_RT In p. -J-In CL (4.57)
nF 'J nF
Equation (4.57) was first used by Lingane around 1940 and is sometimes
called the Lingane equation.
-0.60
E1/2N
-0.65
-0.70
-0.75
Figure 4.44 Polarograph-
ic half-wave potentials
(us. s.c.e.) for Pb(II) in
KN03iNaOH solutions at -0.80
25°C; data from reference 0.0 0.5 1.0 1.5 2.0
(50).
-log[OH-]
§4.9 Applications 231
Analytical Applications
Polarography was the first partially automated instrumental
analytical technique and the first technique capable of routine analysis
at the submillimolar level; as such polarography had a tremendous
impact on analytical chemistry and was widely used from the 1930's.
With the development of competing instrumental methods with even
greater sensitivity and specificity (e.g., atomic absorption
spectrophotometry), analytical applications of polarography went into a
decline in the 1960's. Classical d.c. polarography is limited to analyte
concentrations in the 10 11M - 1 mM range, but the detection limits are
lowered considerably with the pulse techniques (or the a.c. methods to be
discussed in §6.4), which suppress the interference of the capacitive
232 Voltammetry of Reversible Systems
EIJ2 (lD)a,b
Medium Cu(II) Pb(II) Cd(II) Zn(II)
3.0
(a) i/flA (b)
o
2.5
0.25 flAI
2.0
1.5
1.0
0.29 mM BzSH
0.5
0.056 mM BuSH
0.0
-0.2 -0.4 -0.6 -0.8 -1.0 0 0.1 0.2 0.3 0.4 0.5
EIV C/mM
REFERENCES
(Reference numbers preceded by a letter, e.g. (D4), refer to a book listed in
the Bibliography.)
PROBLEMS
4.1 Show that eqs (4.2) are indeed solutions to the diffusion equations
and that they satisfy the boundary conditions.
4.3 Show that inclusion of the activity coefficient ratio in eq (4.1) leads
to another term in the expression for E 112
E 112 = FJO _.B:L In !2s! + MIn 'Yo
2nF DR nF 'YR
4.4 Use the standard potential for the Fe3 +lFe 2 + couple, diffusion
coefficients, estimated from the molar ionic conductivities of Fe3 +
and Fe 2 +, and activity coefficients, estimated using the Debye-
Huckel limiting law, to compute the half-wave potential for the
reduction of Fe3+ to Fe2+ in 0.1 M KCI04 at 25°C.
4.13 (a) Use eqs (4.29), (4.35) and (4.37) to derive an expression for
semiderivative of the current as a function of time in a linear
sweep voltammetry experiment.
(b) How does the scan rate dependence of the peak signals in
semi-integral and semi-derivative presentation compare with the
peak current in linear sweep voltammetry?
(a) Assuming that the electrode process is reversible and that the
shift in half-wave potential is due to solution iR drop, determine
the "true" value of E1I2.
(b) If the limiting current for ro = 500 rad s-1 is 42.0 1lA, what is the
effective solution resistance?
Problems 243
0.8
ilnA
0.6
0.4
0.2
4.25 The diffusion current constants given in Table 4.1 are uniformly
smaller for polarograms in tartrate buffer solutions than for the
other media listed. Give a qualitative explanation for this
phenomenon.
4.26 Using the data of Figure 4.44, estimate the equilibrium constant
for
Pb 2+ + 3 OR- ~ HPb02 2- + H20
The half-wave potential for the polarographic reduction of Pb 2 + in
acid solution is -0.388 V us. s.c.e.
5.1 INTRODUCTION
Multielectron Processes
In electrode processes, we sometimes write a step such as
o +ne- ~ R
and treat it as if it were an elementary process. In fact, electrode
processes rarely, if ever, involve the concerted transfer of more than one
electron. Gas-phase electron attachment or ionization reactions always
proceed in discrete one-electron steps. Even if the same molecular
orbital is populated or ionized, electron repulsion will cause the two
steps to occur at well separated energies. For electron-transfer reactions
at an electrode-solution interface, solvation effects may bring the two
electron-transfer steps closer together in energy, but only in exceptional
cases would we expect the two steps to coincide.
Electrode processes involving two or more electrons are analogous
to rate laws for gas-phase reactions with overall kinetic orders of three
or more: they provide evidence for mechanisms of two or more
elementary steps. However, a process involving two one-electron steps
250 Mechanisms of Electrode Processes
5.2 SPECTROELECTROCHEMISTRY
1st derivative
of absorption
counter
electrode Luggin probe
to solution connecting
reservoir reference
electrode
t.
Hg pool
cathode
Figure 5.3 Extra muros electro- ESR
lysis cell for ESR spectro- sample
electrochemical experiment. tube
larger chamber above. Although such cells work for many applications,
there are some serious problems. Because of ohmic potential drop along
the tube, most of the current flows at the tip of the working electrode.
The current is usually small, on the order of a few IlA so that the steady-
state concentration of short-lived radicals is low. This problem can be
partially solved by replacing the wire by a Pt gauze electrode in a thin-
layer cell. However, in a long thin cell with the reference electrode
separated from the working electrode, there is very little control of the
working electrode potential; indeed many workers simply eliminate the
counter
electrode '/- Luggin probe
connecting
,.,.' reference
~, electrode
microwwave
~--~r---~ca~ty
(b)
ESR active
volume
.1.·
N
.).
wire
• •
l.~ '~::::
CN CN CN CN
CN ] .2
Second wave: [NchCN
2- CN]
[NchCN 3-
CN CN CN CN
I -C~
+ H+
[NC}yCNj-
CN CN
Third wave:
[NC}yCNJ- [NC}yCN] 2-
CN CN CN CN
Thus the trianion loses cyanide ion to generate a highly basic
carbanion which abstracts a proton from the solvent. The
resulting 1,1,3,3-tetracyanopropenide anion is then reduced to a
dianion radical at the potential of the third polarographic
wave.
Sec. 5.2 Spectroelectrochemistry 263
Figure 5.6 (a) ESR spectrum
a1 ,
obtained on reduction of 1 mM
Me4N [(NChCC(CN)C(CN)2J i!:.J •
in DMF at a Hg cathode. (b)
"Stick spectrum" showing the (a)
l !
I
.I .I .1
positions of the hyperfine T T1 I
components. Reprinted with I 1'
permission from P. H.
Rieger, I. Bernal, W. H.
Reinmuth, and G. K. (b)
Fraenkel, J. Am. Chern. Soc. I I I II II II II I I I
1963,85, 683, copyright 1963, Magnetic Field/Gauss
American Chemical Society.
(a) j
I 20G I
I(
(b) )
If
Magnetic Field 0.2 0.0 -0.2 -0.4
EIV
Figure 5.7 ESR spectra of (a) Figure 5.8 Cyclic voltammogram of
[ArCr(CO)2(PhCCH)]+ and (b) ArCr(CO)2=C=C(SiMe3)2 in CH2CI2
[ArCr(CO)2(Me3SiCCSiMe3)1+ in solution, v = 100 mV s-l. Reprinted
CH2Cl2 solution at room temp- with permiss-ion from N. G.
erature. Reprinted with permiss- Connelly, et a!., J. Chern. Soc.,
ion from N. G. Connelly, et a!., J. Chern. Cornrnun.1992, 1293,
Chern. Soc., Chern. Cornrnun. 1992, copyright 1992 Royal Society of
1293, copyright 1992 Royal Society Chemistry.
of Chemistry.
264 Mechanisms of Electrode Processes
* -,
different stable conformations:
V+
OC.. Cr", Oc.. Cr",
I .... ~ I"~
C ~C(SiMeg)2 C ~C(SiMeg)2
o o
tslow l fast
V,Sill.,_._..;,..+.;:;..e- V,S'M" +
OC_Cr..... C OC.. Cr....... C
I ......J" I .......11'
C C C C
o I o I
SiMeg SiMeg
Sec. 5.2 Spectroelectrochemistry 265
In/TaredSpectroscopy
back plate
salt
o
D
1'efzel gaskets o 0
(a) (d)
where
XR
,
=XR t a nh XR
-XD = XR --'=--'-=--=:...--=--O-----=---=-=7
exp(xDIXR) - exp(-xDIXR)
exp{xDIXR) + exp(-xDIXR}
(5.10)
and XR' ~ XD as Xn/XR ~ o. Using eq (4.45) for XD, we can also write
~~ = 1.61(ijY'6 yk 1 ~ k-l
Substituting eqs (5.8) and (5.9) into eqs (5.7), we have the surface fluxes
J.,=_D(dCY(X») =_~[C*_CS~]
dt x=O 1 +K XD XR'
(~)
dr r .. ro
= -C'·V +i'-}R
0
Substituting these in eqs (5.7), we obtain eqs (5.13), the same expressions
for the surface fluxes as found for r.d.e. electrodes, but with different
definitions for the mass-transport rate constant!
kD = Dlro (5.15)
and the kinetic parameter
A. = XR + ro (5.16)
XR
A. = EL (5.18)
XR'
100
10
100 10
o>/rad s·l
Figure 5.12 The steady· state kinetic parameter A for a range of rate
constants, ki + k.I. as a function of (a) r.d.e. rotation speed ro, (b)
microelectrode radius ro, and (c) d.m.e. drop time td.
Note that the dynamic range for the d.m.e. is rather small, that for
the r.d.e. is larger, and for microdisk electrodes very large indeed. This
is mainly a reflection of the fact that the relevant time scales are
proportional to td, I/o>, and 1/r02 for the d.m.e., r.d.e., and microelectrode,
respectively (see §5.1), and these parameters can be varied over about
factors of 10, 100, and 10 4 . The d.m.e. is sensitive to slower rate processes
than the r.d.e. or a microelectrode, again reflecting the difference in time
scales: up to 10 s for a d.m.e., 1 s for an r.d.e., and 50 ms for a
microelectrode.
274 Mechanisms of Electrode Processes
2
DP
log K Io.,----------~
-2
2.0
absence of P; the ratio of these limiting currents is A from which the rate
constant can be extracted.
For polarography with A» 1, we expect a current enhancement
factor of '/(71[/3) ktd. An analysis specific to the d.m.e. by Birke and
Marzluff (25) gives a rather smaller factor, V1.34 ktd, though the
dependence on rate constant and drop time are the same. Birke and
Marzluff also give equations to correct for departure from pseudo-first-
order conditions.
For a catalytic process with A » 1, the limiting current in a steady-
state voltammogram or polarogram is increased by a factor of A, i.e., by a
factor proportional to the square root of the pseudo-first order rate
constant k'. Since k' = k [P], a means is provided for the analysis of
electroinactive species such as H202 which react with electrode products
such as Fe 3 +. Numerous polarographic methods have been developed
which take advantage of catalytic processes, and most of these should be
extendable to steady-state voltammetry using r.d.e.'s or microelectrodes.
ECE Mechanism: 01 + e- ~ R1
k
R1 ~ R2
02+ e- ~ R2
For an ECE process with E1° < E2° (case IA), 02 is reduced at
potentials where R1 is formed and, if the chemical step is fast enough,
the overall process will appear to involve two electrons. When E1° > E2°
(case IB), 02 is stable at potentials near E1°, and the process is
indistinguishable from an EC mechanism. However, 02 will be reduced
at more negative potentials, so that for E < E2°, case IB is identical to
case IA.
For an ECE process with E1° < E2° (case lIA), R2 is stable at
potentials where R1 is formed and the process is again indistinguishable
from the EC mechani~m. Case IIA can be distinguished from an EC
process by reversal techniques (double potential step
chronoamperometry, cyclic voltammetry, etc.) since the oxidation of R2
would then contribute to the reversal current. When E1° > E2° (case
lIB), 01 is converted to 02 via Rl and R2. The overall process then is
neither an oxidation nor a reduction and, if k is large enough, the
current may be nearly zero.
In all of these cases, the homogeneous electron-transfer
(disproportionation) reaction
01 + R2 ~ 02 + R1
may be important. The effect of this reaction depends on the case and on
the technique involved.
Sec. 5.3 Steady-State Voltammetry and Polarography 283
For the ECE mechanism, we identify R1 and 02 with Z and Y,
respectively, in the general scheme and assume that only 01 has a non-
zero bulk concentration so that C* = O. Using eqs (5.1) and (5.13), the
fluxes are
JOlS = - kD (C0 1* - C0 1 S)
JRl s == - kD AC's
J0 2s = kD(CS + AC'S)
JR2s = kDCR2s
Since JRls = -JO l s and JR2s = -J0 2s, and the surface concentrations of 01
and R1 and of 02 and R2 are related by the Nernst equations for the two
couples, we have four equations in the surface concentrations:
- AC's = (COl * - C01S)
Cs + AC's = - CR2S
COlS/CRlS = 81
C02s/CR2s = 82
Solving for the surface concentrations, we obtain
Go~= ~COI*
A + 91
C s- (A-I) C *
R2 - (1 + 82)(A + 81) 01
i = FAhDC 01 * (21..-1)
1..+81
which corresponds to a wave of shape given by the Heyrovsky-Ilkovic
equation with iL given by eq (5.29) and E 1/2 by eq (5.26). Thus the
apparent number of electrons is (2 - III..) and the wave is shifted to a
more positive potential, as expected from the EC part of the mechanism.
The disproportionation reaction
01 + R2 f:! 02 + R1
has an equilibrium constant
K = exp F(E1°-E2°)
~--=-----=---'-
RT
and, for case lA, K «0. Consider the situation when E "" E1°. The effect
of the R1 ~ 02 reaction is to reduce the R1 concentration at the electrode
surface, resulting in a positive shift of the half-wave potential and an
increase in current (02 is almost entirely reduced to R2 at the electrode).
When disproportionation is allowed, 02 is scavenged by R1 to produce 01
and R2, but since there is very little 02 in solution, the effect on the
current is very small. Voltammograms computed from eq (5.28) are
shown in Figure 5.15a for A = 1, 2, and 20.
Sec. 5.3 Steady-State Voltammetry and Polarography 285
Case IE. When Elo > E2° for the BCE mechanism, two waves are
expected from eq (5.28). The first wave involves one electron and is
indistinguishable from that given by an EC mechanism; the second
wave has El/2 = E2° and n2 = 1- 111.. Although the disproportionation
equilibrium constant is very large, the Rl ~ 02 reaction is only a second-
order perturbation of the position of equilibrium, and the inclusion of
disproportionation has a negligible effect on the current. Computed
voltammograms are shown in Figure 5.15b.
Curves showing napp as a function of the rate constant of the
chemical step have been computed by Marcoux, Adams, and Feldberg
(26) for cases IA and IB for various values of the equilibrium constant K.
2 2
(b)
i/iD
1 1
o o
·0.2 ·0.4 ·0.6 0.2 -0.0 -0.2 -0.4 -0.6
1 1
i/iD Cd)
o o
-0.2 ·0.4 -0.6 0.2 -0.0 -0.2 -0.4 -0.6
ElY ElY
Figure 5.15 Computed steady-state voltammograms for the ECE
mechanisms for A = 1, 2, and 20. (a) Case IA; (b) case IB; (c) case IIA;
and (d) case lIB (the solid and dotted lines correspond, respectively, to
infinitely slow and infinitely fast disproportionation equilibria). For
cases IA and IIA, El ° = -0.4 V, E2° = 0.0 V, and for Cases IB and lIB, E10
= 0.0 V, E2° =-0.4 V.
286 Mechanisms of Electrode Processes
Case llA. When E1° < E2° for the ECE mechanism, 81 » 82, and
the second term of eq (5.30) will be close to 0 whenever the first term is
significantly greater than zero. Thus case IIA is really just an EC
process as far as steady-state voltammetry is concerned. A one-electron
wave is expected, which shifts to more positive potentials as the rate of
the chemical step increases. The effects of disproportionation are
completely negligible. Voltammograms, computed using eq (5.30) are
shown in Figure 5.I5c for A. = 1, 2, and 20.
Case lIB. When E 1 ° > E2° for the ECE mechanism and the
potential is in the vicinity of E1°, we have 81 «82. We then expect a
wave with n1 = I/A. and half-wave potential shifted as expected for an EC
process. A second wave is expected with E1I2 = E2° and n2 = 1 - I/A. (the
total limiting current corresponds to one electron). When E1° > E > E2°,
the effects of disproportionation are very significant. Here 01 is reduced
irreversibly to R1, which reacts to form R2. At the electrode, R2 is
oxidized and the net current is diminished. Disproportionation
increases the efficiency of this process since any R2 which escapes from
the electrode is scavenged by 01, leading to a further decrease in the
current. Steady-state voltammograms, computed using eq (5.30) and by a
digital simulation method (assuming that the disproportionation
reaction is infinitely fast), are shown in Figure 5.I5d.
Having examined the CE, EC, EC', and ECE mechanisms from the
point of view of steady-state voltammetry, we are now prepared to see
how electron-transfer processes influenced by homogeneous chemical
reactions behave when studied by chronoamperometry-current-time
curves at constant potential-and chronopotentiometry-potential-time
curves at constant current.
1 Since we assume the electron transfer process is nernstian and therefore infinitely
fast, the current should be infinite at zero time due to the small amount of 0 initially
present; for small K, this initial transient is of negligible importance after a few
milliseconds.
288 Mechanisms of Electrode Processes
4
0 (b)
a ................ O&.
0.1 kt
.......... 0.2
------- 0.4 3
t =400 ms
0 t =300 ms
IJ t =200 ms
o 1 2 3 o ro 100 100 200
tlms
Figure 5.17 (a) Theoretical working curves for double potential step
chronoamperometry applied to the EC mechanism. The curves correspond to various
values of tit, where t is the time of the second potential step and t is the current
measurement time after each potential step. (b) Kinetic plot for the rearrangement of
hydrazobenzene followed by double potential step chronoamperometry: 1.0 mM
azobenzene in 1.98 M HCl04 in 50% aqueous ethanol, t = 200, 300, and 400 ms.
Reprinted with permission from W. M. Schwartz and I. Shain, J. Phys. Chern. 1965,
69, 30, copyright 1965 American Chemical Society.
Sec. 5.4 Chronoamperometry and Chronopotentiometry 289
Q-NH-NH-o k H2N-O-ONH2
Hydrazobenzene is oxidized at about the same potential where
azobenzene is reduced, but benzidine is electroinactive in this
potential range. Thus the double potential step method could
be used to monitor the amount of hydrazobenzene remaining at
various times after formation. Anodic-to-cathodic current
ratios were measured for tl't = 0.1, 0.2, 0.3, 0.4, and 0.5 for
various values of t, the switching time. Values of k1: were read
off the working curves, multiplied by th to obtain kt, and kt
plotted us. t as shown in Figure 5.17b for data obtained for 1: =
200, 300 and 400 ms, [azobenzene] = 1 mM in 1.98 M HCI04 in
50% aqueous ethanol solutions. The slope of this plot is the rate
constant k = 23 s-l. Similar plots were obtained for other acid
concentrations with rate constants ranging from 0.6 to 90 s-l.
290 Mechanisms of Electrode Processes
0.3
'tIt
0.2
0.1
Figure 5.18 Working curves for
determination of rate constants
from current-reversal chronopot-
entiometric measure- ments for
the EC reaction scheme. 1: is the 0.0
transition time after current o 1 2 3 4 5
reversal at time t. kt
Sec. 5.4 Chronoamperometry and Chronopotentiometry 291
40
tis 't 2
Figure 5.19 Current-reversal
3J !l
T
chronopotentiometric curves for the
oxidation of p-aminophenol (1 mM in
0.1 M H2S04 solution) and the ID
hydrolysis of p-benzoQuinone-imine; t
is the time of anodic electrolysis, 'tl t
1
and 't2 are the transition times 10
corresponding to reduction of the imine
and Quinone. Reprinted with permiss- 0
ion from A. C. Testa and W. H.
Reinmuth, Anal. Chern. 1960,32, 1512,
copyright 1960 American Chemical 0.8 0.6 0.4 0.2 0.0 -0.2
Society. EIV
292 Mechanisms of Electrode Processes
Catalytic Reactions
For the EC' mechanism
o +ne- ~ R
k
R+P ~ O+Q
the current is enhanced by the feedback of R to 0 and
chronoamperometry provides a rather direct measure of the rate
constant k.
When the potential is stepped to a sufficiently negative value that
CoCO) = 0, the limiting current can be shown to be (29)
iL = nFACO* [YDohtt exp(-k't) + VDok' erfYk't] (5.34)
or
iLliD = exp(-k't) + Yd't erfYk't (5.35)
where k' = kCp and iD is the unperturbed diffusion-limited current, eq
(4.7). For short times, the error function is near zero, the exponential
close to one, and iL "" iD. For long times, k't » 1, the exponential
approaches 0, the error function goes to 1, and eq (5.34) becomes
4
k' = 10
i/iD
...,
W 3
'S 3
;::j
....>.
2
...,....CIl
..... 2
..//~/
..0
....
CIl
~
k' =1
1 1
k' = 0.1
k' =0 /
0 o ~..~.~~~~~~~~~~~
0 1 2 3 0.0 0.5 1.0 1.5 2.0 2.5
tis (kt)1J2
(5.36)
independent of time. The predicted behavior is shown in Figure 5.20 for
several values of k'. The current should decay at very long times since P
is consumed and pseudo-first-order conditions no longer prevail;
however, if Cp* » Co* and semi-infinite planar diffusion applies, the
current may be constant for quite a long time. Thus if the rate constant
is to be extracted, it is best to measure the current in the presence and
absence of P and to compute iLliD. In this way, k' = kCp can be
determined without havi~to know the diffusion coefficient or electrode
area. A plot ofiLiiD us. Yk't, shown in Figure 5.21, serves as a working
curve for such an experiment. Figure 5.21 also serves as a reaction zone
diagram with the pure diffusion (DP) zone defined by k't < 0.05, the pure
kinetic (KP) zone by k't > 1.5, and the intermediate kinetic (Kl) zone by
0.05 < k't < 1.5.
with E1° < E2° (case IIA), a potential step to E < E1°, E2° is expected to
involve only one electron regardless of the rate of the chemical step and
so should be independent of the existence of the homogeneous electron-
transfer process. In case IIB (E1° > E2°), on the other hand, the Alberts-
Shain theory predicts
napp = exp(-kt) (5.38)
but Feldberg and Jeftic (32) find that for kt > 1, napp is significantly
smaller than predicted by eq (5.38) when disproportionation is
considered and even becomes negative for kt > 2.5. Early in the
experiment, the current is due mostly to the reduction of 01 to R1. Later
when 01 is polarized at the electrode, the R2 oxidation dominates and the
current becomes negative if the conversion of R1 to R2 is fast enough.
Some computed working curves are shown in Figure 5.22.
napp
0.5
O=O=NH
-
~
+2H+
H0-o-/, NH2
\\ //
The corrected datal are plotted in Figure 5.23 as iIFAC*D1I2 us.
t- 1/2 , along with curves, computed using eq (5.37) for k = 0.4, 0.6,
and 0.8 s-l. The data fit the curve for k = 0.6 s-l reasonably well;
Alberts and Shain report k = 0.59 ± 0.07 s-l. The small
discrepancy between theory and experiment is probably due to
neglect of the homogeneous disproportionation reaction.
3
, ,,
,,,
,,
,
,,
"
,
,,
,,
Figure 5.23 Chronoampero- , ,,
metric data for the , ,,
reduction of p-nitroso- , ,,
phenol in 20% aqueous 1 ,,
,,
ethanol solution, pH 4.8.
, ,,
Reprinted with permission
, ,,
from G. S. Alberts and I. ,,
Shain, Anal. Chern. 1963, ,,
35, 1859, copyright 1963 o
American Chemical 0.0 0.5 1.0 1.5 2.0
Society
(t/s)-1I2
1 The experiment employed a hanging mercury drop electrode; the theory for a
spherical electrode is considerably more complex than that for a planar electrode,
although the qualitative effect is similar to that described by eq (5.37). The data
plotted in Figure 5.24 have been corrected to correspond to a planar electrode.
296 Mechanisms of Electrode Processes
1.5
~
rn
.§
~ 1.0
C
CIS 0.5
....,
I-<
:EI-<
~ 0.0
~
Figure 5.24 Cyclic
voltammograms at :S -0.5
constant scan rate for
the CE reaction scheme
for K = 0.2 and A = 0.25 -1.0
(solid line), 2.5 0.2 0.1 -0.0 -0.1
(dashes), and 25 (dots).
(E -E1/2)V
0.6
0.4
0.2
-2.0 -1.5 -1.0 -0.5 0.0 0.5 -1 o 1 2
log k't log A.
Figure 5.26 (a) Ratio of anodic to cathodic peak currents for an EC
process as a function of kt. (b) Shift. of the cathodic peak potential as a
function of A for an EC process. Reproduced with permission from R. S.
Nicholson and I. Shain, Anal. Chern. 1964,36, 706, copyright 1964
American Chemical Society.
the rate constant. A working curve, obtained from the work of Nicholson
and Shain (33), is shown in Figure 5.26a (the parameter 't is the time
required to scan the potential from the cathodic peak and the switching
point).
As we might have expected from the results for steady-state
voltammetry, the cathodic peak shifts to more positive potentials for very
fast following chemical steps. If the unperturbed peak potential is
known, the peak potential shift can be used to estimate the rate constant.
A working curve, obtained from the work of Nicholson and Shain (33), is
shown in Figure 5.26b.
When the following reaction is second-order and involves an
electroinactive species, an interesting experimental variation is possible.
Consider the reaction scheme
O+e- ~ R
k
R+X ~ Y
When k[X] is large, an anodic shift in the peak potential is expected.
However, if less than a stoichiometric quantity of X is available, the
reaction proceeds until X is exhausted, after which R remains
unreacted. If k is sufficiently large, a shifted peak will precede the
unperturbed current peak. The separation of this prepeak and the
reversible peak then gives an accurate measure of the rate constant. In
§5.5 Cyclic Yoltammetry 299
order to resolve the prepeak, llEp must be more than about 80 mY. The
working curve of Figure 5.26b then gives A. ~ 250 or k[X]/v ~ 104 y-1. If[X]
'" 1 mM and v = 0.1 Y s-l, we require k ~ 10 6 L mol-Is-I. The method thus
is restricted to very fast rates, but, because the reversible peak provides
an internal standard for the measurement of the kinetic shift, the
results can be quite precise. Although the working curve of Figure 5.26b
gives a qualitative indication of the shift, in practice digital simulation
using conditions close to the actual experiment should be used to refine
the determination.
This effect was first discussed by Jensen and Parker (34) in the
context of conventional cyclic voltammetry, but a considerable
improvement in accuracy is obtained by using first-derivative (or
semiderivative) presentation. Because the prepeak and reversible peak
are better resolved in a first-derivative voltammogram, peak separations
can be measured much more accurately.
Example 5.10 Parker and Tilset (35) have studied the reaction
of 9-phenylanthracene cation radicals (PA+) with 4-
methylpyridine (Nu) using derivative linear sweep
voltammetry. Sample results are shown in Figure 5.27 for
[Nu]I[PA] concentration ratios of 0, V4 and V2.
Catalytic Reactions
Linear scan voltammograms at a stationary electrode are closely
analogous to potential step chronoamperograms for catalytic processes.
The kinetic zone parameter is again
A=RTk' (5.41)
nFv
where k' is the pseudo-first order rate constant. In the pure kinetic
zone, defined by A > 1, the cyclic voltammogram reduces to a steady-state
wave with limiting current given by eq (5.35), independent of scan rate,
with E p/2 = E 112 equation. The reverse scan in cyclic experiments
virtually retraces the current-potential curve of the forWard scan. When
the homogeneous rate becomes slower (or the scan faster) cathodic and
anodic peaks develop and the half-peak potentials shift toward their
unperturbed values (E1I2 ± 28.5/n mV at 25°C). Some representative
curves are shown in Figure 5.28.
Since the OI/RI and 02/R2 couples are observed separately, cyclic
voltammetry is a particularly powerful technique for the study of ECE
processes,
ECE Mechanism:
ECE Mechanism:
The relative size of the reduction (or oxidation) peaks thus usually can
give an indication of the magnitude of the rate constant for the chemical
step. Examples of cyclic voltammograms for several values of the rate
(a) (c)
0.2 -0.0 -0.2 -0.4 -0.6 0.2 -0.0 -0.2 -0.4 -0.6
EIV EIV
Figure 5.29 Cyclic voltammograms computed by digital simulation for
case IA of the ECE mechanism. The rate parameter A = 0 (a), 0.025 (b),
0.25 (c), and 2.5 (d); E10 =-0.4 V, E2° = 0.0 V.
302 Mechanisms of Electrode Processes
(c)
01 + R2 +Z 02 + Rl
has K« 1 in case lA, so that 02 and Rl tend to scavenge one another.
The principal effect of disproportionation on case IA cyclic
voltammograms is to suppress the curve crossing seen for intermediate
rates (32,36). Amatore, Pinson, Saveant, and Thiebault (37) have pointed
out that disproportionation can have the opposite effect when the
chemical step is exceedingly fast. If K is not too small and Rl is very
short-lived, the disproportionation reaction may be pulled to the right
(uphill energetically) to supply more 02 and Rl (which is quickly
converted to 02). The effect is again to give a cathodic current in the
interpeak region.
In case IB (the ECE mechanism with El° > E2°, Figure 5.30), the
first negative-going scan shows cathodic peaks at both El° and E2°. The
height of the second peak approaches that of the first peak as the
reaction rate increases. The reverse scan shows an R2 oxidation peak
comparable in size to the 02 reduction; if the reaction is fast, little Rl
remains to be oxidized, so that the Rl oxidation peak may be missing.
Disproportionation has little effect on the first negative-going scan, but
the Rl oxidation peak is somewhat smaller on the reverse scan. When
the Rl ~ 02 conversion is not very fast (Figure 5.30b), the
(a)
0.2 -0.0 -0.2 -0.4 -0.6 0.2 -0.0 -0.2 -0.4 -C.6
EIV EIV
Figure 5.31 Cyclic voltammograms computed by digital simulation for
case I1A of the ECE mechanism. The rate parameters and standard
potentials are as described for Figure 5.29.
304 Mechanisms of Electrode Processes
0.2 -0.0 -0.2 -0.4 -0.6 0.2 -0.0 -0.2 -0.4 -0.6
EIV EIV
Figure 5.32 Cyclic voltammograms computed by digital simulation for
case lIB of the ECE mechanism. The rate parameters and standard
potentials are as described for Figure 5.30. The solid lines were
computed neglecting disproportionation. The dotted lines in (b) - (d) show
the effect of infinitely fast disproportionation.
§5.5 Cyclic Voltammetry 305
Example 5.11 Hershberger, Klingler, and Kochi (38) used
cyclic voltammetry to study the oxidation of (Tl 5 -
CHgC5H4)Mn(CO)2L (Mn-L). Cyclic voltammograms of the
derivatives with L = NCCHg and PPhg (traces a and d,
respectively, in Figure 5.33) are uncomplicated and apparently
reversible with peak current ratios of 1.0 and normal peak
potential separations. However, addition of PPhg to a solution
of Mn-NCCHg results in a dramatic change (traces band c of
Figure 5.33). On the first positive-going scan, the oxidation
peak due to the substrate decreases in size and an oxidation
peak due to the product, Mn-PPhg, appears. On the reverse
scan, the product cation is reduced but the cathodic peak
expected for the substrate cation is completely absent.
Comparison of traces b and c of Figure 5.33 with traces bod of
Figure 5.32 suggests that these results can be understood as an
ECE process-the analog of case IIB, making appropriate
allowance for the fact that the first step is an oxidation rather
than a reduction. The mechanism can be written as follows:
Mn-NCCHa =-e-
(Mn-NCCHa"t
~ V;;l PPha
~rCHaCN
Mn-PPha =-e [Mn-PPhal+
0.52 V
As we have seen in Figure 5.32, the shape of a cyclic
voltammogram for such a system is very sensitive to the rate
constant k 1 and somewhat dependent on k 2 as well.
Hershberger, et aI., carried out a systematic investigation
using digital simulation techniques and were able to establish
the rate constants: kl = (1.3 ± 0.2) x 104 M-1s-l, k2 > 104 M-1s-l.
(a)
!\...
I I
(a)
n:.J~
~~CO
~lrL
+ + e-
[Mn-L] :;::::::: Mn-L
-0.7 V
308 Mechanisms of Electrode Processes
REFERENCES
PROBLEMS
0.55
jtA m- 2
0.50
0.45
0.40
5.6 Show that the steady-state current for an EE process where both
steps are assumed to be nernstian, i.e.,
is given by
i = FACo*kD ( 2 + 62 )
1 + 62 + 6162
where 61 and 62 are the nernstian concentration ratios for the
first and second couples.
5.7 (a) Given the result of Problem 5.6, derive an expression for (iL -
i)/i.
(b) If E2° > E1°, a single wave is expected. Compute the half-wave
potential.
(c) Under what circumstances does the answer to part (a) reduce
to the Heyrovsky-Ilkovic equation for a two-electron wave?
(d) Compute E1I4 - E3/4 for E2° - E1° = 0, 50, 100, and 200 mY.
What value would be expected for a two-electron wave according to
the Tomes criterion for reversibility?
312 Mechanisms of Electrode Processes
2.0
1.5
$ 1.0
'§ 0.5
.~... -0.5
electrode process with
scan rate v = 400 mV s-l.
The dotted curve shows a
cyclic voltammogram of -1.0
the same system
uncomplicated by the -1.5
following chemical -0.2 -0.4 -0.6 -0.8
reaction.
EIV
-40
--_ ... --' -- ,
, , 4
\
-20 2.0
;>
S 0 \
0'
'--'
....1:>..
~
rn
.....
+'
§
2
,
' .
::::.::. \
0 \ >.
~
\
, 1.5 I-<
CIS
20
\
,, I-<
,;
I ..... 0
+' ,
!S,I:>.. ,..c ........ .,-- .. -.............. :
-'..,
\ ,,~
,
\
I-<
,,
\
."
CIS ,
\
\ :;::; ,,
40 \
\
\
\
1.0 -2 '' '
,, ''
'
.
','
60
-2 -1 0 1 2 0.0 -0.5 -1.0 -1.5
log A EIV
Figure 5.37 Theoretical working Figure 5.38 Cyclic voltammo-
curves for cyclic voltammograms grams for an ECE process. The
for case IA of the ECE mechanism: potential scan rate v is 100 mV s-1
peak potential shift (dashed curve) (solid curves), 500 mV s-1 (dashed
and peak current (solid curve) vs. curves), and 2.5 V s-1 (dotted
the kinetic parameter I.. = RTkIFv. curves).
-~Ga:j:
ka =Zexp - - - (6.2b)
RT
where the ~Gt's are free energies of activation. The pre-exponential
factors Z can be estimated theoretically (1) but we will be content with
noting that the Z's in eqs (6.2a) and (6.2b) must be equal to satisfy the
principle of microscopic reversibility; otherwise, we will treat Z simply
as an empirical parameter with units of m s-l.
Consider now the special case of equilibrium at the electrode. The
net current must be zero and the surface and bulk concentrations
should be equal. Substituting eqs (6.2) into eq (6.1), taking logs and
rearranging, we have
In Co(O,t) = In ka,o = ~Ge:j: - ~Ga:j: (6.3)
CR(O,t) ke,o RT
where ke,o and ka,o are the rate constants at zero current. The ratio of
the reactant and product concentrations at equilibrium is given by the
Nernst equation,
In Co* = F(Ee - EO) (6.4)
CR* RT
where Ee and EO are the equilibrium and standard half-cell potentials,
respectively. Combining eqs (6.3) and (6.4), we have
~Get - ~Gat = F(Ee - EO)
This expression is a statement of a familiar idea from chemical
kinetics: as shown graphically in Figure 6.1, the difference between the
forward and reverse activation free energies is equal to the standard free
energy change for the reaction. The rates of the cathodic and anodic
processes depend on the electrode potential. By varying the potential, we
can change the free energies of 0 and R at the electrode surface. Thus
the activation free energies must depend on the potential. The details of
this dependence might be rather complicated, but for the moment we
§6.1 Kinetics of Electron Transfer 317
T
T
!J.Gc :t:
!J.Ga :t:
k k ~F(Ee - EO)
a,O: 0 exp '------R'---T-- (6.6b)
where
-!J.G:j:
ko: Z exp __0_ (6.7)
RT
When a net current flows through the electrochemical cell, the cell
is not at equilibrium. The deviation of the half-cell potential from the
equilibrium value is called the overpotential or overvoltage, 11.
E: Ee + 11 (6.8)
We assume that the form of eqs (6.5) is retained under nonequilibrium
conditions so that the cathodic activation free energy changes by an
318 Electron-Transfer Kinetics
amount aFTl and the anodic activation free energy by an amount -~FTl.
The rate constants then differ from the equilibrium values according to
the relations
(6.9a)
~FTl
ka = ka,o exp RT (6.9b)
l. = lO - a.F'tl
. [CoCO,t) exp - ~FTl]
- - CR(O,t) exp - - (6.10)
Co* RT CR* RT
where io, the exchange current, is equal to the cathodic current (and to
the negative of the anodic current) at equilibrium:
F'AkJ' * - aF(Ee
.
lO .
= le,e = uuO exp RT- EO) (6
.11a )
If transport is really fast so that CO(O,t) = CO* and CR(O,t) = CR*, then eq
(6.10) reduces to
. . [exp~-exp
l=lO
- a.F'tl ~FTl
RT
] (6.12)
iii 0
10
o
Figure 6.2 Current-potential
curve predicted by the
Butler-Volmer equation for
a = 0.4. The dashed lines -10
represent the anodic and
cathodic components of the
current. The dotted line
represents the linear -20
(ohmic) region for small 200 100 0 -100 -200
overpoten tials.
11/mV
1 Since a and ~ govern the symmetry of a current-potential curve, some authors refer
to transfer coefficients as symmetry factors.
320 Electron-Transfer Kinetics
Ca) (b)
Cc)
Go* = Wo
Similarly, there will be work Wr required to bring R from bulk solution to
the electrode and affect any gross changes required before electron
transfer. With the zero of free energy defined as above, the bulk solution
free energy of R is F(E - EO). Thus
GR* = Wr + F(E _EO)
Substituting GO*, GR*, and 'A in eq (6.13), we have
~Gct = Gt - Go = Gt
Double-Layer Effects
- Z Fc'P
CR(a) = CR* exp ~T a
where Zo and Zr are the charges on 0 and Rand c'Pa is the potential at x =
a (relative to the bulk solution). If <l>a is significantly different from zero,
then we should use CoCa) and CR(a) in eqs (6.11) for the exchange
current rather than the bulk concentrations Co * and C R * .
Furthermore, the potential difference contributing to the activation free
energy needs to be corrected by subtracting c'P a from Ee. Thus eqs (6.11)
become
- zaFc'P - aF\E - EO - c'P )
io = FAkcf;o* exp RT a exp eRT a
.
lO =
F' "k J l
.n. uvR
* exp - zrF<l>a
RT exp
pF(Ee - EO - c'Pa}
RT
324 Electron-Transfer Kinetics
The apparent exchange current given by eqs (6.11) then is related to the
true exchange current by
.\
(lO/app= . (a - zo)FcI>a (6 18 )
loexp RT . a
. ) loexp
(lOapp= . - h~ +RT
zr}FcI>a
(6. 18b)
or, at 25°C,
log i = log io - 16.90 001
Equation (6.21) is known as the Tafel equation. This logarithmic
current-potential relationship was discovered empirically by Tafel in
1905 (8), some years before the theory of electrode kinetics was developed.
The Tafel equation suggests the means by which the exchange
current and the transfer coefficient may be determined. If, for an
equilibrium mixture of 0 and R, the current is measured as a function
of overpotential and then plotted as log i vs.", a linear region should be
found. Extrapolation of the linear portions of the plot to zero
overpotential yields the log of the exchange current as the intercept; the
slopes should be -16.90 a. and 16.90 13. Such a plot is shown in Figure 6.5.
The Tafel equation and Figure 6.5 suggest that the current
increases exponentially with increasing overpotential. There must be a
point at which the current becomes limited by the rate of transport and
log ius." plots begin to flatten out. The point at which this happens
326 Electron-Transfer Kinetics
2.0
1.5
1.0
·2
........
0.5
-
'E0D
0.0
Thus measurement of Ret (and thus io) for a series of solutions with
constant Cot and variable CRt (and thus variable Ee) allows the
evaluation of u.
Multistep Mechanisms
Thus far we have assumed that the electrode process is simply the
addition of an electron to a single molecule of 0 to produce a single
molecule of R. Consider a process with the stoichiometric half-cell
reaction
Vo 0 + n e- fZ Vr R
The concerted addition of two or more electrons in a single step is highly
unlikely (see further discussion of this point in §5.1). In general, we
expect that any electrode process involving two or more electrons
necessarily involves two or more elementary steps. If one of these steps
is rate limiting, the rate laws should have the relatively simple forms:
anodic rate = kaCRnorCOno.
cathodic rate = kcCRn.,.CO ••
where the n's are the orders of the reactions. Equations (6.11) then are
-6 I I I -8.5
(a)
x
-7
~~
~ ~ -9.0
S S
~-8
::::>
XO~o ~
-
b:O ~
0
_ x _ ..s -9.5
t
-9
<>[Mn2+] = 0.01 M
x [Mn2+] = 0.001 M
-10
, , , ,<>
-10.0
0.6 0.8 1.0 1.2 1.4 1.6 1.40 1.45 1.50 1.55 1.60
EIV Ee /V
Figure 6.6 (a) Tafel plot (cathodic branch) for the Mn 3+/Mn 2+ couple in
7.5 M H2S04 solution at 25°C. The upper points correspond to [Mn 3+] =
0.01 M, the lower ~oints to [Mn 3+] =0.001 M. (b) Exchange current density
for the Mn 3 +/Mn + couple in 7.5 M H2S04 solution at 25°C. Reprinted
with permission from K. J. Vetter and G. Manecke, Z. phys. Chern. 1950,
195,270, copyright 1950 R. Oldenbourg Verlag.
or
(aIn
aEe
co) CR
= ~
vJlT
Combining the partial derivatives, we obtain
(adE
In iO) = L (nnco _ a) (6.25a)
Ce R
RT VO
(aaEe
In iO) =_L(nncr + <X)
Co RT Vr
(6.25c)
or
(aaE
In iO) =_L(nnar _ p) (6.25d)
e c 0
RT Vr
0.16
~
t...
0.12
0.01 M, [13-] variable
(diamonds); [13 -] = f5
0.0086 M, [1-] variable
(squares). Reproduced 0.10
with permission from
K. J. Vetter, Z. phys.
Chern. 1952,199, 285, noo n~ nw nffi nw nM
copyright 1952 R.
Olden bourg Verlag. Eei V
13- ~ 12 + 1-
12 ~ 21
2 (I + e- ~ I-)
where the last step is rate limiting. The small transfer
coefficient is not surprising; eq (6.17a) suggests that a. should
be less than 112 when a lot of work is required to get 0 ready for
electron transfer. The surface potential on platinum is
probably positive in this experiment, so that eq (6.20) would also
suggest a decrease in a. from 112.
1+ e- ~ R E2°
The standard potential for the overall two-electron process is
E12° = (E10 + E2°)/2
and the Nernst equation for the overall process gives the equilibrium
RIO concentration ratio
CR
-=exp -2F(Ee- E 12°) = ee
12
Co RT
where
which has the form of the Butler-Volmer equation, but with aapp = a,
l3app = 1 + 13. We thus have the interesting result: For a two-electron
process where the first step is rate limiting, the apparent cathodic and
anodic transfer coefficients are expected to be approximately 112 and 3/2,
respectively.l Notice that the exchange current in this case is a
complicated function of the electron-transfer rate constant ko, the
transfer coefficient a, and the standard potentials for the two steps, E 1°
and E2°. The individual standard potentials are usually unknown, so
that ko cannot be determined directly.
The situation would be just reversed if the second step were rate
limiting; the apparent cathodic and anodic transfer coefficients would be
3/2 and 1/2, respectively.
Example 6.4 Bockris, Drazic, and Despic (11) studied the rate
of anodic dissolution of iron in 0.5 M FeS04, 0.5 M Na2S04.
Figure 6.8 shows Tafel plots for the process
Fe(s) -+ Fe2+ + 2 e-
at pH 2.0, 3.1, and 4.0. The slopes of the three lines are
approximately equal and give
1 Some authors use the term "transfer coefficient" to refer to these empirical
"apparent transfer coefficients," reserving the term "symmetry factor" for the
transfer coefficient of an elementary process.
334 Electron-Transfer Kinetics
or
k [H+] - aF(E - EO) k C PF(E - EO)
o exp RT '" 0 H exp RT
Thus the surface concentration is
C [H+] F(E _EO)
H'" exp RT
and the net current is given by
. 2F'Ak [H+'2 -2F(E-EO)
t = .l'1 C j exp RT
§6.2 Current-Overpotential Curves 337
or
.. - 2Frt
l=loexp RT
Thus the apparent cathodic transfer coefficient for this mechanism is
expected to be about 2, consistent with the results for Pd.
A strong M-H bond is expected to make step (i) more favorable but
hinder step (iia) or (iib). Conversely, a weak M-H bond would make step
(i) slow and (iia) or (iib) fast. Thus we might expect a correlation
between exchange current density and M-H bond strength which would
distinguish between the two kinds of mechanism. The exchange
current density for several metals is plotted as a function of the
estimated M-H bond strength in Figure 6.9. The exchange current
density increases with increasing bond strength in the series Tl - Co,
presumably because step (i) of the mechanism above is getting faster, but
is still rate limiting (i.e., slower than steps iia or iib). In the series Pt -
Ta, the exchange current density decreases with increasing bond
strength, presumably because step (iia) or (iib) is rate limiting.
0
C'I
S
Figure 6.9 Exchange
<
"0 -2
-
current density for H2 :.:>
b.O
evolution from aqueous 0
solutions for various
electrode materials as a -4
function of estimated M-
H bond strength. Repro-
duced with permission
from S. Trasatti, J. Elect-
roanal. Chern. 1972,39, 100 150 200 250 300 350
163, copyright 1972 Else-
vier Science Publishers. M-H bond enthalpy/kJ mol- 1
338 Electron-Transfer Kinetics
Steady-State Voltammetry
1.0
ilin
0.8
0.6
0.4
0.12
~ 0.08
~
~
Figure 6.11 Kinetic I
diagram showing the o
shift in E 112 as a ~ 0.04
function of the width
of the wave; solid
curves correspond to
constant IX and
variable log kolkn; -0.00
dashed curves to
constant log kolkn 0.05 0.10 0.15 0.20
and variable IX. (E 114 -E3/4 )V
Thus if the reversible half-wave potential is known and El/2 and El/4-
E3/4 are measured, the corresponding point on the diagram determines
the kinetic parameters.
For r.d.e. voltammograms, kn = Dlxn, where xn is a function of
rotation frequency and is given by eq (4.45); thus
kn = 0.62 D2/3y-1I6CJ)1I2
For D = 10-9 m 2 s- 1 , y = 106 m 2 s- 1 , and CJ) = 10 - 1000 rad s-l, kn ranges from
2.0 x 10- 5 to 2.0 x 10-4 m s-l. Thus the experimentally accessible quasi-
reversible region has 2 x 10-6 < ko < 2 x 10-3 m s-1.
For steady-state voltammograms using a microdisk electrode, kn is
given by eq (4.51)
kn = 4Dlnro
where ro is the radius of the microdisk. For D = 10-9 m 2 s- 1 and r ranging
from 0.25 11m to 25 Ilffi, we have 5 x 10-3 m s-l < kD < 5 x 10-5 m s-l. The
experimentally accessible quasi-reversible region then has 5 x 10-6 < ko <
5 x 10-2 m s-l. Note that the range extends to larger ko than with an r.d.e.
Furthermore, since experiments are much less troubled by ohmic
potential drop, microelectrode voltammetry is in principle the preferred
technique for measurement of electron-transfer kinetics. In practice,
however, electrode surface imperfections-poorly defined effective
electrode radius-lead to a potentially large systematic error.
In an experiment, a series of voltammograms is recorded using a
range of microelectrodes of varying sizes. If ko is near the upper end of
342 Electron-Transfer Kinetics
(6.34)
where
§6.3 Electron-Transfer Rates from Voltammetry 343
i = 2FAknCo*
1 + 9192
or, with iD = 2FAkDCO*, we have
in-i
- . - = 99
12
l
which, on taking logs, is the Heyrovsky-Ilkovic equation with n = 2 and
EV2 = (E10 + E2°)/2.
Either of the two electron-transfer steps could be rate limiting.
Consider first the case where step one is slow and step two fast. For 92
« 1 and kdk2c « 1, eq (6.34) becomes
i = __ 2F_:A_"=n_C-",o,-*_
1 + 9192 + knlk1c
which is very similar to eq (6.30) except that the limiting current
corresponds to two electrons. There is an important difference, however.
In the reversible case (kD1klc « 1), the Tomes criterion gives
E V4 - E3/4 = 56.51n m V
or 28.2 mV for a two-electron wave at 25°C. But when the process is
completely irreversible, we have
E V4 - E3/4 = 56.5/a m V
or about 113 m V if a = 1/2. Thus the increase in width with decreasing
electron-transfer rate constant (or increasing mass transport rate) is
rather more dramatic than in the one-electron case. Computed waves
for E2° - E10 = 0.2 V and a = 112 are shown in Figure 6.13a for several
values of kDlko.
When step 2 is rate limiting, the wave shape is sensitive to the
difference in potentials, E2° - El ° and the electron-transfer rate. If the
separation is large compared with the kinetic shift, a single
symmetrical two-electron wave is expected with a width of
EV4 - E3/4 = 56.5/(1 + a) mV
This is analogous to the two-electron process discussed in §6.2, where we
found an apparent transfer coefficient of 3/2.
When the kinetic shift is larger than the separation of standard
potentials, a rather unsymmetrical wave results. If the second electron-
transfer step is intrinsically very slow, the current from the first
electron transfer may be nearly limiting before the overpotential is big
enough to provide a significant rate for the second step. The transition
from a reversible two-electron wave to this unsymmetrical situation is
shown in Figure 6.13b.
§6.3 Electron-Transfer Rates from Voltammetry 345
2.0
2ilin
1.5
1.0
0.5
0.0
0.0 -0.2 -0.4 0.0 -0.2 -0.4
(~ - ~1/2(rev)1V (~ - ~1/2(rev)1V
drop time), the ratio kD1ko is affected. If the process is reversible, the
shape of the curve is unaffected and the apparent half-wave potential is
constant, but variations in El/2 will be found for quasi-reversible and
irreversible waves.
(6.38)
The current is obtained by substituting the surface concentrations, eqs
(6.36), into eq (6.35), obtaining
1.0
F(A-.Jt)
0.8 ----
...,
Ul
3
·S;::l
0.6 >.
.......,e
I-<
2
0.4
1::til
'-'
1
0.2
....................... --_ ............. -- --- --- _.. ..
0.0
o 1 2 3 o 1 2 3 4 5
A-.Jt tIs
Figure 6.14 The function f('J..:Vt) Figure 6.15 Current-time curves
VS. 'J..Vt for potential step chronoamper-
ometry with a = 0.5, 11 = -0.1 V,
DO = 10. 9 m 2 s· 1 , and ko = 00
i(O) = FAkeCO*
so that the current is initially limited by the rate of electron transfer.
When we make the overpotential sufficiently negative that ke »ka, A'"
kelW and eq (6.39) becomes
i = FACO*VDlrtt /tAVi) (6.40)
For a sufficiently negative overpotential or at long times, AVi » 1, /tA vi)
'" 1, and eq (6.40) reduces to the Cottrell equation, eq (4.7); the current is
then diffusion controlled.
Notice that eq (6.40) has the same form as eq (5.31) for the current
limited by a homogeneous chemical step preceding electron transfer. As
we remarked in §5.4, an electron transfer process preceded by a slow
reaction behaves very much as if the electron transfer step itself were
slow. However, we now see that the two phenomena are easily
distinguished: the current for a CE process is independent of
overpotential for EO - E greater than about 100 mV (where 0 is reduced
immediately on reaching the electrode), but when the current is limited
by slow electron-transfer kinetics, ke continues to increas with
increasing overpotential.
In an experiment, the potential is stepped to some negative value
and the current measured as a function of time. With increasing time,
AVi increases, and f(A Vi) -+ 1. Thus for long enough time, current decay
348 Electron-Transfer Kinetics
Cyclic Voltammetry
0.4
0.3
~
en
.....,
0.2
'2
::l
0.1
t'
ell
.....,
1-0
0.0
.....
..c
1-0
-0.1
~
-0.2
-0.3
Figure 6.16 Computed cyclic voltammograms for (a) A = 10, 1, 0.1, and
0.01, a. = 0.5, and (b) A = 0.3 and C1. = 0.25 (dashed curve), 0.5 (solid curve),
and 0.75 (dot-dash curve).
350 Electron-Transfer Kinetics
200
6
Figure 6.18 Excess
(kinetic) separations of
the anodic and cathodic
>150 f- 6
peaks in cyclic voltam- ~
<.>
mograms of ferrocene Q,
in acetone (D), propio-
nitrile (0), and butyro-
~ 100
I
f-
6
6
g
D
0 -
nitrile (6) at 298 K (open <:!
symbols) and 198 K ~Q, 6 g
'-'
(filled
Reproduced
symbols).
with
50 I- D 6 -
D
permission from L. K. 6
!
0
Safford and M. J. 0
Weaver, J. Electroanal. 0 .B. Q D.
Chern. 1992 ,331, 857,
copyright 1992 Elsevier 0 25 50 75 100
Science Publishers. (uN s·1)112
or
dd<l> = Rsloro cos rot + 10 sin rot (6.44)
t Cs
The solution to this differential equation can be written
<l>(t) = ~<l> sin(rot + <p) + <l>dc
so that
~ = ro~<l> cos (rot + <p) = ro~<l> (cos rot cos <p + sin rot sin <p)
Equating the coefficients of sin rot and cos rot, we have
ro~<l> cos <p = ro IoRs
ro~<l> sin <p = IoiCs
so that the phase angle <p is given by
cot <p = cos <p/sin <p = roRsCs (6.45)
and the faradaic impedance Zr is
IZrl = ~<l> = -JRs2 + lI(roCs )2
10 (6.46)
§6.4 Faradaic Impedance 353
Thus the magnitude of the faradaic impedance is given by the ratio of
the amplitudes of the a.c. potential and a.c. current and the resistive and
capacitive components can be separated if the phase angle is known.
(6.47b)
Ret = OTt
al
=-B:!
FLO
(6.49)
where
a---L(~-~)
- FA V2DR V2Do
or, using eqs (6.50),
a-~( 1 + 1 ) (6.54)
- F2A CR*V2DR Co*V2Do
We can now identify the equivalent circuit parameters Rs and Cs by
comparison of eqs (6.44) and (6.53):
§6A Faradaic Impedance 355
cr/Vw = 478 Q
Cs = 1/('5'{(j) =3.33 JlF
With <p = 35°, cot <p = 1.428, we have
RetVw/cr = 0.428
Ret = 205 Q
356 Electron-Transfer Kinetics
A.C. Polarography
In a.c. polarography, a small sinusoidal a.c. voltage is
superimposed on the d.c. ramp voltage and the a.c. component of the cell
current is measured using a lock-in detector so that the in-phase and
out-of-phase current components can be determined separately.! In
other words, the cell potential is modulated by an amount l:!.E and the
resulting current modulation !::.i is measured. For small values of l:!.E,
the a.c. polarogram should approximate the first derivative of the d.c.
polarogram (see Problems). The a.c. modulation voltage is typically in
the audio-frequency range, ca. 10 Hz - 10 kHz, with an amplitude of 5-10
mY. In practice, the a.c. current oscillates with drop growth and
dislodgement; the current is usually measured at maximum drop size,
i.e., just before drop fall.
! Alternatively, the magnitude of the a.c. current and the phase angle can be
determined. Fourier transform methods can also be applied to a.c. polarography.
For further details on the theory and practice of a.c. polarography, see Smith (18).
§6.4 Faradaic Impedance 357
We can derive the a.c. current - d.c. potential curve for a reversible
process from the results of our discussion of the faradaic impedance. If
the process is reversible, the charge transfer resistance is negligible
and, according to eq (6.57), the faradaic impedance is
I Zrl = crY 2/ro
In general the surface concentrations CoCO,t) and CR(O,t) will be very
different from their bulk solution values. However, the sinusoidal
variations are about mean values determined by the d.c. potential. Thus
we can replace Co* and CR* in eq (6.54) by the mean values given by eqs
(4.3), which we can rewrite in slightly more compact form in terms of
the parameter
e' = exp F(E - EJ/2)
--'--=-=---=
RT
The mean surface concentrations (for CR* = 0) are
Co(O,t) = Co* ~
1+ e'
CR(O,t) = Co* YDolDR
1+ e'
Substituting these expressions in eq (6.54) in place of Co* and CR*, we
have
If the potential is
E(t) = Edc + till sin(rot)
the a.c. component of the current will be
I(t) = 10 sin(wt + 7U'4)
since cot <p = 1 for Ret = O. The current amplitude is
10=1~
or
10 = F 2ACo*YJ5O(.;) till e' (6.59)
RT (1 + e'f
Although the problem can be formulated to account for the expanding
spherical drop of the d.m.e., it is usually more convenient to make the
required corrections empirically by determining the ratio of the a.c. to
358 Electron-Transfer Kinetics
d.c. currents, measured for the same drop time on the same solution.
Using the Cottrell equation for iD, we obtain
10 FfiW)[!ill 8'
(6.60)
iD = RT(l + 8,)2
1.0
0.8
Q
:-s 0.6
r..
0
~Q.. 0.4
.....
0.2
(6.63)
('D
~) =~ vrttlD t1E ko
2RT
Cl)=oo
(6.65)
The electron-transfer rate constant can also be extracted from the
phase angle. According to eq (6.64), cot <p is linear in "f'iO)/'A.. If a. = 0.5, the
peak current corresponds to e' = 1 so that eq (6.38) reduces to A. = 2kc!W.
Thus a plot of cot <p us. Yo) should have a zero-frequency intercept of 1 and
a slope of VDl2lko.
50 I
2.5
I
[pillA I
I
I cot cp
I
40 I
I
I
2.0
I
I
I
I
30 I
I
I
1.5
""
ID
1.0
10
0 0.5
0 50 100 150 o 50 100 150
(OY'S-l )1/2 (OY'S-l )1/2
Figure 6.21 Frequency dependence Figure 6.22 Plots of cot cp us. 00 112
of the a.c. polarographic peak for a.c. polarographic currents
current for kO = (dashed line) 00 measured for trans-stilbene (0),
and ko = 0.01, 0.001, and 0.0001 m trimethyl-trans-stilbene (Il), and
s-l. Other parameters are as in hexamethyl-trans-stilbene (<». Re-
Figure 6.20. produced with permission from R.
Dietz and M. E. Peover, Disc. Far-
aday Soc. 1968,45, 154, copyright
1968 Royal Society of Chemistry.
e-
(1,5-COT)CoCp P (1,5-COT)CoCp- ( 1,5-COT)CoCp
Scheme I Scheme II
REFERENCES
PROBLEMS
6.1 Show that either eq (6.lla) or (6.llb) can be rearranged (with the
help ofthe Nernst equation) to
io = FAkoCOPCRP
6.2 Derive an equation for the current, analogous to eq (6.30), for the
case that Co* and CR* are both nonzero.
6.4 Show that the data discussed in Example 6.2 are consistent with
the simple mechanism
Mn3 + + e- ~ Mn 2+
when subjected to the more complete analysis of Example 6.3.
6.5 Figure 6.25 shows a Tafel plot for the reduction ofH+(aq) in a 0.1 M
solution of HCI at a nickel electrode in the presence of 1 bar H2(g) at
20°C.
(a) Determine the exchange current density jo and the apparent
transfer coefficient.
(b) Compute the electron-transfer rate constant ko.
1
c:1
S
~o
b.O ,
o
, ,,
,,
- -1
,
"",
-2
6.7 Bockris, Drazic, and Despic (see Example 6.4) found that the
exchange current density for the oxidation of iron depends on the
concentration of Fe2 + according to
dIn IJol -08
d In [Fe2+] - .
Show that the mechanism discussed in Example 6.4 leads to
d In liol = 1- a/2
dInGo
and thus is consistent with the experimental result.
6.8 The exchange current for an EE reduction with E2° »E1° and the
first electron-transfer step rate-limiting is given by eqs (6.27) and
the net current is given by the Butler-Volmer equation with J3app = 1
+ J3 where J3 is the anodic transfer coefficient for the rate-limIting
step. Derive analogous expressions for the case where the second
electron-transfer step is rate-limiting.
6.10 The Tome~ criterion for a reversible one-electron wave is !!.E = E1/4
-E3/4 = 56.51n mV at 25°C.
(a) Compute !!.E for several values of kD1ko in the range 0.1 to 2
using eq (6.31) (a = 0.5) and plot!!.E us. log(ko/kD).
(b) A r.d.e. voltammogram gave a value of !!.E = 75 mV after
correction for solution iR drop. Ifv = 10- 6 m 2 s- 1, D = 10-9 m 2 s- 1, and
ro = 500 s-1, what is the approximate value of ko?
6.11 (a) Show that the chronocoulometric charge for a potential step
experiment (see Problem 4.11) involving an electron-transfer step of
finite rate is
370 Electron-Transfer Kinetics
Q = FACO*kcA.-2(exp(A,2t)[I- erflNI"t)] + 2A,M -I}
where A, is given by eq (6.38). Hint: Integrate eq (6.39) by parts.
(b) For sufficiently long times, the first term in the bracketted
expression goes to zero and Q becomes linear in Vt. The linear
region of a plot of Q vs. Vt can be extrapolated to t = 0 or to Q = O.
Obtain expressions forQ extrapolated to t = 0 and for Vt
extrapolated to Q = O. Describe how such data could be used to
obtain the electron-transfer rate constant ko.
6.12 The following data were obtained for a potential step
chronocoulometry experiment. The solution contained 1 mM 0
which is reduced by one electron, and the potential step was to E =
Em - 50 mY. Assume that DO =10-9 m2 s- 1, (l =0.5,A =3.1 x 10-6 m 2
'tIs 1 3 5 7 9
Q/mC 1.13 2.10 2.78 3.33 3.81
371
372 Electrolysis
In modem terms, we would say that one mole of electrons (one Faraday)
will reduce one mole of Na+, one-half mole of Ca2 +, or one-third mole of
AI3+. Faraday's work not only put electrolysis on a quantitative basis but
laid the groundwork for the more complete understanding of the role of
electrons in chemistry which began to emerge 50 years later.
Exhaustive Electrolysis
When an electrolyte solution is subjected to electrolysis, either to
separate a component of the solution or to oxidize or reduce a substrate,
rather large amounts of electrical charge must be passed through the
cell. As we saw in Chapter 4 (Example 4.7), the net electrolysis at a
small electrode (e.g., a d.m.e.) in an unstirred solution is negligible, even
after several hours. Thus our first concern is to see how the current can
be increased to obtain significant net electrolysis in a reasonable time.
One obvious way to increase the current through an electrolysis cell
is to stir the solution, thus increasing the rate of mass transport. In a
stirred solution, the diffusion layer thickness can be reduced to as little
as tens of microns. Suppose that the electrode process is
O+n e- ~ R
and that a linear concentration gradient in 0 is set up across a diffusion
layer of constant thickness. If the potential of the working electrode is
sufficiently negative that the concentration of 0 is zero at the electrode
surface, then the current is given by eq (4.14b),
i = nFAkDCo* (4. 14b)
where kD = DlxD is the mass transport rate constant. The number of
moles of 0 reduced is related to the charge transferred by
dno=- dQ
nF
Since dQ =idt and dno =VdCo* (V is the solution volume), the change in
bulk concentration on passage of current for a time interval dt is
dCo*=-~dt
nFV
Substituting eq (4.14b) for i, we have
dCo* = - AkDCO* dt
V
§7.1 Bulk Electrolysis 373
Integration of this expression gives
CO*(t) = CO*(O) exp(-kt) (7.1)
where
k =knAfV (7.2)
Thus we expect the bulk concentration C* to fall exponentially with a
rate proportional to the rate of mass transport and to the ratio of the
electrode area to the solution volume. Thus if the process is to be
completed in minimum time, the general strategy for exhaustive
electrolysis should include efficient stirring and a high area-to-volume
ratio. Since the stirring must be particularly efficient at the working
electrode surface, careful attention to electrode geometry is required for
optimum performance.
Example 7.1 Suppose that 250 cm3 of 0.1 M CUS04 is
electrolyzed using a cathode of area 250 cm2 and an initial
current of 5 A. How long will it take to remove 99% of the
copper from solution? 99.99%? Assume that the cathode
process is rate limiting. From eq (4.14b), we obtain the
diffusional rate constant
kD = il2FACO*
Substituting the electrode area and the initial current and
concentration, we have
kD = 1.04 x 10-5 m s-l
With the area-to-volume ratio of 100 m- 1, eq (7.2) gives
k = 1.04 x 10-3 s-l
To reduce the copper concentration from 0.1 M to 0.001 M thus
requires a time
t =- [In(O.Ol)Vk =4440 s (74 minutes)
To reduce the concentration to 10- 5 M amounts to another
factor of 100 decrease in concentration and thus will require
another 74 minutes or 148 minutes in all.
Electroseparation
Electrolysis is often used to separate the components of a mixture.
For example, a desired metallic component might be plated out at the
cathode or a nonmetal could be oxidized to a gas or deposited as an
374 Electrolysis
1.0
"C
0.8
~
~
"C
Q) 0.6
~
d
.....
....0 0.4
~
e
c:.l
0.2
Figure 7.1 Optimum
separation is achieved in
electroseparations when 0.0
the potential is set 0.3 0.2 0.1 0.0 -0.1 -0.2 -0.3
midway between half-
wave potentials. (E - E 1I2 avg )N
The fractions Xl and X2 are plotted us. potential in Figure 7.1 for a
difference in standard potential, E1° - E2° =0.20 V and n1 =n2 = 1. It is
fairly easy to prove (see Problems) that, when n1 = n2, the optimum
separation is achieved, (i.e., Xl -X2 is maximized) when
E = (E1° + E2°)/2 (7.5)
When this condition is met,X2 = 1-X1 and
E I -E20 = RT In Xl (7.6)
2nF X2
or, at 25°C,
EI-E20 = QJl8,llog X I
n X2
Thus for X1/X2 = 999 (i.e., Xl = 0.999, X2 = 0.001), the difference in
standard potentials must be 0.355/n V.
In practice the situation is usually more complicated. Processes
are rarely strictly nernstian when large currents are flowing, and if the
interfering couple has very slow electrode kinetics, the potential
separation can be considerably smaller than predicted from eq (7.6). If
one or both of the couples involves deposition of a metal on the cathode,
the activity of the plated metal initially may be proportional to the
fraction reduced, but as a monolayer of atoms is formed on the cathode,
the activity approaches unity, i.e., the electrode surface approaches the
376 Electrolysis
property of the pure plated metal. In the limit of aR = 1, the electrode
potential is
E =Eo +RI.ln Co
nF
so that the electrode potential for small Co may be rather more negative
than expected from eq (7.4a).
Despite these complications, the simple model is useful in
predicting electroseparation efficiencies when both couples have similar
properties. Thus if both couples involve the deposition of a metal and
they have similar rates, eq (7.6) gives a semiquantitative guide to the
required difference in standard potentials.
Oliver Wolcott Gibbs (1822-1908) taught at City College of New York from
1849 until his appointment as Rumford Professor at Harvard in 1863.
The following decade saw many contributions to the analytical chemistry
of the platinum metals as well as the development of electrogravimetry.
A curricular change at Harvard in 1871 effectively ended Gibbs' career
as a research scientist.
salt
bridge
reference
electrode
icell = i 1 + i2
Because of the capacitive feedback loop, the potential at the inverting
input of the operational amplifier is a virtual ground, 0 V. Thus the
potentials across resistors Rl and R2 must be equal:
ilRl = i~2
With this relationship, we obtain
. . Rl + R2
lcell = l2-'--"'='-~
Rl
382 Electrolysis
cell
power
r
supply
Since the input impedance of the operational amplifier is very large, the
current i2 must produce a charge Q2 on the capacitor C:
. dQ2
L2=di
But the potential at the output is proportional to this charge
Q2 = C <Pout
Thus
Rearranging, we have
or
Qcell = - R 1 + R2 C<Prut
Rl
By selecting appropriate values of RI, R2, and C, charges ranging from
less than 1 IlC up to about 1 C are easily measured. The method requires
accurately known resistances R I and R2 and capacitance C as well as
an accurate potential measurement.
§ 7.2 Analytical Applications of Electrolysis 383
2 02N-Q-0H+34H++34e- - <:>-NH-NHQOH
cell
power
supply
Figure 7.5 A simple circuit for
constant current electrolysis.
+
Reference
Monitor
lOne of the earliest applications of the biamperometric method was in the titration of
iodine solutions with standard sodium thiosulfate solution from a buret. The 82032-
/S406 2 - couple is irreversible, so that a steadily decreasing current is observed up to
the endpoint, with essentially zero current beyond. Such a titration was called a
dead-stop titration, but this name is inappropriate when the titrant furnishes the
reversible couple.
§7.2 Analytical Applications of Electrolysis 389
Stripping Voltammetry
NOa-+H20+2e- ~ N02-+20H-
The local increase in pH led to the precipitation of various metal
hydroxides on the electrode. On the anodic scan, these gave rise to a
mercury oxidation peak:
HgO) + M(OH)2 ~ Hg(OH)2 + M2+ + 2 e-
Another source of trouble is the formation of intermetallic
compounds in the mercury phase. Thus reduction of solutions
containing gold and cadmium or zinc leads to the intermetallic
compounds AuCd, AuaCd, and AuZn. These species oxidize at a
different potential than cadmium or zinc amalgam, so that the normal
cadmium and zinc peaks in thevoltammetric scan are smaller than
expected and extra peaks are observed due to oxidation of the
intermetallic compounds.
The stripping voltammetry strategy has been most commonly used
to concentrate metals in a mercury drop or thin-layer electrode, followed
by anodic stripping. The same strategy can often be applied in reverse,
depositing the analyte on the electrode surface in an oxidation step; the
analysis is then based on a cathodic stripping voltammetric scan.
7.3 ELECTROSYNTHESIS
-2X-
The reduction is stereospecific. Thus, for example, reduction of d,l-3,4-
dibromobutane in DMF (Rg cathode, -1.1 V vs. s.c.e., BU4NBF 4
electrolyte) gives 100% cis-2-butene. Reduction of the meso isomer under
the same conditions gives trans-2-butene (18).
With 1,3- and 1,4-dihalides, cyclization competes with proton
abstraction. The detailed mechanism is unclear. It may be that two-
electron reduction of one end of the molecule leads to a carbanion,
392 Electrolysis
which then displaces X- from the other end, forming a ring. However,
there is some evidence to suggest that C-C bond formation is concerted
with ionization of the halide ions and electron transfer. The product
distribution depends on the reduction potential. Thus, for example,
reduction of 1,4-dibromobutane at -1.75 V (us. s.c.e.) in DMF solution
gives 26% cyclobutane and 74% butane; under these conditions, step-wise
reduction of the two C-Br bonds seems to occur with protonation
competing successfully with cyclization (19). However, if the reduction is
carried out at -2.3 V, the products are 90% cyclobutane and only 10%
butane, suggesting that the concerted mechanism may be operative at
the more negative potential. Electrochemical reductive elimination
processes are particularly useful for formation of strained ring systems.
Thus reduction of 1,3-dibromo-2,2-di(bromomethyl)propane gives first
1,1-di(bromomethyDcyclopropane and then, on further reduction,
spiropentane (19):
CH2Br CH B
I +2e- 2 r +2e-
BrCHr-C-CH2Br - ~ - ~
I - 2 Br- (../\ - 2 Br- V""-J
CH 2Br CH2Br
~~c¢o~
~+H+ ~.
I~~~
~ +H+ vyv
H HH
If the potential is controlled, reduction stops at this point since the
isolated benzene rings are much more difficult to reduce than
anthracene. Other aromatic hydrocarbons behave similarly; thus,
naphthalene is reduced to 1,4-dihydronaphthalene and phenanthrene to
9,10-dihydrophenanthrene. Pyrene can be reduced to 4,5-dihydropyrene
and, at more negative potentials, to 4,5,9,10-tetrahydropyrene:
§7.3 Electrosynthesis 393
[ Ph_N_%_Ph]Z. -OH-
:..!: Ph-N=N-Ph
o
I I
0
When the solution is alkaline, protonation of the nitrosobenzene radical
anion apparently is slow and this species dimerizes to give, after
protonation and loss ofOH-, azoxybenzene, which can be isolated in high
yield. Further reduction at low pH yields hydrazobenzene which
rearranges to benzidine in strong acid:
+4e- H+ ~
Ph-N=~-Ph ~ Ph-NH-NH-Ph - H2N~NH2
-H20
At higher temperatures or in more strongly acidic media, PhNHOH
rearranges to p-aminophenol. At low pH, and at a more negative
potential, it can be reduced to aniline:
HO -0-
~ ~
H+
NH2 - - PhNHOH
high
+2e-
- - PhNH a+
+ 3 H+
temperature
In general, the reduction of nitro compounds thus can lead to a wide
variety of products, and by judicious control of solution polarity, acidity,
and temperature, high specificities can be achieved.
The reduction of carbonyl compounds proceeds analogously. For
example, benzophenone is reduced in neutral or alkaline media to
benzhydrol
394 Electrolysis
- + e-
- HOI
OH
oII
Ph2
- -e
+ Nu-
, 1
NupC-C-Nu
I \
\ I I I
Nu-C-C-C-C-Nu
I I I \
or to substitution, via proton loss, oxidation and nucleophilic addition to
the allyl cation:
I I
-C-C=C
1-
-e- , I -e- [,
. c-c=c
I 1- + + Nu- I I
c-=-=-c-=c
I] - -c-c=c
1
I \ - H+ I \ I \ Nu \
Examples of all these routes can be found. Thus electrolysis of
cyclohexene in acetic acid with Et4NOTs supporting electrolyte led to
55% 3-acetoxycyclohexene and 12% trans-1,2-diacetoxycyclohexane (21).
Oxidation of styrene in methanol solution containing sodium methoxide
and NaCI04 supporting electrolyte gave 64% 1,4-dimethoxy-1,4-
diphenylbutane (22).
Oxidation of conjugated olefins leads primarily to 1,2- or 1,4-addition
via allylic radical/cation intermediates:
§7.3 Electrosynthesis 395
c-e
I I .
C
.... C~ ........ C~ ... -
::>:
I I
,]+
I ,-e'
-C-C-C=C -
I . '
Nu
I I [,
-c-c--=c:.:.;c
I
Nu
I I
\ -
+Nu' I I I I
Nu
I I
-C-c=C-C- + -C-C-C=C
I
Nu
I I
Nu
I I
I
"-
Nu
ArH -
-e.
[ArH]
+ + Nu-
-
~H
A\. -
-e- [ "H
A\
1+ --H +
ArNu
Nu Nu
but the substitution product is often more easily oxidized than the
starting hydrocarbon, leading to reduced yields. Better results are
obtained when the product is less easily oxidized. Thus, for example,
oxidation of 1,4-dimethoxybenzene in methanol with KOH leads to a 75%
yield of quinone bisacetal (24):
CHgO OCHg
OCHg
CHgOH/KOH
•
Q
CHgO OCHs
,
radical intermediate, which rapidly loses C02 and dimerizes to the
hydrocarbon product:
-e-
RCO£ - RC0 2" - R" + C02
R-R
Yields vary from 50 to 90%, quite a bit higher than might be expected
considering the reactions open to an alkyl radical; indeed, there is good
396 Electrolysis
reason to believe that the radicals are adsorbed on the electrode surface
and protected to some extent from the solvent (from which hydrogen
abstraction might be expected). Three kinds of side reactions compete
with radical dimerization: (1) In cases where radical rearrangement
can occur, a variety of coupling products is obtained. Thus, for example,
a f3;y-unsaturated acid leads to an allylic radical intermediate,
R-CH=CH-CH 2 • R-CH-CH=CH2
and three coupling products are obtained:
R-CH=CH-CH2-CH2-CH=CH-R
CH2=CH -CHR-CHR-CH=CH2
R-CH=CH2-CHR-CH=CH2
(2) When the radical intermediate is secondary or tertiary, or substituted
such that the carbonium ion is stabilized, further oxidation takes place:
-e-
R·-R+
and the eventual products are derived from carbonium ion reactions. (3)
In cases where the acyloxy radical is stabilized by conjugation (aromatic
or a.,f3-unsaturated acids), Kolbe products are not obtained, most likely
because the acyloxy radical survives long enough to escape from the
electrode surface. Thus benzoic acid gives benzene under Kolbe
conditions, presumably through loss of C02 and hydrogen abstraction
from the solvent by the resulting phenyl radical.
Electroplating
An important application of electrodeposition is the plating of a
layer of metal on an object to improve the appearance or to impart
hardness or corrosion resistance. Metals commonly used in
electroplating include Cr, Ni, Zn, Cd, Cu, Ag, Au, Sn, and Pb. Some
alloys such as brass (Cu/Zn) and bronze (Cu/Sn) also can be
electroplated.
The basic theory of electroplating is extremely simple. For example,
to plate copper on a steel substrate, the object to be plated is made the
cathode in an electrolysis cell where a piece of copper is used as the
anode. There are many subtleties, however, which must be considered
in practice. Thus, in plating copper on steel, for example, we
immediately recognize that the reaction
Cu2 +(aq) + Fe(s) ~ Cu(s) + Fe 2 +(aq)
is spontaneous. To prevent the dissolution of iron from the substrate, the
activity of Cu2 +(aq) must be reduced by the addition of a complexing
agent which coordinates strongly to Cu(Il) but much less so than Fe(II).
In electroplating applications, it is usually desirable to deposit a
layer of uniform thickness. This requirement is not difficult to meet if
the substrate has a simple geometry. If there are holes or recesses,
however, a uniform layer can be quite difficult to achieve. When a
potential is applied across an electrolysis cell, the potential drop is the
sum of several contributions:
(1) the equilibrium anode-solution and solution-cathode potentials (i.e.,
the equilibrium cell potential);
(2) the activation and polarization overpotentials at the anode and
cathode; and
(3) the solution iR drop.
§7.4 Industrial Electrolysis Processes 399
Anodization
The electrochemical formation of a protective oxide layer on a metal
surface is called anodization. The process is most commonly applied to
aluminum but sometimes also to copper, titanium, and steel. The metal
object to be anodized is made the anode of an electrochemical cell and
current is passed through an acidic electrolyte solution, so that the
electrode process (for aluminum) is
2 Al + 3 H20 --+ Al203(S) + 6 H +(aq) + 6 e-
The process is much like electroplating, although the cell polarity is
reversed.
The formation of the oxide layer involves a complicated electrode
process which tends to be slow, requiring a large overpotential,
particularly after a thin oxide layer is formed. For this reason, throwing
power is not a problem in anodization; the oxide layer grows most
rapidly where it is thinnest.
Aluminum is anodized using sulfuric acid, chromic acid, or oxalic
acid baths, depending on the nature of the surface desired. Anodized
aluminum surfaces can be colored by adsorbing an organic dye on the
surface. Alternatively, transition metals can be deposited in the pores of
the oxide layer; colors then arise from interference effects.
Carbon Anode
Molten Cryolite
& Alumina
Molten
Aluminum
Carbon Cathode
<E--Hg
50% NaOH
The oldest technology divides the overall process into two parts by
using a cell with a mercury cathode as shown in Figure 7.8. Since the
1 Hydrogen is also produced from coal or natural gas as "synthesis gas" (a mixture
of H2 and CO) or, when very high purity is required, from the electrolysis of water.
§7.4 Industrial Electrolysis Processes 405
overpotential for H2(g) production on Hg is so high, the cathode process
is the formation of sodium amalgam,
Na+(aq) + e- -+ Na(Hg)
The anode is titanium, coated with RU02, C0203, or other transition
metal oxides to catalyze the evolution of chlorine gas,
2 CI- -+ CI2(g) + 2 e-
Since the evolution of 02 is thermodynamically more favorable than that
of CI2(g), the anode must have a high overpotential for water oxidation so
as to prevent contamination of the Cl2 product. The reaction of sodium
amalgam with water is thermodynamically favorable but is very slow,
owing to the large overpotential for H2 production. Sodium amalgam is
withdrawn from the cell at about 0.5% Na content, washed with water to
remove NaCI and passed with water through a column packed with
graphite impregnated with Fe, Ni, or some other catalyst. Hydrogen is
then evolved and collected at the top of the column. Mercury and 50%
NaOH solution emerge at the bottom of the column in quite pure form.
The NaCI feedstock is typically 50% brine which has been treated with
base to precipitate group II and transition metal hydroxides and
reacidified to pH 4. The cell normally operates at about 60°C.
The mercury cell technology produces high-purity products with
little additional treatment required. However, there is a potentially
serious problem of mercury contamination in the plant environment
and in the wastewater effluent. Because Na+ is reduced in the
electrolysis step rather than H+, the cell potential-and thus energy
consumption-is significantly higher than should be necessary (in
principle) for the electrolysis of NaCI to form Ch(g) and H2(g).
Beginning in the 1950's, environmental concerns prompted the
development of another approach, the so-called diaphragm cell. In this
method, the cathode material is steel gauze coated with a catalyst to
reduce the H2 overpotential. An asbestos-based diaphragm is used to
separate the anode and cathode compartments to prevent mixing of the
hydrogen and chlorine gases and to somewhat reduce mixing of the
N aCI feedstock on the anode side with the N aOH solution of the cathode
side. The arrangement is shown schematically in Figure 7.9. While this
eliminates the mercury problem, the cell potential is still much larger
than theoretical because of an iR drop of about 1 V across the
diaphragm. This is not the only problem: (1) OH- diffusing into the
anolyte reacts with Cl2 to form hypochlorite ion:
Cl2 + 2 OH- -+ H20 + CIO- + CI-
reducing the yield of C12; (2) when the anolyte pH goes up, the potential
for formation of 02 at the anode becomes more favorable and
contamination of the CI2(g) results; (3) to reduce the magnitude of
problems (1) and (2), the catholyte cannot be allowed to exceed 10% in
406 Electrolysis
NaOH concentration; (4) chloride ion diffusion from the anolyte to the
catholyte results in a very considerable contamination of the NaOH
solution; (5) to deal with problems (3) and (4), evaporation of the caustic
H2 Cl 2 Cl 2
t t t
~
0
CIS
--
Q
Z
CIS
Q
Z
CIS
Z
~ ~ ~
0 0
0 C',) C',)
,...;
Electrosynthesis of Adiponitrile
The largest volume organic electro synthetic process presently in
use is the production of adiponitrile, a precursor to nylon, from
acrylonitrile. The stoichiometric half-cell reactions are
2 CH2=CHCN + 2 H20 + 2 e- ~ NC-(CH2)6-CN + 2 OH-
2 H20 ~ 02(g) + 4 H + + 4 e-
which sum to the overall reaction stoichiometry
4 CH2=CHCN + 2 H20 ~ 02(g) + 2 NC-(CH2)6-CN
The actual mechanism of the reaction is considerably more complex and
probably involves several parallel pathways. Some of the possible
reaction pathways are shown in Figure 7.10. The major side reactions
which must be minimized are the production of propionitrile (the
monomeric reduction product) and trimers and higher polymers.
,
CH2CHCW
.
l
CH2CHCN
+e-
trimer,
- etc .
.CH2CH2CN
+SH
-S t+ e-
CH2CH2CN"" +SH
t + H+
-S
CH3CH2CN
since this avoids the neutral radical, which is the most likely route to
higher polymers. On the other hand, the pH cannot be allowed to get too
high since acrylonitrile reacts with hydroxide ion to form ~
hydroxypropionitrile and bis-(2-cyanoethyl)ether. These considerations
led to a definite strategy for the process:
(1) The concentration of acrylonitrile should be as high as possible to
promote dimerization; in practice an aqueous emulsion is used so
that the aqueous phase is saturated, about 7% by weight. The
organic phase is a mixture of acrylonitrile and adiponitrile.
(2) The aqueous phase must be buffered to prevent reaction of
acrylonitrile with base; in practice the emulsion contains about 15%
Na2HP04.
(3) The electrode surface pH should be high to prevent protonation of
the initially formed radical anion. This can be achieved by
adsorbing positive ions on the electrode surface (see §2.2); the
industrial process uses the diquaternary ammonium ion, BU2EtN-
(CH2)6-NEtBu22+.
(4) Hydrogen evolution at the cathode must be suppressed so that the
cathode material should have a large overpotential for H+
reduction; in practice, cadmium-plated steel is used.
Oxygen is produced at a steel anode, spaced 2 mm from the cathode l
to reduce solution iR drop. Corrosion of the anode is a problem since
deposition of iron on the cathode would promote hydrogen reduction and
a decrease in current efficiency. Additives are used to suppress anode
dissolution and EDTA is added to scavenge transition metals before they
plate out on the cathode.
The two-phase mixture is cycled rapidly between the electrolysis
cell and a reservoir. Solution is continually withdrawn from the
reservoir for product isolation, removal of by-products and metal ions,
and recycling of unreacted acrylonitrile. The overall process has about
90% selectivity toward adiponitrile. The reversible cell potential is about
-2.5 V; the overpotentials and solution iR drop are relatively small at the
current density used (about 2000 A m- 2) so that the working cell potential
is about -3.8 V.
1 About 100 anode-cathode pairs are used in the cell to allow for greater output.
§7.4 Industrial Electrolysis Processes 411
Lead Rod
Electrode
Lead Shot
(anode)
tIl--+- Polypropylene
diaphram
1 With the phase-out of leaded gasoline in the United States and Europe, this process
has decreased in importance; tetraethyllead is still used as a motor fuel additive in
much of the rest of the world.
412 Electrolysis
stripped from the solution and recycled to the Grignard reactor, the
MgCl2 is extracted with water, and the ether solvent is separated from
the lead tetraalkyl, purified by azeotropic distillation and recycled to the
Grignard reactor.
7.5 CORROSION
1.5
:: ~
E/v·:S.
J.; ••••••
...........~:;;:;.............
Fe 2+(aq)
0.0 ••••••••••••••••••••••
-0.5
Figure 7.12 Predominance Fe(s)
area diagram for iron. The
dashed lines correspond to -1.0
the H+(aq)/H2(g) and o 5
pH
10 15
02(g)/H20 couples.
2
o
~
S
S0
C'l
------
,~.-----=-"-::-~-
<
-
< ~-2
:;:;:'-2 o
-0.2 -0.3 -0.4 -0.5 -0.6 -0.7 -0.1 -0.2 -0.3 -0.4 -0.5 -0.6 -0.7
ElY ElY
Figure 7.13 Current density us. Figure 7.14 Tafel plot corresponding
potential curve for iron in contact to the current density us. potential
with an aqueous acidic solution (pH curve of Figure 7.13.
2, [Fe 2+] = 1 mM) saturated with H2
(1 bar partial pressure).
1 We assume that the experiment is done in an oxygen-free solution and that the
anion is not electroactive.
§7.5 Corrosion 415
Passivation
Some metals-most notably aluminum, chromium, and nickel-
are essentially unreactive in air-saturated water despite a large
thermodynamic driving force. Aluminum, for example, can be used in
cookware despite an expected free energy of oxidation of -717 kJ mol-I.
Metals are said to be passive when they corrode anomalously slowly.
Many other metals, including iron, show passive behavior under some
circumstances and "normal behavior" under others. It has been known
since the eighteenth century that iron is attacked rapidly by dilute nitric
acid but is essentially immune to attack by concentrated nitric acid.
Treatment of iron with concentrated nitric acid imparts temporary
corrosion resistance when the iron is immersed in dilute acid. Iron can
also be temporarily passivated by electrochemical anodization. Whereas
iron loses passivity in a matter of minutes, chromium and nickel show
long-term passive behavior.
Differential Aeration
Although ionic conduction across a passivating metal oxide film is
very slow, electronic conduction may be much faster. Thus even at a
passivated metal surface, reduction of oxygen may occur provided that a
source of electrons is available. Consider the experimental
arrangement shown in Figure 7.16. Two initially identical pieces of
metal are dipped into identical electrolyte solutions. Nitrogen is bubbled
over one electrode and oxygen over the other and we suppose that the
metal exposed to oxygen is passivated. If we connect the two electrodes
by closing the switch, we have set up an electrochemical cell in which, at
the left-hand electrode, metal is oxidized,
M(s) ~ Mn+ + n e-
and at the right-hand electrode, oxygen is reduced,
02(g)+4H++4e- ~ 2 H20
This overall cell reaction will almost always be spontaneous and the rate
may be quite significant, despite the fact that the right-hand electrode is
passivated.
+-°2
Prevention of Corrosion
Perhaps the most common approach to the prevention of corrosion
is to cover the metal surface with a protective coating, e.g., paint, plastic,
ceramic, or an electroplated layer of a passive metal such as chromium,
nickel, or tin. A mechanical barrier is effective, of course, only as long
as it remains a barrier. Bare metal exposed by a scratch is susceptible to
corrosion. Once corrosion begins at a breach in the protective layer, it
often continues under the layer through the differential aeration
mechanism, lifting the coating and eventually leading to massive
damage. Differential aeration is particularly effective with damaged
metal-plated surfaces since oxygen reduction can occur over the entire
undamaged surface to balance the base metal oxidation at the point of
damage.
Coatings can be made more effective by the addition of corrosion
inhibitors. A corrosion inhibitor works by decreasing the rate of metal
oxidation, by decreasing the rate of oxygen (or water) reduction, or by
shifting the surface potential into the pastsive region. Thus rust-
inhibiting paints often contain chromate salts which passivate the metal
surface by making the surface potential positive. Other common paint
additives are aliphatic or aromatic amines which appear to adsorb on
the metal, slowing the metal oxidntion half-reaction. Since inhibitors
impart a resistance to corrosion over and above the effect of the physical
barrier, small scratches do not usually lead to massive damage.
Another way to protect a metal object from corrosion is to attach a
sacrificial anode made of zinc, magnesium, or an aluminum alloy. In
effect a short-circuited galvanic cell is produced where the protected
object is the cathode (oxygen is reduced) and the more active metal is
§7.5 Corrosion 421
REFERENCES
PROBLEMS
7.2 Repeat the calculation of Example 7.1 assuming that the solution
is unstirred and that the current remains diffusion-controlled
indefinitely. The diffusion coefficient of Cu2+ is ca. 7 x 10- 10 m 2 s- 1.
7.10 Suppose that you have a constant current source which can
deliver 1, 2, 5, 10, 20, 50, or 100 rnA (±O.25%), an endpoint detection
system and timer which is reliable to ±1 s, and a collection of
pipets (1,2,5,10,25, and 50 mL), each of which is accurate to ±0.02
mL. You wish to do a coulometric titration of a solution where the
analyte concentration is approximately 1 mM. What volume and
current should be chosen if the analytical uncertainty must be
less than 1.0% and it is desirable to use minimum sample size
and do the titration in minimum time? Assume a one-electron
oxidation is involved.
7.14 Two common ways of protecting steel from corrosion are tin
plating and zinc coating. Typical examples of this approach are
"tin cans" and galvanized buckets. Discuss qualitatively the
chemical reactions which would occur if pieces of tin-plated steel
and zinc-coated steel were scratched to expose the steel and placed
in a wet oxygen-rich environment. Indicate clearly where
(relative to the scratch) each reaction would take place.
427
428 Appendix 1
B4 D. A. MacInnes, The Principles of Electrochemistry, New York:
Dover, 1966 (corrected reprint of 1947 edition).
B5 G. Kortum, Treatise on Electrochemistry, 2nd ed, Amsterdam:
Elsevier, 1965.
136 B. E. Conway, Theory and Principles of Electrode Processes, New
York: Ronald Press, 1965.
B7 K. J. Vetter, Electrochemical Kinetics, New York: Academic Press,
1967.
B8 J. O'M. Bockris and A. Reddy, Modern Electrochemistry, New York:
Plenum Press, 1970.
139 R. A. Robinson and R. H. Stokes, Electrolyte Solutions, 2nd ed (rev),
London: Butterworths, 1970.
B10 J. S. Newman, Electrochemical Systems, Englewood Cliffs, NJ:
Prentice-Hall,1972.
Bll L. I. Antropov, Theoretical Electrochemistry, Moscow: Mir, 1972.
B12 A. J. Bard and L. R. Faulkner, Electrochemical Methods, New
York: John Wiley, 1980.
B13 R. Greef, R. Peat, L. M. Peter, D. Pletcher, and J. Robinson,
Instrumental Methods in Electrochemistry, Chicester: Ellis
Horwood, 1985.
B14 J. Koryta and J. Dvorak, Principles of Electrochemistry, New York:
John Wiley, 1987.
c. Specialized Books and Monographs
Cl W. Ostwald, Electrochemistry, History and Theory, Leipzig: Veit,
1896. Republished in English translation for the Smithsonian
Institution, New Delhi: Amerind Publishing, 1980.
C2 V. G. Levich, Physiochemical Hydrodynamics, Englewood Cliffs,
NJ: Prentice-Hall, 1962.
C3 R. N. Adams, Electrochemistry at Solid Electrodes, New York:
Marcel Dekker, 1969.
C4 C. K. Mann and K. K. Barnes, Electrochemical Reactions in
Nonaqueous Systems, New York: Marcel Dekker, 1970.
C5 A. Weissberger and B. W. Rossiter, eds, Physical Methods of
Chemistry, Vol. 1 (Techniques of Chemistry), Parts IIA and lIB
(Electrochemical Methods), New York: John Wiley, 1971.
C6 J. S. Mattson, H. B. Mark, Jr., and H. C. MacDonald, Jr.,
Electrochemistry; Calculations, Simulation, and Instrumentation
(Computers in Chemistry and Instrumentation, Vol. 2), New York:
Marcel Dekker, 1972.
C7 Yu. V. Pleskov and V. Yu. Filinovskii, The Rotating Disc Electrode,
New York: Consultants Bureau, 1976.
CB J. O'M. Bockris and S. U. M. Khan, Quantum Electrochemistry,
New York: Plenum Press, 1979.
Bibliography 429
D. Electroanalytical Methods
D1 1. M. Kolthoff and J. J. Lingane, Polarography, 2nd ed, New York:
Interscience, 1952.
D2 J. J. Lingane, Electroanalytical Chemistry, 2nd ed, New York:
Interscience, 1958.
D3 W. C. Purdy, Electroanalytical Methods in Biochemistry, New
York: McGraw-Hill,1965.
D4 L. Meites, Polarographic Techniques, 2nd ed, New York: John
Wiley, 1965.
D5 J. Heyrovsky and J. Kuta, Principles of Polarography, New York:
Academic Press, 1966.
D6 H. Rossotti, Chemical Applications of Potentiometry, Princeton, NJ:
Van Nostrand, 1969.
D7 R. G. Bates, Determination of pH: Theory and Practice, 2nd ed, New
York: John Wiley, 1973.
DB Z. Galus, Fundamentals of Electrochemical Analysis, Chichester:
Ellis Harwood, 1976.
D9 G. Dryhurst, Electrochemistry of Biological Molecules, New York:
Academic Press, 1977.
D10 C. C. Westcott, pH Measurements, New York: Academic Press,
1978.
Dll J. Vesely, D. Weiss, and K. Stulik, Analysis with Ion-Selective
Electrodes, Chichester: Ellis Horwood, 1978.
D12 A. M. Bond, Modern Polarographic Methods in Analytical
Chemistry, New York: Marcel Dekker, 1980.
D13 J. A. Plambeck, Electroanalytical Chemistry, New York: John
Wiley, 1982.
D14 J. Koryta and K. Stulik, Ion-Selective Electrodes, 2nd ed, London:
Cambridge University Press, 1983.
E. Organic Electrosynthesis
El M. R. Rifi and F. H. Covitz, Introduction to Organic
Electrochemistry, New York: Marcel Dekker, 1974.
430 Appendix 1
F. Experimental Methods
F1 D. J. G. Ives and G. J. Janz, eds, Reference Electrodes, Theory and
Practice, New York: Academic Press, 1961.
F2 W. J. Albery and M. L. Hitchman, Ring-Disc Electrodes, Oxford:
Clarendon Press, 1971.
F3 D. T. Sawyer and J. L. Roberts, Jr., Experimental Electrochemistry
for Chemists, New York: John Wiley, 1974.
F4 E. Gileadi, E. Kirowa-Eisner, and J. Penciner, Interfacial
Electrochemistry--An Experimental Approach, Reading, MA:
Addison-Wesley, 1975.
F5 D. D. MacDonald, Transient Techniques in Electrochemistry, New
York: Plenum Press, 1977.
F6 P. T. Kissinger and W. R. Heineman, eds, Laboratory Techniques in
Electroanalytical Chemistry, New York: Marcel Dekker, 1984.
G. Technological Applications of Electrochemistry
G1 J. O'M. Bockris and S. Srinivasan, Fuel Cells: Their
Electrochemistry, New York: McGraw-Hill,1969.
G2 C. L. Mantell, Batteries and Energy Systems, New York: McGraw-
Hill,1970.
G3 F. A. Lowenheim, ed, Modern Electroplating, New York: John
Wiley, 1974.
G4 S. W. Angrist, Direct Energy Conversion, Boston: Allyn and Bacon,
1976.
G5 V. S. Bagotzky and A. M. Skundin, Chemical Power Sources, New
York: Academic Press, 1980.
G6 D. Pletcher, Industrial Electrochemistry, London: Chapman and
Hall,1982.
Bibliography 431
G7 N. L. Weinberg and B. V. Tilak, eds, Technique of Electroorganic
Synthesis (Technique of Chemistry, Vol. V), Part III, New York:
John Wiley, 1982.
G8 H. H. Uhlig and R. W. Revie, Corrosion and Corrosion Control, New
York: John Wiley, 1984.
G9 H. V. Ventatasetty, ed, Lithium Battery Technology, New York:
John Wiley, 1984.
G10 R. E. White, ed, Electrochemical Cell Design, New York: Plenum
Press, 1984.
Gll Z. Nagy, Electrochemical Synthesis of Inorganic Compounds, New
York: Plenum Press, 1985.
H. Electrochemical Data
HI W. M. Latimer, Oxidation Potentials, 2nd ed, Englewood Cliffs, NJ:
Prentice-Hall, 1952.
H2 B. E. Conway, Electrochemical Data, Amsterdam: Elsevier, 1952.
H3 R. Parsons, Handbook of Electrochemical Data, London:
Butterworths, 1959.
H4 A. J. de Bethune and N. A. S. Loud, Standard Aqueous Electrode
Potentials and Temperature Coefficients at 25°C, Skokie, IL:
Hampel, 1964.
H5 M. Pourbaix, Atlas of Electrochemical Equilibria, New York:
Pergamon Press, 1966.
H6 G. J. Janz and R. P. T. Tomkins, Nonaqueous Electrolytes
Handbook, New York: Academic Press, 1972.
H7 A. J. Bard and H. Lund, eds, The Encyclopedia of the
Electrochemistry of the Elements, New York: Marcel Dekker, 1973.
H8 L. Meites and P. Zuman, Electrochemical Data. Part I. Organic,
Organometallic, and Biochemical Systems, New York: John Wiley,
1974.
H9 D. Dobos, Electrochemical Data, Amsterdam: Elsevier, 1975.
H10 G. Milazzo and S. Caroli, Tables of Standard Electrode Potentials,
New York: John Wiley, 1977.
Hll L. Meites and P. Zuman, eds, CRC Handbook Series in Organic
Electrochemistry, Boca Raton, FL: CRC Press, 1977-.
H12 L. Meites, P. Zuman, E. B. Rupp and A. Narayanan, eds, CRC
Handbook Series in Inorganic Electrochemistry, Boca Raton, FL:
CRC Press, 1980-.
H13 A. J. Bard, R. Parsons, and J. Jordan, eds, Standard Potentials in
Aqueous Solution, New York: Marcel Dekker, 1985.
H14 A. J. Bard, R. Parsons, and J. Jordan, eds, Oxidation-Reduction
Potentials in Aqueous Solution, Oxford: Blackwell, 1986.
432 Appendix 1
1. Review Series
11 J. O'M. Bockris and B. E. Conway, eds, Modern Aspects of
Electrochemistry, New York: Plenum Press, from 1954.
12 P. Delahay (Vols. 1-9), C. W. Tobias (Vols. 1- ), and H. Gerischer
(Vols. 10- ), Advances in Electrochemistry and Electrochemical
Engineering, New York: John Wiley, from 1961.
13 A. J. Bard, ed, Electroanalytical Chemistry, New York: Marcel
Dekker, from 1966.
14 E. B. Yeager and A. J. Salkind, eds, Techniques of Electrochemistry,
New York: John Wiley, from 1972.
15 G. J. Hills (Vols. 1-3) and H. R. Thirsk (Vol. 4- ), Senior Reporters,
Electrochemistry, A Specialist Periodical Report, London: Royal
Society of Chemistry, from 1971.
16 Analytical Chemistry, Fundamental Annual Reviews (April issue
of even-numbered years), Washington: American Chemical
Society.
2
-zc>< SYMBOLS AND UNITS
w
a..
a..
til(
433
434 Appendix 2
438
Electrochemical Data 439
Half-Cell Reaction
C02+ + 2 e- --+ Co(s) ....().2:l7
NiOis) + 4 H+ + 2 e- --+ Ni2+ + 2 H20 1.593
Ni2+ + 2 e- --+ Ni(s) ....().257
Ni(OH>2(s) + 2 e- --+ Ni(s) + 2 OH- ....().72
Cu2+ + e- --+ Cu+ 0.159
CuCI(s) + e- --+ Cu(s) + CI- 0.121
Cu2+ + 2 e- --+ Cu(s) 0.340
Cu(NHa)42+ + 2 e- --+ Cu(s) + 4 NHa ....().oo
Ag2+ + e- --+ Ag+ 1.980
Ag+ + e- --+ Ag(s) 0.7991
AgCI(s) + e- --+ Ag(s) + CI- 02223
Zn2+ + 2 e- --+ Zn(s) ....().7626
Zn(OH)42- + 2 e- --+ Zn(s) + 4 OH- -1.285
Cd 2+ + 2 e- --+ Cd(s) ""().4025
2 Hg2+ + 2 e- --+ Hg22+ 0.9110
Hg22+ + 2 e- --+ 2 Hg(l) 0.7960
Hg2CI2(S) + 2 e- --+ 2 Rg(l) + 2 CI- 0.26816
EO'/V
Couple 1MHCI04 1MHCl
Ag(l)lAg( 0) 0.792 0.228 0.77
As(V)1As(III) 0.577 0.577
Ce(IV)/Ce(III) 1.70 1.28 1.44
Fe(IIl)lFe(II) 0.732 0.700 0.68
Ag(l)lAg( 0) 0.792 0.228 0.77
As(V)1As(III) 0.577 0.577
Ce(IV)/Ce(III) 1.70 1.28 1.44
Fe(IIl)lFe(II) 0.732 0.700 0.68
H(I)IH(O) -0.005 -0.005
Hg(II)IHg(1) 0.907
Hg(l)IHg(O) 0.776 0.274 0.674
Mn(IV)/Mn(II) 1.24
Pb(II)lPb(O) -0.14 -0.29
Sn(lV)/Sn(II) 0.14
Sn(Il)/Sn( 0) -0.16
Conductivities from Robinson and Stokes (B9) and Dobos (H9) in units of
10- 4 S m 2mol- 1, aqueous solutions at infinite dilution, 25 DC.
446 Appendix 3
448
Laplace Transform Methods 449
The Laplace tranform of a constant is
L(a) = als
Functions of variables other than t behave as constants in the
transformation
L[H(x)] = H(x)ls
The utility of Lapace transforms in the solution of differential equations
is that the transform of a derivative is a simple function
L[dF(t)/dt] = s f{s) - F(O) (AAa)
L[d 2F(t)/dt 2 ] = s2f{s) - s F(D) - (dFldt)o (A4b)
A short selection of Laplace transforms are found in Table All.
Two properties of the Laplace transformation are sometimes useful
in finding the inverse transform. The shift theorem allows the zero of s
to be displaced by a constant:
L-llfts + a)) = F(t) exp(~t) (A5)
The convolution theorem is useful when the inverse transformation of
f(s)cannot be found, but f(s) can be written as the product of two
functions, f{s) = g(s)h(s), the inverse transforms of which can be found.
If
G(t) = L-lfg(s)]
H(t) = L-l[h(s)]
then
(A.6)
t s-2 cos at s.
a 2 + s2
t n- 1 _a_
s-n sinh at s2-a 2
(n -I)!
~
l/'fii s-1I2 cosh at s2-a 2
_I_
2fiTiC s-3/2 exp at s-a
F(t) f{s)
exp(at) - exp(bt) 1
a-b (s-a)(s-b)
ifI
exp(at) erfW -IS(s - a)
1
exp(at)[l- erfYat]
We need initial and boundary conditions to solve eq (A,B), and these differ
from one problem to another.
Derivation of eq (3.22). Let us start with the problem posed in
§3.3. We considered a solution layered on pure solvent so that the initial
condition was C =C* for x < 0, C = 0 for x > O. The boundary condition is
C -7 C* as x -7 - 0 0 , C -7 0 as x -7 +00. We will divide the problem into two
regimes, - 0 0 < X < 0 and 0 < x < +00, with the requirement that C(x,t) and
J(x,t) be continuous at x = 0 for t > O. Thus for x > 0, we have
D a2c(x,S) () 0
sex,s =
ax 2
the general solution to which is
c(x,s) =A(s) exp(-YslD) + B(s) exp(+YslDx) (A.9a)
where A(s) and B(s) are to be determined from the boundary conditions.
One of the boundary conditions requires c(x,s) -7 0 as x -7 00 so that B(s) =
O.
For x < 0, eq (A,B) gives
D ic'(x,s)
dX 2
s
,
C (x,s) +
C*
= °
452 Appendix 4
Since the error function is an odd function of the argument, i.e., erf(-'I') =
-erf('I'), we see that these two functions are in fact identical:
C(x,t) =C*2 [1- erf 2,Dt
.~]
(3.22)
or
B(s) =- ~A(s) (AI4)
where
~= "IDo/DR
Setting x = 0 in eqs (All) and substituting in eq (A.I3a) gives
Co*/s + A(s) = eB(s) = - ~e A(s)
so that
A(s) =_ Co*
+ ~e)
s(1
Thus the transformed concentrations are
Co*
( ) = -s-
cox,s [1 exP{-Ys7i50x}]
1 + ~e
( ) - C* = D [d C(r,S) + .2.
2
scr,s rdC(r,S)]
--
dr 2 dr
This differential equation can be converted to a more familiar form by
the substitution
vCr,s) = r c(r,s)
2
s vCr,s) _ rC* = D d vCr,s)
dr 2
Remembering that the range of r is ro to 00, the solution analogous to eq
(AIO) is
Laplace Transform Methods 455
Reverting to the transformed concentration function, we have
c(r,s) = ~ + A~) exp [-VI (r-rol]
This expression is consistent with the boundary condition
C(r,t) -+ C*, c(r,s) -+ C*/s as r -+ 00
we get
i = nFACo*fl5Qiii(-L-..l..) t >t (4.26)
vt-t Yt
Derivation of eq (4.27). In a constant current experiment such
as chronopotentiometry, the flux of 0 at the electrode surface is constant
up to the transition time when Co(O,t) -+ 0 and the potential swings
negative. Thus the boundary condition is
Do OCo(O,t) ---.L
ax nFA
Taking the Laplace transform, we have
Do aco(O,s) =~
ax nFAs
=0, we have
Differentiating eq (A.10) and setting x
-Doh/Do A(s) =nJAs
Solving for A(s) and substituting in eq (A. 10) with x =0:
cO<O,s) =C0 * .
I
S nFAsYsDo
Taking the inverse transform, we get
CO<O,t) =Co* _ 2ifi
nFAVreDo
Apparently, Co(O,t) goes to zero when
nFA VreDo Co* =2ifi
so that the transition time is given by
fC = nFA VreDo Co*/2 i (4.27)
Derivation of eq (4.29). For a linear potential scan experiment,
the boundary conditions are similar to those used in deriving eq (4.2)
except that the potential is time-dependent. Thus we can proceed as
before up through eq (A.14), but A(s) cannot be so simply evaluated. We
can write the transformed current as
Laplace Transform Methods 457
Co(O,t) = Co. - 1
nFA Y7r.Do
t i(t) dt
10 yt-t
S ) K·V s
'V K% + Kki
Neglecting K compared with 1 and assuming that K 2 s «Kkl,l the
transformed flux of 0 is
;o(O,s) = -C* w s +fKki.
YKklS
Taking the inverse Laplace transform and converting to current, we
have
(5.31)
CR(O,s) = keCo*
VDR seA + -IS)
Taking the inverse transforms, we have
Co(O,t) = Co* - kcCo* (1- exp(A2t}[1- erf(A Vt)))
'A.VDo
sin rot
fit
t ~ cos
Jo ft
CJ)'t d't - ~ t ~ sin
fit Jo ft
CJ)'t d't
Thus we have
CO<O,t) = Co. _ 10 {sin rot - cos rot)
FAY2roDo (6.47a)
Equation (6.47b) for CR(O,t) results from a similar development.
5
-c>< DIGITAL SIMULATION
z
W
Il.
METHODS
Il.
<
462
Digital Simulation Methods 463
lJ.1 flux = 2.2 x 10-3 mole Na+ s-l; firing rate =2.2 s-l
L12 (a) E' =+0.043, +0.353, -0.154 V
(b) !lG' ;;; -8.3, -68.1, +29.7 kJ moP
(c) K =28, 8.6 x 1011, 6.2 x 10-8
2.14 E = 0.767 V
3.5 0.203
Answers to Selected Problems 469
4.8 Co = 0.14 mM
4.26 13 '" 4 x 10 12
5.11 (a) 60
(b) 1.3
(c) 5.1
7.6 126C
7.7 (a)0.234C
(b)48.5 J.!M
472
Author Index 473
Cruikshank, W. 371, 372, 377, 421 Gerischer, H. 432
Curran, D. J. 387, 422 Geske, D. H. 259,308
Dalrymple-Alford, 193, 238 Gibbs, O. W. 377,421
Damaskin, B. B. 427 Gileadi, E. 430
Daniell, J. F. 4 Given, P. H. 259,308
Davis, D. G. 181,238 Goldberg, I. B. 258, 308
Davy, H. 4 Gosser, D. K 266,309,465
de Bethune, A J. 431 Goto, M. 193,238
de Montauzon, D. 226,238,247,308 Gouy, L.-G. 60,61, 105
de Smet, M. 79, 80, 105 Grahame, D. C. 60,87,88,89,105
Debye,P.J. W.90,91,105, 124,147 Greef, R. 428
DeFord, D. D. 383,421 Gregory, D. P. 45,54
Delahay, P. 26, 54, 179, 181,237,292, Gritzer, G. 171,237
309, 427, 432 Gross, M. 247,308
Deryagin, B. V. 73, 105 Grove, W. R. 48, 49, 54
Despic, A R. 333,367, 369,415 Grunwald, R. A 236, 239
Dietz, R. 361, 367 Grzeszczuk, M. 365, 367
Dobos, D. 431, 445 Guggenheim, E. A 142, 147
Donnan, F. G. 142, 147, 158 Haber, F. 36, 54
Dordesch, K V. 45, 54 Hajdu, J. 391,422
Drazic, D. 333,367,369,415 Hall, C. M. 402
Dryhurst, G. 171, 237, 429 Hamaker, H. C. 69, 105
Dukhin, S. S. 73, 105 Hamed, H. S. 97, 105, lO7,427
Dvorak, J. 428 Harrar, J. E. 383, 421
Ebsworth, E. A V. 20, 54 Harris, M. D. 236, 239
Edison, T. A. 52 Hawkridge, F. M. 173, 237
Ehlers, R. W. 97, 105 Hawley, M. D. 247, 293,308,309
Eisenberg, M. 50, 54 Headridge,J.B.427
Engles, R. 394, 422 Healy, T. W. 70, 105
Enke, C. G. 113, 146 Heineman, W. R. 389,422,430
Erdey-Gruz, T. 318,367 Heinze,J.218,220,222,238
Erman,P.113,115,146,371 Helmholtz, H. von 60, 61, 105
Falkenhagen, H. 124, 147 Henderson, P. 139, 147
Faraday, M. 4,52,371,421 Hershberger, J. W. 305, 309
Faulkner,L. R. 204,238,266,309,428 Heubert, B. J. 322,367
Feldberg,S. W.286,293,294,303,304, Heydweiller, A 126, 147
309,462 Heyrovsky, J. 157,158, 165, 194,232,
Fick, A E. 130 237,239,429
Filinovskii, V. Yu. 428 Heroult, P. L. T. 402
Fillenz, M. 236, 239 Hibbert, D. 427
Flato, J. D. 232,239 Hills, G. J. 432
Fraenkel, G. K 259, 261, 308 Hitchman, M. L. 430
Frost, A. A 20, 54 Hittorf, J. W. 116, 117, 146
Frumkin, A N. 324, 367 Hogg, R. 70, 105
Fry, A J. 169,237,428,430 Hoijtink, G. J. 225,238
Fuoss, R. M. 122, 146 Holler, F. J. 113, 146
Furman, N. H. 39,54,175,237 Howell, J. O. 222,238
Furstenau, D. W. 70, 105 Huston, R. 107
Gale, R. J. 429 Huckel, E. 90, 91, 105
Galus, Z. 171, 237, 429 Ilkovic, D. 157, 158, 196,237,238
Galvani, L. 4 Ingram, D. J. E. 259,308
Gardner, A. W. 202, 238 Ishii, D. 238
Geary, C. G. 107 Israel, Y. 348,367
Geiger, W. E. 247,308,363,365,367 Ives, D. J. G. 430
474 Author Index
James, A. M. 427 Manecke, G. 327, 367
Janata, J. 39,54 Mann, C. K 169,237, 428, 447
Janz,G.J.43~431 Mantell, C. L. 430
Jefti~L.293,294,303,309 Marcoux, L. S. 286, 309
Jensen, B. S.299,309 Marcus, R. A. 316, 319, 367
Jones, R. D. 207,238 Mark, H. B., Jr. 389, 422, 428
Jordan,J.431,441 Martinchek, G. A. 261,309
Joslin, T. 465 Marzlutf, W. F., Jr. 281, 309
KekuIe, A. 396 Mattson, J. S. 428
Kern, D. M. H. 279,309 Mazorra,M.235,239
Khan, S. U. M. 428 McAllister, D. L. 171,237
Kim, S. 306, 309 McBreen, J. 45, 54
Kirowa-Eisner, E. 204, 234, 238, 239, McCollum, P. A. 448
245,430 McCormick, M. J. 253, 308
Kissinger, P. T. 430 McKinney, T. M. 258,308
Klemenciewicz, Z. 36, 54 Meites, J. 384, 421
Klingler, R. J. 305, 309 Meites, L. 194, 233, 238, 239, 348, 367,
Kochi, J. K 305,309 384,421,429,431
Kohlrausch, F. 114, 115, 126, 146, 147 Meng, Q. 306, 309
Kolbe, H. 395, 396,422 Milazzo, G. 431
Kolthotf, I. M. 207, 234,238, 429 Miller, J. W. 383, 421
Kornberg, H. L. 443 Milner, P. C. 53, 54
Kortum, G. 428 Mohammad, M. 391, 422
Koryta, J. 236,239, 427, 428, 429 Mohilner, D. M. 60, 105
Kosaka, T. 394, 422 Moraczewski, J. 363,367
Kosower, E. M. 391, 422 Morris J. R. 204,238
Koutecky, J. 278,309 Murray, R. W. 39, 54, 427
Krebs, H. A. 443 Myland, J. C. 341,367
Kublik, Z. 389, 422 Nagy, Z. 431
Kuta, J. 171,237, 429 Narayanan, A. 431
Kyriacou, D. K 430 Nernst, W. 11, 158,318
Laitinen, H. A. 207,238,392,422 Neto, C. C. see Camaioni-Neto, C.
LaMer, V. K 98, 105 Newman,J.220,238
Latimer, W. M. 18,431 Newman, J. S. 428
Le Bel, J. A. 396 Nicholson, R. S. 183, 187, 238, 296, 298,
Le Chatelier, H. 401 309, 348, 366, 367
Leclanche, G. 45 Nicholson, W. 4, 113, 371
Lehnhotf,N.S.215,238,266,309 Nishiguchi, I. 395, 422
Lemoine, P. 247,308 Nixon, F. E. 448
Leon, L. E. 233,239 O'Brien, P. 342, 367
Levich, V. G. 212, 238, 428 O'Halloran, R. J. 366, 367
Lewis, G. N. 18 O'Neill, R. D. 236, 239
Lingane,J.J.230,234,238,384,421,429 Ohkawa, M. 395, 422
Lingane,P.J.217,238 Oldham, K B. 193,215,217,238,274,
Loud, N. A. S. 431 309,341,367
Loveland, J. W. 128, 147 Onsager,L. 77,105,122, 123, 146
Lowenheim, F. A. 430 Orpen, A. G. 263, 309
Lund, H. 430, 431 Ostapczuk, P. 389, 422
Lyons, E. H., Jr. 427 Osteryoung, J. 204,234,238,239,245
MacDonald, D. D. 430 Osteryoung, R. A. 206,238, 390, 422
MacDonald, H. C., Jr. 428 Ostwald, F. W. 5,48,54, 156,428
MacInnes, D. A. 427 Overman, R. F. 56
Maki, A. H. 259, 308 Owen, B. B.427
Maloy, J. T. 462 Parker, V. D. 191,238,299,309
Author Index 475
Parsons,R.60,105,324,367,431,441 Selley, N. J. 427
Patriarche, G. J. 236,239 Shain, I. 183, 187, 190,238,289,293,296,
Peat, R. 428 298,309,389,422
Penciner, J. 430 Shannon, R. D. 121, 146
Peover,~.E.259,308,361,367 Shaw, D. J. 69,70,85,105
Perkins, R. S. 89, 105 Shedlovsky, L. 113, 146
Peter, L.~. 428 Shedlovsky, T. 113, 146
Pinson,J.303,309 Shimizu, K 390, 422
Plambeck, J. A. 429 Shono,T. 394, 395,422,430
Plante, G. 52 Siegerman, H. 236, 239
Pleskov, Yu. V. 428 Simon, A. C. 51, 54
Pletcher, D. 428,430,465 Skundin, A. ~. 430
Poilblanc, R. 226,238,247,308 Smith, D. E. 322, 356, 365, 366, 367
Pourbaix,~.26,54,431 Smith, ~. G. 448
Prada!; J. 236, 239 Somo~,Z.380,421
Prade~va,J.236,239 Soos, Z. G. 217,238
Purdy, W. C. 429 Spiro, ~. 116, 146
Reddy, A. 428 Srinivasan, S. 430
Reeves, R. ~. 60, 105 Steckhan, E. 394, 422
Reilley, C. N. 175,237,427 Stefani, S. 465
Reinmuth, W. H. 179,237, 259, 261, 290, Steihl, G. L. 292, 309
308 Stem, O. 60,61, 105
Revie, R. W. 431 Stock, J. T. 177,237
Riddick, J. A. 446 Stokes, G. G. 121
Riddiford, A. C. 213,238 Stokes, R. H. 428, 445
Rieger, A. L. 263, 309 Stone, N. J., see N. S. Lehnhoff
Rieger,P. H. 258,259, 261,263,266,308, SWrzbach, ~. 220, 222, 238
308,309,322,367,465 Streitwieser, A. 226, 238
Rifi, ~. R. 392, 422, 429 Stulik, K. 429
Riley, T. 427 Sweigart, D. A. 215, 238, 266,306,309,
Robbins, J. 427 342,367
Roberts, J. L., Jr. 430 Swift, E. H. 444
Robinson, J. 428 Szebelledy, L.380,421
Robinson, R. A. 428, 445 Tachikawa, H. 266, 309
Roe, D. K. 167,238 Tafel, J. 325, 326,367
Rogers, J. R. 391, 422 Talmor, D. 244
Rosair, G. ~. 263, 309 Tanaka, N. 377,421
Rossiter, B. W. 428 Testa, A. C. 290, 309
Rossotti, H. 429 Thiebault, A. 303, 309
Roston, D. A. 389, 422 Thirsk, H. R. 432
Rupp, E. B. 431 Thomas, U. B. 53,54
Rutgers, A. J. 79, 80, 105 Tilak, B. V. 430
Ruzic, I. 366, 367 Tilset, ~. 299, 309
Rysselberghe, P. van 26, 54 Tiselius, A. 84, 105, 117
Sack, H. 124,147 Tobias, C. W. 432
Safford, L. K. 350, 367 Tome'§, J. 158,237
Salkind, A. J. 432 Tomkins, R. P. T. 431
Sand,H.T.S.182,238 Tomlinson, C. 427
Saveant,J.~.303,308,309 Trasatti, S. 337, 367
Sawyer,D.T.233,239,430 Tremillon, B. 427
Schafer, R. 394, 422 Tropp, C. 280,309
Schwartz, W. M. 289, 309 Turner, J. A. 206,238
Scott, C. J. 263,309 Uhlig, H. H. 418,422,431
Seeber, R. 465 Underwood, A. L. 237,239
476 Author Index
van't Hoff, J. H. 396
Vandenbalck, J. L 236,239
Verhoef, J. C. 388,422
Vesely, J. 429
Vetter, K J. 327, 330, 367, 428
Visco, S. J. 308,309
Volmer, M. 318,367
Volta, A 4, 371
Vukovic, M. 206, 238
Waller, A G. 258, 308
Wawzonek,S.39,54,392,422
Weaver,M.J.350,367
Weinberg, N. L. 395, 422, 429,430
Weiss, D. 429
Weissberger, A. 428
Westcott, C. C. 429
Weston, E. 7
Whewell, W. 4
White, R. E. 431
Wien, M. 124, 147
Wiese, G. R. 70,105
Wiesner, K 278, 280, 309
Wightman, R. M. 215, 223,238
Willihnganz, E. 51,54
Winograd, N. 171, 237
Wipf, D. O. 215,238
Wise, J. A. 389, 422
Wopschall, R. H. 190, 238
Yeager, E. B. 124, 147,432
Yoshida, K 430
Zana, R. 124, 147
Zhan~ Y.266,309,342,367
Zoski, C. G. 215, 217,238,341,367
Zuman, P. 236,239,247,308,431
SUBJECT INDEX
477
478 Subject Index
at high frequency 124 diffusion 128-136
ionic conductivity 115 Fick's first law 130
of pure water 126 Fick's second law 131
relation to diffusion 134 flux 130
relation to mobility 119 random walk model 128-130
theory of 122-124 to a microdisk electrode 216-219
units 109 to a planar electrode 153-162
conductometric titrations 128 to a spherical electrode 162-165
convection 152 diffusion coefficient 130
r.d.e. forced convection 208 relation to conductivity 135
conventions relation to frictional coefficient 134
potential scale zero 12 relation to mobility 133-135
sign of current 151, 195 diffusion layer thickness 158, 195, 211
sign of potential 7 diffusion-limited current 152-164, 196
sign of work 5 digital simulation 462-466
writing half-cell reactions 11 Donnan membrane potential 141-143
corrosion 412-421 double layer
corrosion inhibitors 420 effect on
differential aeration 419 colloid stability 68
passivation 417-418, 421 electron-transfer rate 323-324
prevention of corrosion 420-421 interfacial tension 86-88
reaction of metal with air 416 Gouy-Chapman theory 59-68
reaction of metal with water 413-416 surface pH 72, 410
sacrificial anodes 420 thickness of 66, 92
corrosion potential 414 double layer capacitance 88-90, 110,351
Cottrell equation 156 effect on
coulometric titrations 387 chronoamperometry 180, 220
coulometry 380-388 cyclic voltammetry 185, 223
constant potential coulometry 380 polarography 197, 202-206
constant current coulometry 384 dropping Hg electrode 172, 194-201
current-potential curve 155-162 effect of potential on drop time 86
cyclic voltammetry 183-194 homogeneous kinetics at 272
adsorption effects 189-190 see polarography
capacitive charging current 185, 223 dry cell 45
CE mechanism 296 Ebsworth diagrams 20
EC mechanism 296-300 EC mechanism 256, 279, 288-291, 296-
EC' mechanism 300 300, 362-366
ECE mechanisms 301-308 EC' mechanism 256, 280-281, 292-293,
derivative presentation 190-192 300
irreversible,quasi-reversible 348 ECE mechanisms 256, 282-287,293-295,
ohmic potential drop 185 301-308
semiderivative presentation 192-194 Edison cell 52
time scale 250 EE mechanism 332-334, 342-345
Daniell cell 2 efficiency of fuel cells 47
dead-stop titration 388 electric migration 152
Debye length 66, 92 electrical circuits
Debye-Falkenhagen effect 124 a.c. bridge 112
Debye-Huckel theory 90-98 cell equivalent circuit 110, 351
comparison with experiment 95-98 current source 168
extensions from 102-104 electrometer 30
limiting law 95 constant current source 386
relation to conductivity 122 for measuring a cell potential 28, 29
dielectric constant 61, 169 potentiometer 29
table 446 operational amplifier 30, 166
Subject Index 479
current follower 165 electrolysis 391
ramp generator 167 analytical applications 376-390
voltage follower 165 current efficiency 374
coulometer 382 electroseparation 374-375
potentiostat 165 electrosynthesis 390-396
galvanostat 386 industrial processes 396-412
voltage integrator 167, 382 electrometer 30
Wheatstone bridge 112 electromotive force 6
electrocapillarity 86-88 electron spin resonance 258-264
electrochemical cell potentials electron-transfer rate 316-319, 321-324,
measurement of 27 360,366
electrochemical cells Frumkin effect 324
conductance cells 113 Marcus theory 319-322
for coulometry 380 electroosmosis 75
Daniell cell 2, 151 electroosmotic pressure 75
Edison cell 52 electrophoresis 82-85
for electrogravimetry 379 electrophoretic painting 400
electrolysis cells 5, 151 electroplating 398-400
equivalent circuit 110,351 electrorefining 403
fuel cells 47-50 electrosynthesis 390-396
galvanic cells 2 Kolbe hydrocarbon synthesis 395-396
lead-acid cell 51 oxidation of olefins 394-395
Leclanche cell (dry cell) 45 reduction of aromatics 392
mercury cell 46 reductive elimination reactions 391
nickel-cadmium cell 53 equation
production of aluminum 402 Boltzmann distribution law 62
production of Cl2 & NaOH 405 Butler-Volmer equation 318
production of lead tetraalkyls 411 Cottrell equation 156
silver cell 47 Debye-Huckel limiting law 95
silver-zinc cell 58 Einstein relation 135
sodium-sulfur cell 53 Fick's first law 130
storage batteries 50-54 Fick's second law 131
thermodynamics 5 Gibbs-Duhem equation 86, 101
three-electrode configuration 165 Henderson equation 140
for voltammetry 173 Heyrovsky-Ilcovic equation 157
Weston cell 7-12, 29 Ilkovic equation 196
electrochemical potential 134 Kohlrausch equation 114
electrodes Levich equation 212
anode 4 Lingane equation 230
auxiliary electrode 165 Nernst equation 10-13
cathode 4 Nernst-Einstein equation 135
in conductance cell 113 Nernst-Planck equation 134
for stripping analysis 389 Ohm's law 76
see indicator electrodes Ostwald's dilution law 126
see reference electrodes Poiseuille's equation 76
electrogravimetric analysis 377-379 Poisson equation 61
electrokinetic phenomena 73-80 Poisson-Boltzmann equation 63
electroosmosis 75 Sand equation 182
electroosmotic pressure 75 Stokes law 82
streaming current 73 Tafel equation 325
streaming potential 74 equilibrium constants
theory of 76-79 from cell potential data 14
zeta potential 75, 79-80, 83 from conductance data 125
from polarographic data 227-231
480 Subject Index
error function 131-132 rotating platinum electrode 207
exchange current 318 rotating ring-disk electrode 214
faradaic impedance 110, 351-356 static Hg drop electrode 206-207
Faraday's laws of electrolysis 371, 380 industrial processes 396-412
Fermi level 1 anodization 400
ferrocene electrophoretic painting 400
as a potential reference 171 electroplating 398-399
rate of oxidation 350 electrorefining 403
Fick's laws of diffusion 128 hydrometallurgical processes 403
Flade potential 417 organic syntheses 412
formal potentials production of
table 444 adiponitrile 409-410
Franck-Condon principle 319 alkalies & alkaline earths 403
free energy-oxidation state diagrams aluminum 401-402
19-21 chlorates and bromates 407
frictional coefficient 120 Cl2 and NaOH 404-407
relation to diffusion coefficient 134 fluorine 407
Frost diagrams 19-21 lead tetraalkyls 410-412
Frumkin effect 324 manganese dioxide 408
fuel cells 47-50 perchlorates 407
galvanic cells 2 potassium dichromate 408
Gibbs free energy potassium permanganate 408
of activation 316-317 infrared spectroscopy 265-268
from cell potential data 14, 224-227 ion-selective electrode 35-39, 143-146
relation to electrical work 5 ionic conductivity
Gouy layer 60 table 445
Gouy-Chapman theory 59-68 ionic radii
half-cell potentials crystal radii 121
table 438 Stokes law radii 120-121
half-cell reactions 8 ionic strength 63
half-wave potential 157, 212 kinetic current 274
correlation with IR frequencies 226 kinetic zones
correlation with MO theory 224 CE mechanism 275-276
stability constants from 227-231 EC' mechanism 292-293
table 233 kinetics of electron transfer 315-324
Hall-Heroult process 401 Kohlrausch law of independent ionic
Helmholtz layer 60 migration 115
heterogeneous rate constants 316 Kolbe hydrocarbon synthesis 395
Heyrovsky-IlcovU: equation 157 Laplace transforms 448-461
history of electrochemistry table 450
double layer theory 60 Latimer diagrams 18
conductivity 114-115 lead-acid cell 51
electrolysis & Faraday's laws 371 Lec1anche cell 45
origins 3-4 liquid junction potentials 2, 136-141
polarography 194 London force 69
hydrodynamic layer thickness 209 lyotropic series 71
hydrogen evolution kinetics 334-337 Marcus theory 319-322
indicator electrodes 171-172 mass transport rate constants 159, 211,
dropping Hg electrode 172, 194-201 218
glass electrode 35 maximum suppressor 200-201
ion-selective electrode 35-39, 143-146 mechanisms 247
microelectrodes 215-223 CE 256, 274-278, 287, 296
quinhydrone electrode 56 EC 256, 279, 288-291, 296-300, 362-366
rotating-disk electrode 172, 207-215 Ee' 256, 280-281, 292-293, 300
Subject Index 481
ECE 256, 282-287, 293-295, 301-308 overpotential (overvoltage) 317
EE 332-334, 342-345 oxidation state diagrams 19
bond cleavage 251 peak polarogram 183
electrophilic attack 252 pH measurements 40
multi-electron processes 249 phase angle 352
multi-step processes 328-334 planar diffusion 153-162
reactions of olefin radical cations polarography 176, 194-201,231-234
394 a.c. polarography 322, 356-366
rearrangement 253 adsorption effects 198-199
reductive elimination reactions 391 analytical applications 231-234
reduction of aromatic anodic waves 235
hydrocarbons, nitro and capacitive charging current 197,
carbonyl compounds 293 ID2-200
mechanistic data on CE mechanism 274
hydrogen evolution 334-337 criteria for reversibility 345
oxidation of current maxima 201
p-aminophenol291 differential pulse 204
ArCr(CO)2(alkyne) 263 EC mechanism 279
iron 333 EC' mechanism 281
(mesitylene)W(CO)3 266 ECE mechanisms 282
Mn(CO)3(dppm)CI 253 instrumentation 165-168
CpMn(CO>2L 305 irreversible waves 348
9-phenylanthracene 299 maximum suppressors 200
reduction of peak polarography 183
azobenzene 289 polarographic wave 157, 161
(COT)CoCp 363 pulse polargraphy 202
cyclooctatetraene 322 resolution 162, 205
triiodide ion 330 reverse pulse polarography 204
Mn3 + 327 sampled d.c. (tast) polarography 201
CpMn(NO)(CO)2 306 sensitivity 197,201-206
p-nitrosophenol 295 square wave 206
1,1,2,3,3-pentacyanopropenide 261 stability constants from 227-231
mercury cell 46 static Hg drop electrode 206
microelectrode voltammetry time scale 250
see steady-state voltammetry, cyclic total electrolysis in 198
voltammetry potential
microelectrodes 215-223 corrosion potential 414
electron transfer kinetics 338, 350 between dissimilar conductors 1
homogeneous reactions at 272 Donnan membrane potential 141-
mobility 143
electrophoretic mobility 82-84 electrochemical cell potential 6
ionic mobility 118-119 ferrocene as a potential standard
relation to conductivity 119 171
relation to diffusion coefficient 133 Flade potential 417
molecular orbital theory 224 formal potentials 17, 154
N ernst diffusion layer 158 half-cell potentials 11-16
Nernst equation 10-13 half-wave potential 157,212
nickel-cadmium cell 53 ion-selective membrane 37, 143·146
ohmic potential drop Latimer diagrams 18
in cyclic voltammetry 185, 191, 193 liquid junction potential 2, 136-141
in steady· state voltammetry 213, measurement 27
220,222 potential of zero charge 87-90, 324
Onsager reciprocal relations 77 sedimentation potential 82
osmotic pressure 142 standard reference half-cell 12
482 Subject Index
surface potential 64, 66 spherical diffusion 162-165,216-219
ultrasonic vibration potential 124 square scheme 263, 305, 307, 364
zeta potential 75, 79-80, 83 standard states 9,17,21,91
potential range for solvents 447 static Hg drop electrode 206
potential-pH diagrams 25-26 steady-state voltammetry 176, 212, 220
potentiometric titration 42-44 CE mechanism 274-278
potentiostat 165-166 coupled homogeneous reactions 269
predominance area diagrams criteria for reversibility 345
(Pourbaix diagrams) 26-27, 413 EC mechanism 279
r.d.e. voltammetry EC' mechanism 280-281
see steady-state voltammetry ECE mechanisms 282-287
stripping analysis 389 EE process 342-345
rate laws 247 irreversible/quasi-reversible 338
rate of electron transfer in time scale 250
oxidation of ferrocene 351 Stem model for double layer 60
reduction of Stokes' law 120
(COT)CoCp 363 Stokes' law radii 120-121
cyclooctatetraene 322 storage batteries 50-54
Mn porphyrin complex 342 streaming current 73
substituted stilbenes 361 streaming potential 74
reaction layer thickness· 251, 270-273 stripping voltammetry 389-390
reference electrodes 31, 170-171 supporting electrolyte 153, 170
Ag/AgCI electrode 32, 170 surface charge density 66, 86, 89
calomel electrode 31, 170 surface tension
hydrogen electrode 12 effect of potential on 86-88
Luggin probe 168 symmetry factor 319, 333
table of potentials 444 table
residence time 250, 338 biochemical half-cells 442-443
resistance 109 conductivities, mobilities and
resistivity 109 diffusion coefficients 136
reversibility formal potentials 444
operational definition 28 half-cell potentials 438-441
reversibility, criteria for ion atmosphere thickness 66
a.c. polarography 360 ionic conductivity 445
cyclic voltammetry 188, 348 Laplace transforms 450
steady-state voltammetry 158, 345 list of symbols 434-437
rotating platinum electrode 207 parameters for extended Debye-
rotating ring-disk electrode 214 Huckel theory 104
rotating-disk electrode 172, 207-215 physical constants 434
homogeneous reactions at 270 polarographic data 234
Sack effect 124 potential range 447
salting-out effect 69-71 reference electrode potentials 444
sedimentation potential 82 SI units 433
silver cell 47 solvent properties 446
silver-zinc cell 58 Stokes law and crystal radii 121
silver/silver chloride electrode 32, 170 Tafel plot 325-326
sodium-sulfur cell 53 tast polarography 201
solvents thermodynamics 5
choice of 168-169 thickness
properties 169 ion atmosphere thickness 66
table of properties 446,447 diffusion layer thickness 158, 195,
spectroelectrochemistry 257 210
electron spin resonance 258-264 hydrodynamic layer 209
infrared spectroscopy 265-268 reaction layer thickness 251, 270
Subject Index 483
time scales for experiments 250 diffusion coefficient 130
titration for variables used in text 434-437
amperometric titration 177 heterogeneous rate constants 159,
biamperometric titration 388 316
conductometric titration 128 kinematic viscosity 208
coulometric titration 385 resistivity 109
dead-stop titration 388 SI units 433
Karl Fischer titration 387 van der Waals attraction 69
potentiometric 42, 176 vapor pressure 169
Tomes:riteria for reversibility 158, 345 table 446
transfer coefficient 317-323, 324, 333 viscosity 169
transference numbers 116-118 table 446
Hifforf method 116-117 voltammetry
moving boundary method 117 differential pulse 204
transition state theory 316 linear potential sweep 183
transition time 182 square wave 206
transport impedance 355 see steady-state voltammetry, cyclic
transport processes 128 voltammetry
ultrasonic vibration potential 124 Warburg impedance 355
units Wheatstone bridge 112
cell potential and free energy 7 Wien effect 124
concentration 62 work
conductivity 109 electrical work 5-6
surface tension - area work 86
zeta potential 75, 79-80, 83