Sie sind auf Seite 1von 3

Theory of Chronopotentiometry with Current

Reversal for Measuring Heterogeneous Electron Transfer Rate Constants


Floyd H. Beyerlein a n d Richard S. Nicholson Chemistry Department, Michigan State University, East Lansing, Mich. 48823
Theory of chronopotentiometry with current reversal is used to show that a system which behaves reversibly for some given current density can be made to exhibit kinetic behavior at higher current densities, as indicated by a separation of cathodic and anodic quarterwave potentials. An equation is derived which relates this difference of quarter-wave potentials to the standard rate constant for electron transfer. The method appears to provide a rapid and simple way to evaluate electrode kinetics.

-300

- 200

CHRONOPOTENTIOMETRY reversal provides a useful with current qualitative estimate of reversibility of an electrode process. It should be possible with this technique to obtain quantitative estimates as well. Consider a redox system characterized by its heterogeneous rate constant, k,. For some given current density, k , will be large enough that electrochemical equilibrium will be established during the entire chronopotentiometric experiment. Under these conditions the system will appear Nernstian, and, for example, the difference of quarterwave potentials during the forward and reverse parts of the experiment will be nearly zero. However, with this same system if current density is progressively increased, a point will be reached where it is impossible for electrochemical equilibrium to be maintained. In this case the difference of quarter-wave potentials will be different from zero, and together with the current density at which this happens, should be a measure of ks. The purpose of this paper is to present equations with which exactly such correlations can be made.
THEORY

0.2

0.6

1 .o

\I ,

I .4

t/q
Figure 1. Theoretical chronopotentiograms showing effect of kinetic parameter, il., when equls 0.5 and R equals -1
(Y

Theory of chronopotentiometry with current reversal and electron transfer described by the electrochemical absolute rate equation is given by Anderson and Macero ( I ) ; their results provide the starting point of our treatment. In the following discussion we assume that only the oxidized form of depolarizer is present at the beginning of the experiment. Hence we consider reduction to occur initially with current density, iy, and label the first transition time, rY. After current reversal at ry, the current density is i,, and the second (oxidation) transition time measured from t = r Ywill be labeled, r7. Under these conditions, and with slightly different notation, Equation 4 of Reference ( I ) becomes
O < t I r y

The following definitions apply to Equations 1 and 2


=

iy(DR)a/2/nFksCo*(Do)OL/2
Y = tY 1 .

(3)
(4)
(5)

g(E) = (nF/RT) ( E - E d

illif

(6)

IJ exp[ag(E)]
ry

= 1

- y1/2- Y

1/2

expk(El1

(1)

5 t I rr
=

IJ Rexp[ag(E)]

+ (1 - R) ( y - 1)lj2 - y 1 / 2+
W 2- ~ ~ / 2 l e x ~ k ( E(2)l l

[(I - R) (Y -

(1) L. B. Anderson and D. J. Macero, ANAL. CHEM., 322 37,

(1965).

There Ellz is the conventional polarographic half-wave potential, k , the standard heterogeneous rate constant, a the transfer coefficient, DOand DRdiffusion coefficients of the oxidized and reduced forms respectively of depolarizer, CO*initial bulk concentration of depolarizer, and remaining terms have their usual meaning. Calculation of Potential-Time Curves. Because of the nonlinear form of Equations 1 and 2, except for some limiting cases discussed below, these equations cannot be solved explicitly for g(E), and therefore to calculate &E) as a function of y would require numerical solution of Equations 1 and 2. If the only object is to construct theoretical potential-time curves, however, it is simpler to consider y as the dependent variable. In this case Equations 1 and 2 can be solved explicitly for y

286

ANALYTICAL CHEMISTRY

Tf

5t 5
=

7 7

yl/z

-K(E)

+ { K (E)' + [(l - R)'


(1

- R)2 - 1

- 11 [(l - R)'

+ K(E)']}'/'
(8)

where

K ( E ) = (1 - R+exp[ag(E)1)/(1

+ expMEll}

Equations 7 and 8 can be used to construct theoretical potential-time curves without performing extensive numerical calculations. Potential-time curves calculated in this manner are shown in Figure 1. In general curves like those of Figure 1 depend on the three parameters +, a, and R . Effects of these parameters on potential-time curves are discussed in following sections. Effect of R. The parameter R affects only the anodic portion of potential-time curves (see Equations 7 and 8). In general R affects the shape of the potential-time curve slightly, but the major effect is in terms of T,. Often R would be - 1, but from an experimental point of view sometimes it is convenient to select R so that T and rr will be of similar magnitude. The exact relationship between T,-, r r ,and R is (2)
7,/7/

01 .

0.3

0.5
C Y

0.7

0.9

= 1/[(R

1)'

- 11

(9)

Except where indicated subsequent discussions are limited to R = -1. Effects of $ and (Y.Both $ and affect the potential-time curves of Figure 1, and in general these effects cannot be separated. However, some limiting cases exist where the relationship between g(E), R, +, and a can be stated explicitly. For example for small the redox system always is in equilibrium, and Equations 7 and 8 reduce to the following well known relationships which are independent of kinetic parameters ( I )
(Y

Figure 2. Dependence of AEon charge transfer coefficient, a

+,

A second limiting case arises when is sufficiently large that the processes for oxidation and reduction can be treated separately as the totally irreversible case. Equations 7 and 8 then reduce to the following familiar expressions ( I ) o<tsr,
ag(E) = ln(1 - y112) - In(+)
71 5

(13)

7r

O5tI71

(1 - &(E) = g(E) = ln[(l - yl/z)/yl/z] (10)

In(+> ln[[(l - R ) ( y

- l)l/'

- y1/']/R) (14)

71

5t5
=

71.

g(E)

ln([l

+ (1 - R ) ( y - l)lj2 yl/2]/[y1/2- (1

- R) (Y - 1>1'21} (11)

We find that potential-time curves are described by Equations 10 and 11 within a few millivolts whenever +is less than 0.01. For Equation 10 Eli2 occurs [g(E) = 01 when y = 0.25 (quarter-wave potential). For Equation 11 the value of y at which Eli2 occurs is given by
y g ( E )=0 =

Thus, for example when R equals - 1, from Equation 12 El/' occurs at y = 1,0716,or relative to r r ,when y-1 = 0 . 2 1 5 ~I~. ( ) Potentials corresponding to these two times (0.25 and yo(^) - 0 ) hereafter are referred to as E, and E,, respectively.
(2) R. W. Murray and C. N. Reilley, J . Electroanal. Chem., 3 182 ,
( 1962).

-1

+ { l + [(l - R)'

- 11 [4(1 - R)' 2[(1 - R)' - 11

+ l]]"' ]'

(12)

We find that potential-time curves are described within a few millivolts by Equations 13 and 14 whenever is greater than 2.5. Generally a affects potential-time curves in the expected < 0.01, curves are independent of (Y manner. Thus, for (Equations 10 and 11). For > 2.5, the effect of a is given by Equations 13 and 14. For values of between these limits, a affects both the symmetry of potential-time curves, and their position on the potential axis. This latter behavior is illustrated in Figure 2 where A E ( = E, - El) is plotted cs. CY for two values of Because for the mechanism being considered here CY is typically about 0.5, and rarely outside the range 0.3-0.7, these data of Figure 2 show that for reasonable values of a, the parameter AE tends to be independent of a, the dependence becoming less as decreases. The explanation of this fact is that as CY varies both E, and E, shift in the same direction, and these shifts tend to cancel in terms of AE. Nevertheless, for extreme values of a near 0 or 1, AE is markedly dependent on a . The reason for this effect is that, for example, as a approaches 1, E , tends to be independent of a (see Equation 13), whereas E, tends to vary exponentially with a (see Equation 14).

+.

VOL. 40, NO. 2, FEBRUARY 1968

287

AE as Measure of k,. In the preceding discussion it was concluded that for small values of $ (a quantitative discussion of the limits for $ is given below) and reasonable values of a, AE is independent of a. This fact is important for two reasons. First, when AE is independent of a, AE is determined uniquely by $, and therefore AE is a simple measure of $, and hence k, (see Equation 3.) Second, for the special case of a equal 0.5, Equations 7 and 8 can be solved explicitly for g(E). This means that an equation can be derived for AE which always is valid when a equals 0.5, and which, depending on $, may be exact for any between 0.3 and 0.7. These expressions for g(E) when (Y equals 0.5 take the following form
(Y

O_<t<7,

g ( E ) = 21n{[-$
T/

+ ($ + 4y1l2(1- y1/2))1/2]/2y1/2) (15)

5t
=

g(E)

I rr 2ln x

Limits of Applicability of Equation 17. It may be useful to consider the range of rate constants that can be measured by the above approach. The smallest value of k , that can be determined is set by the restriction $ < 1.0. This condition is equivalent to the statement k , 1 d 7 O O / 2 d 7 1 ,which for reasonable values of Do and 7 /(the maximum value of T / is determined by the time at which convection begins to interfere), recmisec. Of course duces to k , greater than about 4 x for systems having k , smaller than this value, one normally would apply Equation 13 directly without the use of current reversal. The maximum value of k , that can be determined is set by the restriction that $ must be greater than about 0.1 to detect kinetic effects, and the fact that double-layer charging determines the minimum transition time that can be measured experimentally. The condition $ > 0.1 is equivalent t o k , 5 d z / O . 2 6 / . Based on the discussion of Delahay (3), one concludes that transition times of the order of 10 mseconds can be measured under conditions where charging current is at most 1% of the total current. Although this value may be optimistic, it leads to the result that values of k, _< 0.3 cmjsec can be determined.
CONCLUSIONS

To calculate AE as it is defined above, Equation 15 is evaluated at y = 0.25, and Equation 16 is evaluated at given by Equation 12. The result is
AE = -In
($R2
($2

nF

+ 1) - $R + 1)Z - $

For Equation 17 to be generally useful it is important to define the errors that will result if Equation 17 is applied to a system where a is different from 0.5. We find that for AE about 95/n m V ( $ 4 . 0 ) , a value of $ calculated on the basis of Equation 17 always will be too small by about 10% if a is 0.3 or 0.7. For AE about 50jn mV ($-OS) the error is reduced t o about 6 z . Because in terms of rate constants an error of 10% is not large, and often within experimental error, we conclude that Equation 17 can be applied whenever AE is less than about 95/n mV. Thus, t o apply Equation 17 experimentally, conditions (current density, etc.) are selected which give AE less than 95/n mV, and then from this experimental AE, and a plot of Equation 17, $ is determined. If current density is known, k , can be calculated from Equation 4. Actually, to apply Equation 4 rigorously, (DR/DO)~/ must be known, However, except for the unusual case of very large is differences between Do and DR, the quantity very nearly unity, regardless of a.

Chronopotentiometry with current reversal appears to provide a rapid means of measuring heterogeneous rate constants. Both from experimental and theoretical points of view the method is probably the simplest yet developed. The major limitation is interference of double layer charging current, which sets the upper limit of rate constants that can be measured. For cases where charging current is not a limiting factor, the method should be very useful. An interesting result of the theory described here is that the kinetic parameter $ varies with Co* as well as with current density. Thus, in principle a given system at fixed current density may exhibit behavior ranging from nearly Nernstian to totally irreversible simply by changing Co* two orders of magnitude.

RECEIVED review September 20, 1967. Accepted Novemfor ber 17, 1967. Work supported in part by National Science Foundation under Grant No. GP-3830. Additional support received from United States Army Research Office-Durham, Contract No. DA-31-124-ARO-D-308.
(3) P. Delahay, New Instrumental Methods in Electrochemistry,.fl Interscience, New York, 1954, p 207.

288

ANALYTICAL CHEMISTRY

Das könnte Ihnen auch gefallen