You are on page 1of 11

Nernst equation

From Wikipedia, the free encyclopedia

Jump to: navigation, search

In electrochemistry, the Nernst equation is an equation that can be used (in conjunction with
other information) to determine the equilibrium reduction potential of a half-cell in an
electrochemical cell. It can also be used to determine the total voltage (electromotive force) for a
full electrochemical cell. It is named after the German physical chemist who first formulated it,
Walther Nernst.[1][2]

Contents
[hide]

 1 Expression
 2 Nernst potential
 3 Derivation
o 3.1 Using Boltzmann factors
o 3.2 Using entropy and Gibbs energy
 4 Relation to equilibrium
 5 Limitations
 6 See also
 7 References
 8 External links

[edit] Expression
The two (ultimately equivalent) equations for these two cases (half-cell, full cell) are as follows:

   (half-cell reduction potential)

   (total cell potential)

where

 Ered is the half-cell reduction potential at the temperature of interest


 Eored is the standard half-cell reduction potential
 Ecell is the cell potential (electromotive force)
 Eocell is the standard cell potential at the temperature of interest
 R is the universal gas constant: R = 8.314 472(15) J K−1 mol−1
 T is the absolute temperature
 a is the chemical activity for the relevant species, where aRed is the reductant and aOx is
the oxidant. aX = γXcX, where γX is the activity coefficient of species X. (Since activity
coefficients tend to unity at low concentrations, activities in the Nernst equation are
frequently replaced by simple concentrations.)
 F is the Faraday constant, the number of coulombs per mole of electrons: F = 9.648 533 
99(24)×104 C mol−1
 z is the number of moles of electrons transferred in the cell reaction or half-reaction
 Q is the reaction quotient.

At room temperature (25 °C), RT/F may be treated like a constant and replaced by 25.693 mV
for cells.

The Nernst equation is frequently expressed in terms of base 10 logarithms (i.e., common
logarithms) rather than natural logarithms, in which case it is written, for a cell at 25 °C:

The Nernst equation is used in physiology for finding the electric potential of a cell membrane
with respect to one type of ion.

[edit] Nernst potential


Main article: Reversal potential

The Nernst equation has a physiological application when used to calculate the potential of an
ion of charge z across a membrane. This potential is determined using the concentration of the
ion both inside and outside the cell:
ELECTROCHEMISTRY

3.1 The Nernst Equation

The Nernst equation describes the fundamental relationship between the potential applied to
an electrode and the concentration of the redox species at the electrode surface.1 If an
electrode is at equilibrium with the solution in which it is immersed, the electrode will have
a potential, invariant with time, which is thermodynamically related to the composition of
the solution. In solution, species O is capable of being reduced to R at the electrode by the
following reversible electrochemical reaction.

O + ne-  R

The Nernst equation relates the potential, E, which is applied to the electrode and the
concentrations of species O and R at the electrode surface:

E = E  + 0.0591 log [O]


n [R]

where E = potential applied to electrode

E  = formal reduction potential of the couple vs. reference electrode

n = number of electrons

[O] = surface concentration of species O

[R] = surface concentration of species R.

The variation in the ratio of the [O] to [R] as a function of potential is the basis of all
voltammetric methods. The Nernst equation describes the relationship for reversible
equations. In other words, those are systems for which the electrode reaction in the

O + ne-  R is rapid in both directions. Surface concentration responds instantaneously to


any changes in potential (Table 2).2

3.2 Cyclic Voltammetry

Early contributions to cyclic voltammetry were made by investigators including Randles,3


Nicholson and Shain,4,5 and Kalthoff and Tomsicek.6 Cyclic voltammetry is one of the most
versatile electroanalytical techniques for the study of electroactive species. Cyclic
voltammetry has the capability for rapidly observing redox behavior over a wide potential
range. Cyclic voltammetry involves the cycling of the potential of an electrode, which is
immersed in an unstirred solution, and measuring the resulting current. The controlling
potential applied across these two electrodes is an excitation signal. For cyclic voltammetry,
the excitation signal is a linear potential scan with a triangular waveform8 (Figure 14). A
cyclic voltammogram is obtained by measuring the current at the working electrode during
the potential scan. The voltammogram is a display of current versus potential.7

Figure 15 is a cyclic voltammogram of 6 mM K3Fe(CN)6 in 1 M KNO3. The scan was


initiated at 0.80 V (applied at a) vs. SCE in negative direction at 50 mV/s. The area of the
platinum electrode is 2.54 mm2. The cathodic current at b to due to the electrode process:

[FeII(CN)6]3- + e-  [FeIII(CN)6]4-

The cathodic current rapidly increases (b-d) until the concentration of [Fe(CN)6]3- at the
electrode surface approaches zero, and the current peaks at d. The current then decreases (d-
g) as the solution surrounding the electrode is depleted of [Fe(CN)6]3-. The scan direction is
switched to positive at -0.15V (f) for the reverse scan. Anodic current is generated (i-k)
when the electrode becomes a sufficiently strong oxidant, and [Fe(CN)6]4- can be oxidized by
the electrode process:

[FeII(CN)6]4-  [FeIII(CN)6]3- + e-

The anodic current increases until the surface concentration of [Fe(CN)6]4- approaches zero
and the current peaks (j). The current decays (j-k) as the solution surrounding the electrode is
depleted of [Fe(CN)6]4-. The magnitudes of parameters including the anodic peak current,
(ipa), cathodic peak current (ipc), anodic peak potential (Epa), and cathodic peak potential (Epic)
are elucidated from the cyclic voltammogram.

The formal reduction potential E  for an electrochemically reversible couple is:

E  = Epa +Epc
2

For a reversible redox couple, the number of electrons transferred in the electrode reaction
can be determined by the separation between the peak potentials:

 Ep = Epa - Epic  0.059


n

The Randles-Sevcik equation for the forward sweep of the first cycle is:

ip = 2.69 X 105n3/2AD1/2Cv1/2

where ip = peak current, A


n = electron stoichiometry

A = electrode area, cm2

D = diffusion coefficient, cm2/s

C = concentration, mol/cm3

v = scan rate, V/s

Furthermore, ip increases with v1/2 and is directly proportional to concentration. This


relationship becomes particularly important in the study of electrode mechanisms. The ratio
of ipa to ipc should be close to one; however, chemical reactions coupled to the electrode
process can significantly alter the ratio of peak currents:

ipa  1
ipc

3.3 Instrumentation

In cyclic voltammetry, a potentiostat applies a potential to the electrochemical cell and a


current to a voltage converter which measures the resulting current. The current is displayed
on a recorder as a function of the applied potential (Figure 16).2 Operational amplifiers are
circuit elements utilized for enforcing a controlled potential at an electrode or for controlling
the current through a cell (Figure 17).8 The characteristics of an ideal operational amplifier
most importantly include:

1. No current through amp...a high input impedance (Z)

ideally Z = 

really Z ~ 1012

2. E(-) = E (+) This is only true when the op amp is balance

Current loops cause ground problems when a series of instruments are connected to ground
separately (use of three-prong plugs). Current loop problems can be eliminated when a series
of instruments are connected to ground through one common source (use of three-plug
adapters for all instruments expect one). See Figure 18.

Figure 19 illustrates the basic potentiostat used for cyclic voltammetry which incorporates a
three-electrode configuration. The potentiostat applies the desired potential between a
working electrode and a reference electrode. The working electrode is the place where the
reaction being studied takes place. The auxiliary electrode, typically platinum wire,
generates the current (polarized) required to sustain electrolysis at the working electrode.
The potential of the working electrode is controlled versus a reference electrode such as a
Ag/AgCl electrode (Figure 20) which maintains constant potential (polarized).

The half cell constructed may be represented as:

Ag|AgCl (sat’d.), Cl- (1M)

for which the half-reaction is:

AgCl + e-  Ag + Cl- E = +0.2223V

Potential (V) vs NHE

The potential of the reference electrode is measured before and after each experiment.
During an experiment, the potential of the reference electrode should vary no more than
+10mV. The potential of the reference electrode can be easily measured by immersing the
reference electrode in a buffer solution of known pH which is saturated with quinhydrone
(Figure 21).

quinhydrone = QNH + 2H+ + 2e- 

for QNH:

E QNH = 0.699 - 0.0591(pH)

and

E ref = E QNH + 0.0591 log[ ][H+]2


2[]

or

E ref = E QNH - |E reading|

3.4 Ohmic Potential Drop

Also known as iR drop, ohmic potential drop is potential drop due to solution resistance-the
difference in potential required to move ions through the solution. The major effects of iR
drop in cyclic voltammetry include shift in peak potential, decrease in magnitude of current,
and increase in peak separation (Figure 22).9 The reactions of interest occur at the surface of
the working electrode. Controlling the interfacial potential, specifically, the potential across
the interface between the surface of the working electrode and the solution, is imperative.10
The iR drop effects will become more evident as the scan rate is increased due to increasing
current. However, increasing peak separation with increasing scan rate is also characteristic
of a quasi-reversible process-the rate of the heterogeneous electron transfer is slow
compared to the time scale of the experiment.11

The iR drop or kinetic effects can be distinguished by scan rate studies performed at two
different concentrations (the difference should be at least an order of magnitude).12 The
increase in peak separation will be the same if due solely to electron transfer kinetics. The
increase in peak separation will be less at the lower concentration if due to iR drop. The iR
drop can be minimized by using a three-electrode system. By adding a high concentration of
fully dissociated electrolyte to the solution, conductivity will be increased, thereby reducing
resistance. Furthermore, the reference electrode tip should be in close proximity to the
working electrode surface.

The current can also be decreased in an effort to minimize iR drop.13 In a reversible process
peak current is directly proportional to the square root of the scan rate; therefore, the use of
low scan rates will minimize the current. However, higher scan rates are required for
calculation of heterogeneous and homogeneous rate constants. For a reversible process, peak
current is also proportional to concentration, so concentrations on the order of 10-4-10-3 M
analyte solution should be used for cyclic voltammetry. Finally, decreasing the electrode
surface area will also serve to minimize iR drop effects. Ohmic potential drop can have a
significant influence on electrochemical measurements, and these effects need to be
considered during the interpretation of electrochemical data.14

3.5 Mass Transfer

Mass transfer is the movement of material from one location to another in solution.15 In
electrochemical systems, three modes of mass transport are generally considered including
diffusion, convection, and migration. Diffusion removes concentration gradients by the
movement of material from a high concentration to a low concentration. Migration is the
movement of charged species due to a potential gradient. Electrolysis is carried out with a
large excess of inert electrolyte in the solution so the current of electrons through the
external circuit can be balanced by the passage of ions through the solution between the
electrodes, and a minimal amount of the electroactive species will be transported by
migration. Convection is the movement of species due to mechanical forces and usually can
be eliminated on a short time scale.

A concentration-distance profile2 is shown in Figure 23 for voltammetry of a planar


electrode immersed in a stirred solution of species O (1mM). The x-axis is distance from the
electrode into solution, and the y-axis is concentration. The vertical line at distance = 0 is the
interfacial boundary between the electrode and the solution. The dashed horizontal line is the
concentration of R in solution, CR=O. The continuous horizontal line is the concentration of
O in solution, C0 = 1 mM. Three regions of solution flow can be identified. A thin layer of
stagnant solution having a discrete thickness  , called the Nernst diffusion layer, is present
immediately adjacent to the electrode surface. Turbulent flow comprises the bulk of solution.
Laminar flow, a nonturbulent flow in which adjacent layers slide by each other parallel to
the electrode surface, occurs in a region between the Nernst diffusion layer and the solution
bulk. Figure 23a illustrates the homogeneous concentration of O and R throughout the
solution and up to the electrode surface. Either no potential has been applied or a potential
has been applied which is sufficiently positive of E   O,R such that CO/CR > 1000. Figure 23b
illustrates the conditions when E = E   O,R and equal concentrations of O and R are present
at the electrode surface in order to satisfy the Nernst equation. Figure 23c profiles the
conditions when a potential has been applied which is sufficiently negative such that CO/CR <
0.0001, and effectively, the concentration of O at the electrode surface is zero. The curved
lines denote the gradual transition between stagnant and flowing solution.

3.6 The Electrical Double-Layer

The changing arrangement of ions, solvent, and electrons in the interphasial region near the
electrode surface has been the focus of considerable investigation.16 The electrical double-
layer is associated with an ideally polarized electrode which is an electrode in which no
charge transfer can occur regardless of the potential imposed by an external voltage source.
The specific nature of the structure and the interactions of the electrical double-layer should
be considered in the interpretation of electroanalytical data. Various models have been
proposed describing the interphasial region near the electrode surface.

Helmholtz envisaged a "double layer" in which the excess charge on the metal would be
neutralized by a monomolecular layer of ions of opposite charge to that on the metal phase.17
This model did not account for the possible specific adsorption or random motion of ions.
See Figure 24.

The Guoy-Chapan "diffuse-layer model" was derived depicting a distribution of net charge
near the electrode surface which diminishes as the interfacial region stretches toward the
bulk (Figure 25). The random motion of ions results in a diffuse layer of charge in which the
concentration of counter ions is greatest next to the electrode surface and decreases
progressively until a homogeneous distribution appears in the bulk electrolyte. Specific
adsorption of desolvated ions at the electrode is weak; therefore, the adsorbed inner layer of
negative charge only partially shields the positive charge at the electrode. As a result, a layer
of solvated anions is present adjacent in the diffuse layer.

Finally, the Gouy-Chapman-Stern model was proposed which accounts for the finite size of
ions (Figure 26). Specifically adsorbed anions which are desolvated maintain direct contact
with the electrode surface in the inner Helmholtz layer. Electrostatic and chemical
interactions between the electrode and the ion govern the nature of specific adsorption.
Furthermore, potential profiles as well as the kinetics of interfacial interactions may be
significantly altered as the result of specific adsorption. The inner Helmholtz plane (IHP) is
considered to pass through the center of specifically adsorbed ions. The metal-ion
interactions are strong, and the specifically adsorbed charge may exceed the positive charge
on the electrode effectively establishing a layer of solvated cations adjacent in the outer
Helmholtz layer. The outer Helmholtz plane (OHP) is considered to be the approximate site
for electron transfer. Nonspecifically adsorbed ions which have their primary solvation
shells generally reside in the diffuse layer extending some distance from the electrode
surface. The thickness of the diffuse layer is dependent upon ionic strength of the buffer, and
in dilute electrolyte, the diffuse layer can extend to 100Å. The nature of the diffuse layer can
have a significant impact on the rate of electron-transfer since the actual potential felt by a
reactant close to the electrode is dependent upon it. Interfacial interactions must be
considered in the treatment of electrochemical data even though the exact interfacial
structure near the electrode surface is not known.18,19

3.7 Ferricyanide

A three-electrode potentiostat with a Lucite cell of conventional design was used for all
experiments with provision for nitrogen bubbling through a solution volume of
approximately 1 mL. The working electrode was a tin-doped indium oxide film deposited on
glass (Donnelly Corporation). A new electrode was used for each set of determinations.
Pretreatment and cleaning of the working electrode and cell involved successive
ultrasonications in Alconox (5 minutes), ethanol (5 minutes), and distilled water (5 minutes).
The area of the working electrode was approximately 0.3 cm2 in all experiments. A Ag/AgCl
reference electrode and a platinum auxiliary electrode completed the cell. The reference
electrode was calibrated against saturated quinhydrone in the appropriate buffer before each
experiment.

All determinations were carried out under nitrogen (Grade 4.5 from Airco, Inc.) to prevent
oxygen interference. The buffer and sample solutions were purged with nitrogen for
approximately 10 minutes prior to the beginning of a new series of determinations. Between
each scan, initial conditions at the electrode were restored by gently stirring the solution and
then allowing approximately two minutes for the solution to come to rest before obtaining
the cyclic voltammogram. All scans were initiated in the negative direction. The initial
potential was set at 0.80 V and the scan limits at 0.80V and -0.12V. A background of the
supporting electrolyte was obtained prior to each experiment. All chemicals were ACS
reagent grade and used as received. All water used was purified with a Milli RO-4/Milli-Q
system.20 Graphs were drawn with a Hewlett-Packard x-y recorder (model 7015B) which
was interfaced to the potentiostat.

The effect of scan rate (v) on cyclic voltammograms of 1 mM K3Fe(CN)6 in 0.1M KNO3 has
been observed (Figure 27) at 10, 20, 50, 100, and 200 mV/s. The reversibility of the
electrochemical system is indicated by the separation of the peak potentials ( Ep) which is
independent of the scan rate v. According to the Randles-Sevcik equation, ipa and ipc increase
as v1/2 (Figure 28). A plot of this equation will yield a straight line, the slope of which can be
used to determine the diffusion coefficient (D in cm2/s).

Concentration affects the magnitude of the peak current. Figure 29 displays cyclic
voltammograms of 6, 8, 10, and 20 mM K3Fe(CN)6 in 0.1M KNO3 at 20mV/s. Figure 30
illustrates that peak current is directly proportional to concentration.

As reported in the literature, a formal potential is a working number that is dependent on the
conditions of the experiment. The effect of supporting electrolyte on the cyclic
voltammograms of 4 mM K3Fe(CN)6 in 1M Na2SO4 and 1M KNO3 is shown in Figure 31.

3.8 References

1. Bard, A.J.; Faulkner, L.R. Electrochemical Methods: Fundamentals and Applications.

New York: John Wiley & Sons, 1980.

2. Heineman, W.R.; Kissinger, P.T. Laboratory Techniques in Electroanalytical Chemistry.


New York: Deeker, 1984.

3. Randles, J.E.B. Trans. Faraday Soc. 1952, 48, 828-832.

4. Nicholson, R.S.; Shain, I. Anal. Chem. 1964, 36, 706-723.

5. Nicholson, R.S. Anal. Chem. 1965, 37, 1351-1355.

6. Kolthoff, I.M.; Tomsicek, W.J. J. Phys. Chem. 1935, 39, 945-954.

7. Kissinger, P.T.; Heineman, W.R. J. Chem. Ed. 1983, 60, 702-706.

8. Reilley, C.N. J. Chem. Ed. 1962, 39, A853.

9. Britz, D. J. Electroanal. Chem. 1978, 88, 309-352.

10. Wipf, D.O.; Wightman, R.M. Anal. Chem. 1990, 62, 98-102.
11. He, P.; Faulkner, L.R. Anal. Chem. 1982, 54, 1313A-1326A.

12. Oldman, K.B. Anal. Chem. 1969, 41, 1904-1905.

13. Oldman, K.B. Anal. Chem. 1972, 44, 196-198.

14. Evans, D.H. Acct. Chem. Res. 1977, 10, 313-319.

15. Bockris, J.O’M.; Khan, S.U. Surface Electrochemistry: A Molecular Level

Approach. New York: Plenum Press, 1993.

16. Bowden, E.F.; Hawkridge, F.M.; Blount, H.N. Comprehensive Treatise of


Electroanalytical Chemistry. Vol. 10 (S. Srinivasan, Y.A. Chizmadshev, J. O’M. Bockris,
B.E. Conway, E. Yeager, Eds.) New York: Plenum Press, 1985.

17. Greef, R.; Peat, R.; Peter, L.M.; Pletcher, D.; Robinson, J. Instrumental Methods in
Electrochemistry. New York: John Wiley & Sons, 1985.

18. Riger, P.H. Electrochemistry. New York: Chapman & Hall, 1994.

19. Riley, T.; Tomlinson, C. Principles of Electroanalytical Methods. New York: John Wiley
& Sons, 1987.

20. Van Benschoten, J.J.; Lewis, J.Y.; Heineman, W.R.; Roston, D.A.; Kissinger, P.T.

J. Chem. Ed. 1983, 60, 772-776.

You might also like