Вы находитесь на странице: 1из 14

Electrochemistry Lecture 1 (This treatment follows Bard and Faulkner chapter 1)

In some ways, electrochemistry is a relatively easy thing to understand. You can apply a
potential form of some sort and measure current (the example of cyclic voltammetry is
given below), or you can apply a current and measure a potential. Interpreting
electrochemical experiments, however, can be quite complicated. We will focus
primarily on applying a potential form and measuring a current.

i
(-)
E
E (-)
0

t
Figure 1. Potential Program Used in Cyclic Voltammetry (left) and charging
current (right).
Electrical Potential and Electrochemical Reactions
First, we need to remind ourselves what potential means. Electrical potential is
the energy per unit charge required to bring a positive test charge from infinity to a
certain point. Consider bringing a test charge to a negatively charged surface (see the
figures below). There is an attractive force between the positive test charge and the
surface, so the energy required to bring the charge to the surface is negative (the particle
could do work on going to the surface), and we have a negative potential. Note that
charge separation gives rise to electrical potential.
+
+
+
+
+
+
+
Negative surface- negative
potential.
We can do external work by
bringing a positive charge to
this point. Electron energy is
high.

Positive surface-positive
potential.
We must exert external work
by bringing a positive charge
to this point. Electron energy
is high.

Figure 2. Concept of electrical potential.


Additionally, a negative potential will result in a high energy of a negative test
charge. Thus at high negative potentials, the electrons in an electrode have high energy.
In electrochemistry, the potential drop that we measure (actually we can only really
1

measure changes in potential drop because we have to use a reference electrode, but
nonetheless) is the electrode potential minus the solution potential. The more negative
this potential is the more likely reduction is to occur. Reduction takes place at a cathode.
Figure 3 illustrates the processes of electrochemical oxidation and reduction. At the open
circuit potential, no current can flow because we do not have a circuit. Thus, the energy
of the electrons in the electrode should be below the lowest unoccupied molecular orbital
energy and above the highest occupied molecular orbital (Figure 3, left). If we make the
potential more negative (Figure 3, upper right), the electrons will flow from the electrode
to the unoccupied molecular orbital and reduction will occur.
Electron
energy
level in
electrode

Electron Energy

Potential

LUMO
high

low

Electron
energy
level in
electrode

More Negative
Potential
HOMO

HOMO

Solution

Solution

LUMO

Electron Energy

Potential

LUMO
high

low

Electron
energy
level in
electrode

LUMO

More Positive
Potential
HOMO
Solution

Electron
energy
level in
electrode

HOMO
Solution

Figure 3. Schematic representation of reduction (top) and oxidation (bottom).


In contrast for oxidation, the open circuit potential starts in the same spot. If we
make the potential more positive, the electron energy in the electrode moves below the
energy of the occupied molecular orbital and oxidation will occur as electrons flow from
the redox couple to the electrode (Figure 3, bottom right). Lets look at a few specific
examples given in Bard and Faulkner. Suppose we put a Pt electrode in a solution
containing 0.01 M each of Fe3+, Sn4+, and Ni2+ in 1 M HCl. We use a normal hydrogen
electrode as the counter electrode. We can make a plot of standard electrode potentials
for each of the possible reactions as Figure 4 shows. Note that we can only have
reduction of each of these species because we did not add the reduced half of the redox
couple to the solution. Where will the open circuit potential be? For open circuit, we
have zero current. Thus the potential drop at the electrode solution interface has to
greater than 0.77. Otherwise the electron energy in the electrode would be sufficient to
reduce iron. The potential should also be low enough that we dont oxidize water. Thus
it should be between 1.23 and 0.77. When we move the electrode potential more
negative, the first species reduced will be Fe3+. Note that if the electrode is not at the
proper potential for open circuit, we simply see transfer of a few electrons between the
electrode and a redox couple to achieve the open circuit potential. A very small amount
of charge transfer leads to a change in the electrode-solution potential drop.

negative

Eo vs
Hydrogen
electrode

negative

-0.20

Ni2++ 2e- Ni

0.00

2H++ 2e- H2

0.15

Sn4++ 2e- Sn2+

0.77

positive

Eo vs
Hydrogen
electrode

Fe3++ e- Fe2+

2 H2O 4e- +O2+4H+

Figure 4. Appropriate potentials for a Pt


electrode in a solution containing Ni2+, H+,
Sn4+, and Fe3+.

0.54

Oxidation of Sn2+
2I-I2 + 2e-

0.77

Fe2+ Fe3++ e-

1.23

2 H2O 4e- +O2+4H+

positive

Open circuit potential


1.23

0.15

2H++ 2e- H2
Open circuit potential
Sn2+ Sn4++ 2e-

0.00

Figure 5. Appropriate potentials for a Au


electrode in a solution containing Sn2+, H+,
I-, and Fe2+.

Now consider the example in Figure 5. We have an electrode in a solution


containing Sn2+, H+, I-, and Fe2+. Note that in the potential range of interest, we only can
oxidize Sn2+, I-, and Fe2+. We can reduce H+, however. Thus, the open circuit potential
will be greater than the potential for reduction of H+ and less than the potential for
oxidation of Sn2+. If we go negative of this potential, we will see H2 evolution and if we
go positive of this potential, we will initially see Sn2+ oxidation.

Electrode 2

Electrode 1

How Do we Make Electrochemical


Measurements
Typically in electrochemistry we
like to talk about the potential drop at a
single electrode. However, we cannot
make a measurement of even open circuit
potential at a single electrode. We always
require a small amount of current to
measure voltage, and current requires a
complete circuit. Thus we need at least
two electrodes and this means we have at
least two potential drops, as Figure 6
shows. This is a problem because we can
only measure the potential difference
between the two electrodes. We do not
know the potential drops at the two
electrodes. Moreover, when we choose to
apply a potential drop, we can only apply
the drop between two electrodes. The
best we can do is to keep the potential
drop at one of the electrodes constant and
measure changes in potential at the other

Voltmeter

Potential

Measured
potential
difference

Potential
drop at
electrode 1

Potential
drop at
electrode 2

Figure 6. Potential drops in a two electrode


cell. We can only measure the difference in
potential, so we do not know the individual
potential drops.

electrode. If this is possible we can measure changes in


potential drops at a given electrode, but we cannot
Real
i
ref electrode
measure absolute potential drops.
Suppose in Figure 6, we want to measure how
current varies as we change the applied potential across
E
electrode 1. To do this, we need electrode 2 to behave
Ideal
as shown in Figure 7. As the current, i, changes equally
ref electrode
through both electrodes, the potential should not change
at electrode 2. When we can apply a current without
Figure 7. Current-voltage
changing the potential at an electrode, we term the
characteristics of ideal (solid
line) and real (dashed line)
electrode nonpolarizable (i.e. we see no polarization of
reference electrodes.
charge to create an additional potential
drop). Thus we can measure how the
Potentiostat
current varies with changes in potential at
electrode 1.
reference
working
counter
Note, however, that at high currents
electrode electrode
electrode
even the best references electrodes start to
show changes in potential. To overcome
this challenge, we employ the three
solution
electrode cell shown in Figure 8. The
potentiostat controls the potential between
Potential drop
the working and the reference electrode.
working electrode
Nevertheless, we measure the current
Potential drop
E
Potential drop
between the working and the counter
reference electrode
counter electrode
(constant)
electrode. The current through the
reference electrode is negligible compared
position
to the current through the working
electrode because we need only very small
Figure 8. Schematic diagram of a three electrode
currents to measure voltages. Thus, this
system and the accompanying potential drops.
current does not disrupt the measurement
of current between the working and counter
electrode. The potentiostat essentially applies a potential
Ag wire
between the working and counter electrodes such that the
potential between the working and reference electrodes is
the desired value. There is a potential drop at the counter
electrode that maintains the charge transfer at this electrode
AgCl
to complete the circuit, but we do not determine this
potential drop. Using such a setup, we can see how the
current varies with the change in potential between the
3 M NaCl
electrode of interest and the reference electrode.
How does one make a reference electrode?
Consider the most common reference electrode, which is a
Ag/AgCl electrode (Figure 9). The compartment, which is
Glass frit
separated from the solution by a slightly permeable frit,
contains a high concentration of Cl . For current to flow,
Figure 9. Ag/AgCl reference
electrode.

we must generate some AgCl or some Ag(0), depending on the direction of the current
(see equation 1).
(1)
AgCl e Ag Cl
However, generation of these species represents no change in the potential of the cell
because both are solids. We will create or remove a little free Cl-, which could change
the potential of the cell, but the amount generated or removed will be insignificant in
comparison to the amount of Cl- in the compartment. Thus, the potential of the electrode
is constant unless large amounts of current pass. To maintain the concentration of Cl- in
the reference electrode constant, we store it in a 3 M KCl or saturated KCl solution.
Zn wire

Ag wire

AgCl

Zn2+, Cl-

Notation for Electrochemical Cells


We frequently employ a shorthand notation for
electrochemical cells. Consider the system in Figure
10. We have a Zn and a Ag/AgCl electrode. The
shorthand notation is
Zn/Zn2+, Cl-/AgCl/Ag

Figure 10. An example


electrochemical cell.

H2

Pt wire

Ag wire

AgCl

The / represents a phase boundary and the comma


denotes components in the same phase. The standard
of all reference electrodes is the NHE (normal
hydrogen electrode) shown on the left of Figure 11.
Note that the activities of H2 and H+ are 1 for this
electrode. The Ag/AgCl electrode in saturated KCl has
a potential of 0.197 V versus NHE. In shorthand
notation, we would write this cell as
Pt/H2/H+, Cl-/AgCl/Ag

H+, Cl-

Introduction to Faradaic and Charging Current


Weve briefly considered potential. Lets next
Figure 11. An electrochemical cell
consider the current. Current is simply the amount of
with Ag/AgCl and normal hydrogen
charge passing per unit time. For electrodes that are
electrodes.
bigger than the micron scale, current generally
+
scales linearly with electrode area. Above we
E
considered that current would flow due to charge
+
l
+
transfer between an electrode and a redox couple in
e
c
solution. We term this Faradaic current. The
+
t
+
magnitude of Faradaic current depends on several
+
r
factors, so we will temporarily postpone discussing
o
+
this.
d
+
Lets consider the simpler case of applying a
e
+
potential and measuring current when no redox
couple is in solution. No electrons transfer between
the electrode and the solution at the given potential,
Figure 12. The electrochemical double
and ions dont pass into the electrode so we have a

layer, which behaves similarly to a


capacitor.

capacitor. The electrode forms one plate of the capacitor and the ions in solution form
the other plate of the capacitor as shown in Figure 12. Note the excess charge in solution
that forms one plate of this capacitor. This is what we call an ideally polarized electrode,
i.e. the electrode becomes charge polarized to create a potential without charge transfer
In ideal cases, the charge on a capacitor, q, is proportional to potential, E, with the
capacitance, C, being the proportionality constant.
q = CE

(2)

Typical values of an electrochemical capacitance are around 20 F/cm2. (A Farad is a


Coulomb per V.) Consider applying a potential step to make the electrode have a
negative potential with respect to the solution. To charge the capacitor current must flow,
bringing electrons to the surface of the electrode and cations to the solution/electrode
interface (we call this charging or nonFaradaic current).

Ag wire

AgCl
Hg drop
K+, Cl-

CHg Rsolution CAg/AgCl

Equivalent Circuits and Charging


Current
In many instances we can
model electrochemical systems with
equivalent electronic circuits.
Subsequently, we relate the equivalent
circuit elements to physically relevant
parameters of the system. Lets look
at one of the simplest examples:
(Figure 13) a mercury drop electrode
(this is basically an ideally polarizable
electrode) in a circuit with a reference
electrode (essentially a nonpolarizable
electrode). Figure 13 shows the
equivalent circuit for the electrode.
Because no charge transfer can occur,

Figure 13. An electrochemical cell containing a


hanging drop Hg electrode and a Ag/AgCl electrode.

each of the electrodes behave similarly to a capacitor,


but there is resistance to current flow in the solution.
The capacitance of the Ag/AgCl electrode is large so
we can neglect it. (Capacitances in series add
reciprocally). We have a simple series RC circuit.
Lets connect the circuit to a voltage source
(Figure 14). Once we close the switch, charging is
not instantaneous because of the solution resistance.
When we close the circuit, how quickly does the
capacitor charge? Using the loop theorem, we get
equation (3) where E is the battery voltage, q is
6

+
-

Figure 14. A series RC circuit.

charge on the capacitor, and R is the solution resistance.


(3)
Initially, there is no charge on the capacitor and all of the potential drop occurs across the
resistor. With a fully charged capacitor, there is no current because the potential drop
across the capacitor is equal to that across the battery so there is no potential drop across
the resistor. However, at times in between no charge on the capacitor and full charge,
the current is not constant. To determine the transient current, we note that i=dq/dt. Thus
we obtain

= + or =
(4)

Separating variables and integrating we get

(5)
0 = 0 = ln ( ) |0
= + /

Finally, evaluating from 0 to q as shown above, we get

= ln = ln(1

(6)

Taking the exponent of both sides, we get equation (7). This equation shows the charge
exponentially approaches CE, and at t=0, the charge is 0.

exp ( ) = 1 or = (1 exp ( ))
(7)

Now that we know q as a function of time, we can easily differential to get the current.

= = exp ( ) = exp ( )
(8)

Thus, after applying a potential to a


capacitor, the current decreases
q
i
exponentially with time (Figure 15). This is
an important point in electrochemistry where
we apply potential steps and the capacitor is
the electrical double layer. The current
t
t
decreases rapidly, but there is a transient
Figure 15. Evolution of current (left) and
current. The lower the resistance, the faster
charge on the capacitor (right) after closing the
the capacitor charges. At long times, the
switch in the RC circuit of Figure 3.
capacitor behaves as an open circuit, but
when potential is changing, capacitors allow
passage of current. This becomes especially important with ac voltages because the
magnitude of the voltage is always changing.
We just discussed a potential step where the voltage increases instantaneously to a
constant value. What if, however, the value of E increases linearly with time as
illustrated at the beginning of these notes for cyclic voltammetry (see Figure 1)? In this
case,
= =

(9)

where v is the voltage ramp rate in V/sec. The voltage will still be equal to the sum of the
voltage drops across the capacitor and the resistor. This equation is a little harder to
solve, but the solution gives

= (1 ( ))

(10)

Note that at long times, = . This is the steady state solution. You can get this just by
remembering q=CE. Differentiating with respect to t gives

dQ
dE
C
dt
dt

i C

dE
dt

( is thescan rate)

Thus a CV for a solution without electron transfer looks like that in the right of
Figure 1. To calculate the capacitance of an electrode with the steady-state charging
current:
ich arg ing

Most electrodes have a capacitance of about 20 F/cm2. Using this value, you can
estimate whether capacitance will be important in one of your cyclic voltammetry
experiments. Consider that you have a disk electrode with a 3 mm diameter. The
electrode area is (0.15cm)2 or 0.0707 cm2. The charging current for a 0.1 V/sec scan
C
V
C
rate will be i 20
* 0.0707 cm2 * 0.1
0.141
0.141amps
2
sec
sec
Vcm
Remember that a Farad is a C/V and an amp is a C/sec.
Faradaic processes
We need to be aware of charging current, but usually we want to ignore it because
we are more interested in studying electron-transfer reactions at the electrode surface.
When electron transfer occurs, we deem the reactions Faradaic. Lets first define a few
terms as they relate to Faradaic processes.
Galvanic cell- spontaneous reactions occur (battery discharge)
Electrolytic cell- external potential drives a reaction (plating, electrorefining)
Cathode- electrode at which reduction occurs
Cathodic current- reduction current (this term is valid when we are talking about one
electrode)
Anode- electrode at which oxidation occurs
Anodic current- oxidation current (this term is valid when we are talking about one
electrode)
Types of experiments (what variables are there?)
Potentiometric- Measure potential with no current (this is independent of electrode area
and often unaffected by kinetic parameters)
We can also perturb the electrochemical system- control voltage and measure current or
control current and measure voltage.

In electrochemical reactions, current, i, is a measure of the reaction rate. = . For


example, if we are plating a metal, equation (11) describes the moles plated, N, where F
is Faradays constant (units of Coulombs/mole of electrons), Q is the coulombs of charge
applied to plating, and n is the moles of electrons required to plate one mole of metal.

=
(11)

Differentiating, we get

= =
(12)

Thus the moles plated per unit time is proportional to current (at least the current that
went to the plating reaction). The important point is that current is a measure of the rate
of electrochemical reactions.
You may remember that in homogeneous kinetics, we define a rate for a reaction AB

as = where is the concentration of A. To get this rate we are simply


normalizing

by the volume because

is proportional to the reaction volume. In

electrochemistry, however, is not proportional to volume, but to electrode area. Thus,


in electrochemistry for the reduction of A,
1

= =
(13)

electrode

Now that we know that currents and


O*
rates are related (assuming all the
electrons go to the reaction of
O
interest), what controls Faradaic
nediffusion
Bulk current? Consider the reaction +
shown in Figure 16, which
shows how current due to electron
R*
R
transfer originates. For reduction, the
oxidized species O must diffuse to the
electrode-solution interface and
exchange an electron with the surface.
double
The amount of current can be
layer
controlled by either the kinetics of the
Figure 16. Mechanism of reduction at an electrode.
electron transfer or by the rate of
diffusion to the surface. In more
mathematically difficult cases, reaction rates are determined by a combination of
diffusion and kinetics.
Mass Transfer-Controlled Reactions
Lets consider a very fast electrode reaction. Thus at the surface, electron transfer
can occur so fast that the Nernst equation applies.
Fe(CN ) 36 e Fe(CN ) 64
RT O
EE '
ln
R Fe(CN ) 64
nF R
O Fe(CN ) 36
We call these reactions reversible or Nernstian because the surface obeys the Nernst
equation. (However, we are not at equilibrium because there is a concentration gradient
in the solution.) The rate of reaction depends on how fast O can get to the surface and R
can leave because upon reaching the surface because O is immediately reduced to
maintain equilibrium. There are three ways that a molecule can move toward the surface

1. Diffusion: movement of a species under a chemical potential gradient (i.e. a


concentration gradient)
2. Migration: movement of a charged species in an electric field (gradient of electrical
potential).
3. Convection: Stirring or hydrodynamic transport. As the solution moves, it moves the
dissolved molecules.
In one dimension, equation (14) describes transport,

=
+ ()
(14)
2
where for species i, is the flux (mole/cm /s), is the diffusion coefficient, is the
concentration, is the charge, is the electrical potential, () is the solution velocity,
is the Faraday constant, R is the gas constant, and T is temperature. The different terms
on the right hand side of the equation represent diffusion, electrical migration, and
convection from left to right. Usually we design experiments so one of the modes of mass
transfer dominates. This greatly simplifies the treatment of mass transport. Hopefully,
we can talk about the individual modes of transport more later on. In this course, we will
only discuss the case of where diffusion is the dominant term. Lets get back to our
reaction

Fe(CN ) 36 e Fe(CN ) 64
R Fe(CN ) 64

Concentration of O

O Fe(CN ) 36
As potential becomes more and more negative the concentration of O at the surface
decreases due to reduction. This creates a flux of O to the surface as the bulk
concentration is greater than that at the surface. Flux is defined as the amount of a
species arriving at a surface area in a given time (units of moles/(cm2sec)). Because
whatever comes to the surface by diffusion is immediately reduced, and nF converts
moles to coulombs (F=96485 C/mole of electrons).
C*
(15)
i nFAJ O
The negative sign arises because flux to the
electrode is in the negative direction. We are
defining reduction current as positive.
How can we determine JO if diffusion is the mode of
mass transport?
=

(16)

If we assume the concentration profile in Figure 17


for species O (a linear profile where C* is the
concentration of O in the unperturbed bulk of the
solution), equation (16) becomes
=

(=0)

x or distance from electrode

Figure 17. Assumed linear


concentration profile for species O
during its reduction at an electrode.

(17)

10

where ( = 0) is the concentration of O at the electrode surface. (Note we are


assuming a linear concentration profile, which is only an approximation.) If we stir the

solution or rotate the electrode the value of reaches a steady-state value. Defining

as mO, we get = ( ( = 0)). (Actually, mO is more general than and


does not require the strictly linear concentration gradient.)
Finally, we get
= ( ( = 0))
(18)
This is the current in a stirred solution with a constant mass transfer coefficient. This
looks good, but we do not yet know ( = 0). To overcome this, we need the Nernst
equation. However, the Nernst equation assumption requires values for both ( = 0)
and ( = 0). At steady state, the flux of O to the surface must equal the flux of R to
the surface. So
= ( ( = 0) )
(19)
Note that the highest current we can obtain occurs when ( = 0) = 0. (You cant
have a negative concentration.) We call this highest current the cathodic limiting
current, .
=
(20)
Solving equation (18) for ( = 0), we get

( = 0) =

(21)

For anodic limiting currents, which occur when ( = 0) = 0


=
Thus, similar to equation (21), we can write

(22)

( = 0) =

(23)

Using the Nernstian conditions for the surface we have,

(=0)

= + (=0)

(24)

Substituting for the surface concentrations, we get

= +

= + +

(25)

This equation links E and i. It is not especially intuitive, but we can see that as i
approaches , then

will become very small and E will be highly negative. As i

approaches , then E will be highly positive. If limiting currents are the same, then

when i=0, E = + . This would be the case if the concentrations and mass

transport coefficients are the same for O and R, so E = . Figure 18 shows a plot of i
versus E when mass transport coefficients are the same for both species, = 0.5, and
= = 10 . Note that even when we go to very high potentials, we cannot
increase the magnitude of the current because of the diffusion limitation.

11

12

Current (mA)

10
8
6
4
2
0
1

0.8

0.6

0.4

-2

Potential (V)

0.2

-4
-6
-8
-10

Figure 18. Mass transport limited current for a system with

= 0.5, and

= 10 .

Suppose that the reduced species, R, is initially absent from the solution. This leads to
= 0, and we get

= + +

(26)

This yields the current-potential curve in Figure 19.

12

Current (mA)

10
8
6
4
2
0
1

0.8

0.6

Potential (V)

-2

0.4

0.2

-4
-6
-8
-10

Figure 19. Mass transport limited current for a system with

=
, = 0.5, = 10 , and = 0. (The reduced
species is absent in the bulk solution.
Example problem. Suppose you have a solution that contains 5 mM Ce3+ and 3 mM Fe2+,
Sketch a steady state voltammogram in a stirred solution. The ions are dissolved in 1 M
H2SO4.
First, we note that Eo is 0.77 for Fe2+/Fe3+ and 1.44 for Ce3+/Ce4+. We also have the
possibility of Fe2+/Fe0, but Eo is -0.409. We should reduce H+ before we reduce Fe2+,
12

Cathodic H+ reduction
current
Cathodic limit
1.2

0.7

0.2

-0.3
Potential

Fe2+ oxidation
Anodic
current
Ce3+

oxidation (adds
to Fe2+ oxidation
current)

Water oxidation
Anodic limit

Concentration of O

Figure 20. Steady state currents for


the oxidation of 5 mM Ce3+ and 3 mM
Fe2+ in a stirred solution.

Mass-transport Limited Current for a non-stirred


system. Above we discussed fast reactions that are
limited by mass transport in stirred solutions. In this
case, Figure 17 describes a steady-state
concentration profiles. However, the situation is
different in the absence of stirring. Without stirring
the boundary layer will grow with time (Figure 21)
as more and more O is reduced. This increasing
boundary layer thickness results in a current that is
inversely proportional to the square root of time.
t=0

C*

current

1.7

and we have a large excess of H+. If the mass


transfer coefficient are the same for the two ions, the
Ce3+ limiting current should be 1.67 times greater
than the Fe2+ limiting current. The Ce3+ current will
add to the Fe2+ current as Figure 20 shows.

Increasing
time

i(1/t1/2)

time

x or distance from electrode

These diagrams apply to the Case of a Large, Negative Potential Step with no stirring of the solution

Figure 21. Left- concentration profile of the oxidized species after application of a large negative
potential step in a quiescent solution. Right- evolution of the cathodic current with time after the
potential step.

For the non-stirred solution, equations (15), (16), and (17) apply, as current is still
proportional to the flux to the surface. However the boundary layer thickness, which we
term , now varies with time. Thus,
=

(=0)

and
=

(27)

()

(=0)

(28)

()

The average concentration in the boundary layer (for the linear gradient) is

+ (=0)

(29)
Thus the total amount of O that has been reduced at any given time,
is the
product of the boundary layer volume and the difference between the initial concentration
and the average concentration in the boundary layer. This gives equation (30).
2

NOreduced,

13

+ (=0)

(=0)

= () [
(30)
Differentiating, we get

] = ()

()
+ (=0)

(31)

Note that if we just consider the moles of O still present in the solution, NO, we get
() + (=0)

= 2
(31a)

However, we can also assume that O is reduced as soon as it gets to the surface by
diffusion to maintain equilibrium. This gives

= =

(=0)

(32)

()

Equating the two expressions for

()
(=0)

(=0)

, we get
(33)

()

Rearranging yields
() () = 2
We can integrate this equation assuming the () = 0 at t=0 to get
()2 = 4 or () = 2
Now we can substitute this expression into equation (28) to get
=

(=0)

(35)

1/2 (=0)
2

(34)

(36)

Note that the current is indeed proportional to 1/ as shown in Figure 21. If we apply a
large amplitude negative potential step, ( = 0) = 0. Under these conditions, we get a
limiting current.

1/2

=
(37)
2
We assumed that the concentration profile is linear, which is not quite right. If you
rigorously solve the concentration profile and determine the current, you get

1/2

(38)
2

These equation differ by a factor of or 1.13. Thus our simple approximation of a

linear concentration profile predicts correctly how the limiting current varies with time
and predicts the magnitude of the current to within 15%. We are capturing the important
phenomena.

14

Вам также может понравиться