Sie sind auf Seite 1von 10

Analytical solutions of the planar cyclic voltammetry process for two soluble

species with equal diffusivities and fast electron transfer using the method of
eigenfunction expansions
Adib Samin, Erik Lahti, and Jinsuo Zhang

Citation: AIP Advances 5, 087141 (2015); doi: 10.1063/1.4928862


View online: https://doi.org/10.1063/1.4928862
View Table of Contents: http://aip.scitation.org/toc/adv/5/8
Published by the American Institute of Physics

Articles you may be interested in


A one-dimensional stochastic approach to the study of cyclic voltammetry with adsorption effects
AIP Advances 6, 055101 (2016); 10.1063/1.4948698

The thermodynamic and transport properties of GdCl3 in molten eutectic LiCl-KCl derived from the analysis
of cyclic voltammetry signals
Journal of Applied Physics 121, 074904 (2017); 10.1063/1.4976570

Ab initio molecular dynamics study of the properties of cerium in liquid sodium at 1000 K temperature
Journal of Applied Physics 118, 234902 (2015); 10.1063/1.4937910

Analytical solution of the Poisson-Nernst-Planck equations for an electrochemical system close to


electroneutrality
The Journal of Chemical Physics 140, 224113 (2014); 10.1063/1.4881599

Commentary: The Materials Project: A materials genome approach to accelerating materials innovation
APL Materials 1, 011002 (2013); 10.1063/1.4812323

Copper nano-clusters prepared by one-step electrodeposition and its application on nitrate sensing
AIP Advances 5, 041312 (2015); 10.1063/1.4905712
AIP ADVANCES 5, 087141 (2015)

Analytical solutions of the planar cyclic voltammetry


process for two soluble species with equal diffusivities
and fast electron transfer using the method
of eigenfunction expansions
Adib Samin, Erik Lahti, and Jinsuo Zhanga
Nuclear Engineering Program, Department of Mechanical and Aerospace Engineering,
The Ohio State University, 201 W 19th Avenue, Columbus, Ohio 43210, USA
(Received 8 July 2015; accepted 6 August 2015; published online 14 August 2015)

Cyclic voltammetry is a powerful tool that is used for characterizing electrochemical


processes. Models of cyclic voltammetry take into account the mass transport of
species and the kinetics at the electrode surface. Analytical solutions of these models
are not well-known due to the complexity of the boundary conditions. In this study
we present closed form analytical solutions of the planar voltammetry model for
two soluble species with fast electron transfer and equal diffusivities using the
eigenfunction expansion method. Our solution methodology does not incorporate
Laplace transforms and yields good agreement with the numerical solution. This
solution method can be extended to cases that are more general and may be useful
for benchmarking purposes. C 2015 Author(s). All article content, except where
otherwise noted, is licensed under a Creative Commons Attribution 3.0 Unported
License. [http://dx.doi.org/10.1063/1.4928862]

INTRODUCTION
Cyclic voltammetry is one of the main tools used to characterize electrochemical systems and
has a broad range of applications. The method is frequently used for analyzing electroactive species
and surfaces and for the determination of reaction mechanisms and rate constants. Investigators over
the years have sought analytical expressions and relationships to enable a theoretical interpretation
and understanding of experimentally recorded cyclic voltammograms and other electrochemical
characterization techniques.1–6
Bortels et al.7 found analytical solutions for the one dimensional steady state transport of ions
in an electrolyte between two planar electrodes. The solution method takes the implicit form of a
set of nonlinear algebraic equations that must be solved numerically with a small modification. The
authors’ solution method is general and can deal with any kind of overpotential relation at both the
anode and the cathode.
Molina et al.8 derived analytical expressions for the current/potential response and concen-
tration profiles for the reversible ion transfer at the interface between two immiscible electrolyte
solutions. The researchers exploited the assumptions of semi-infinite planar diffusion and excess
ligand in the organic phase. Moreover, all solutions were attained for the equilibrium case, which is
described by the Nernst equation.
Berzins and Delahay9 derived an equation for oscillographic polarographic waves correspond-
ing to the reversible deposition of an insoluble substance and compared their results with experi-
mental data. In the analysis, the Nernst equation was used to describe the kinetics and the Laplace
transform method was used.
White and Lawson4 presented solutions for the voltammetric deposition and dissolution of a
metal onto or from an electrode accounting for spherical effects as well as kinetics, uncompensated

a Author to whom correspondence should be addressed. Electronic mail: zhang.3558@osu.edu.

2158-3226/2015/5(8)/087141/9 5, 087141-1 © Author(s) 2015


087141-2 Samin, Lahti, and Zhang AIP Advances 5, 087141 (2015)

cell resistance, and submonolayer metal deposition. The authors used Laplace transforms to derive
Volterra integral equations for the current as a function of the voltage.
Lantelme and Cherrat3 explored analytical solutions to the cyclic voltammetry process. The
authors utilized the diffusion equation to study mass transport and then applied the Nernst equation
at the surface of the electrode to describe the kinetics. In their analysis, the authors used the Laplace
transform to obtain the solution and expressed the current as a Dawson integral.
Oldham and Myland10 derived mathematical solutions for the cyclic voltammetry of strong
adsorption such that the redox pair is confined to the surface. The authors accounted for the kinetics
through the Butler-Volmer equation but completely removed mass transport from the model and
this ultimately allowed for closed form analytical expressions. In that study, the investigators postu-
lated that the activity of each adsorbate is proportional to its fractional coverage and they explored
different degrees of reversibility. For the quasi-reversible case, numerical integration was required
for the implementation of the presented solutions.
In addition, Oldham and Myland,11 developed a semi-analytical method for describing the
current in cyclic voltammetry. The method relied on using the Laplace transform and a numerical
implementation of the Riemann-Liouville semi-integration. This simple and efficient method, which
is implicit in nature, was implemented with success to a variety of mechanisms including E, EE and
ECE with various degrees of reversibility. However, one of the deficiencies of the method was that it
could not handle homogeneous reactions of orders other than unity and in which case it is necessary
to assume equal diffusivities.
Eswari and Rajendran12 reported analytical expressions for the concentration in an EC reac-
tion for limiting cases of small and large reaction rates at the electrode and small and large time
intervals. In that investigation, the authors employed the diffusion equation in planar geometry to
describe the mass transport and the Butler-Volmer equation at the electrode surface to account
for the kinetics. The asymptotic expressions for the normalized current were derived by using the
Laplace transformation.
Furthermore, Eswari and Rajendran explored analytical solutions for the EC2 cyclic voltam-
metry model.13 In that study, the authors utilized He’s Homotopy perturbation method to arrive at
approximate analytical expressions for limiting cases of small and large time intervals.
In this study we describe a generalized methodology for generating analytical expressions for
planar cyclic voltammetry for two soluble species using the method of eigenfunction expansions.
The method does not use a Laplace transform and can be extended to more complicated processes.

PROBLEM FORMULATION
A general reduction process at the electrode can be represented as:
A + e− → B (1)
In cyclic voltammetry the potential is swept linearly with time from a starting position Ei where
species A is stable to another potential Ev at which species B is formed and then the potential is
swept back to Ei leading to the re-formation of species A.

 Ei − vt f or t ≤ t sw

E=
 Ev + v (t − t sw ) f or t > t sw (2)

where t sw = |E i −E
v
v|
is the switching time.
To demonstrate our method, the one-dimensional, planar diffusion is considered. The mass
transport of the two species, A and B, is described by the corresponding diffusion equations:

∂c A ∂ 2c A
= DA 2 (3)
∂t ∂x
∂cB ∂ 2c B
= DB (4)
∂t ∂ x2
087141-3 Samin, Lahti, and Zhang AIP Advances 5, 087141 (2015)

The first boundary is the surface of the electrode which is defined to be at x = 0. The concentrations
at the electrode surface are functions of the potential applied to the electrode and in general we
write:
c A (x = 0,t > 0) = f (E(t)) (5)
cB (x = 0,t > 0) = g(E(t)) (6)
Furthermore, and by conservation of mass, the amount of A lost at the electrode is exactly the
amount of B formed there and the diffusive fluxes of the two species there are equal and opposite:

∂c A ∂cB
DA = −D B (7)
∂ x x=0 ∂ x x=0

The upper spatial boundary is typically set to be:14

L = 6 D A t ma x

(8)
where t ma x is the is the maximum time the experiment will run for. Typical diffusion coefficients are
on the order of 10−5 cm2/s and the typical experimental time is about 10 s. The effects of diffusion
will not extend this far from the electrode during the experiment and the concentration of each
species at this distance is constant in time and is equal to the bulk value:
c A (x = L,t > 0) = c∗A ; cB (x = L,t > 0) = c∗B (9)
Moreover, the initial concentrations at the beginning of the concentrations are also assumed to be
equal to the bulk values:
c A (x,t = 0) = c∗A ; cB (x,t = 0) = c∗B (10)
We now seek to non-dimensionalize the system variables:
x Dj cj D At
X= ; dj = Cj = ∗ ; T = 2 ( f or j = A, B) (11)
ϵ DA cA ϵ
Here ϵ is the radius of the macrodisc electrode. Additionally, we define a dimensionless potential
and the dimensionless scan rate to be:14

F ( ) ϵ2 F  θ i − σT f or T ≤ Tsw
θ= E − E of ; σ= v ; θ (T) =


 θ v + σ (T − Tsw ) f or T > Tsw (12)
RT D A RT

where E of is the formal potential of the reaction, R is the gas constant, T is the temperature, and F is
the Faraday constant.
Therefore, the transformed problem assumes the form:

∂C A ∂ 2C A ∂CB ∂ 2CB
= ; = d B (13)
∂T ∂X2 ∂T ∂X2
With initial and boundary conditions:
C A (X,T = 0) = C A (X = X ma x ,T) = 1 (14)
c∗
CB (X,T = 0) = CB (X = X ma x ,T) = ∗B (15)
cA
∂C A
∂CB
= −d B (16)
∂X ∂ X

X =0 X =0

where X ma x = 6 Tma x . In this study, and for simplicity, we will assume that CB∗ = 0. However, the
results of the study can be easily extended to cases where CB∗ , 0.
As far as the electrode surface boundary is concerned, we will discuss two cases in this study.
The first case is when Nernstian equilibrium is attained at the electrode surface throughout the
087141-4 Samin, Lahti, and Zhang AIP Advances 5, 087141 (2015)

potential scan, then Nernst’s equation relates the surface concentrations of species A, and B to the
applied potential at the electrode, E:
( )
RT c A(0,t)
E(t) = E f +
o
ln (17)
F cB(0,t)
By rearranging equation (17) and expressing it in non-dimensional variables we arrive at the
final boundary condition which is required to solve the problem:

C A (X = 0,T)
= eθ(T ) (18)
CB (X = 0,T)
The second case we will explore more general situations where the electron exchange reactions
involve finite kinetic electron transfer processes such that the reaction is slow and equilibrium is not
achieved in the time scale of the experiment. In that case the appropriate boundary condition is:

∂c A
( )
DA = [k r ed c A (0,t) − k o x cB (0,t)] (19)
∂ x x=0
Where k r ed and k o x are the reduction and oxidation rate constants. In the Butler-Volmer model we
have:

ed = k o e x = koe
(1−α)θ
k rBV −αθ
; k oBV (20)

where θ is the dimensionless parameter defined in equation (12) and k o is the standard rate
constant( whose )value corresponds
( ) to the value of the rate constants at the formal potential (k o
= k r ed E = E of = k o x E = E of ). Therefore, the final boundary condition when considering finite
kinetics in non-dimensional variables assumes the form:
∂C A
= e−αθ(T )Ko C A (0,T) − eθ(T )CB (0,T)
 
(21)
∂ X X =0

koϵ
Where the non-dimensional standard rate constant is defined as: Ko = DA .

SOLUTION METHODOLOGY
Case1: The problem is fully described by equations (13), (14), (15), (16), and (18). The first
step towards deriving a solution is to linearize the boundary conditions at the electrode. This can be
easily accomplished by approximating the derivative:

∂C j C j (X ma x ,T) − C j (0,T)
≈ f or j = A, B (22)
∂ X X =0

X ma x
This linearization allows us to exploit the bulk boundary conditions in the far field and express the
boundary conditions at the electrode as Dirichlet type boundary conditions.
If we employ equation (22) in equation (16) we arrive at:

C A (0,T) = 1 − d BCB (0,T) (23)

where we have used the fact that CB (X ma x ,T) = 0 and C A (X ma x ,T) = 1. If we then combine
equation (23) with equation (18), we attain:
1
CB (0,T) = ≡ Q(T) (24)
eθ(T ) + d B
Hence, to derive the current voltage response it suffices to solve the diffusion equation for species
B with the boundary condition at the surface of the electrode described by equation (23) i.e. the
problem is fully described by equations (13), (15) and (23).
087141-5 Samin, Lahti, and Zhang AIP Advances 5, 087141 (2015)

To solve this problem we employ the method of eigenfunction expansions. We define a func-
tion:
( )
X
W (X,T) = 1 − Q(T) (25)
X ma x
It can be clearly seen that: W (0,T) = Q(T) and W (X ma x ,T) = 0. Additionally, we let:
CB (X,T) = W (X,T) + V (X,T) (26)
Introducing equation (26) into the diffusion equation and the boundary and initial conditions yields:
∂V (X,T) ∂ 2V (X,T) ∂W (X,T)
= dB − (27)
∂T ∂X2 ∂T
Boundary Conditions:
V (0,T) = 0; V (X ma x ,T) = 0 (28)
Initial Conditions:
V (X, 0) = −W (X, 0) (29)
Using the method of separation of variables on the diffusion equation (27) without the source
term, one can easily derive the eigenfunctions and eigenvalues associated with the boundary condi-
tions for V (X,T) = H (X) Z(T):

Hn (X) Z n (T) = sin(λ n X)e−λ nT ; λ n = f or n = 1, 2, 3, . . . (30)
X ma x
Now we attempt to expand the source term in the eigenfunctions. Let:

∂W ∂Q (T) 
( )
X
S (X,T) = − = −1 = Sn (T) sin (λ n X) (31)
∂T X ma x ∂T n=0

Then using the orthogonality of the eigenfunctions we obtain:


Xma x (
∂Q (T)
)
2 X
Sn (T) = − 1− sin (λ n X) dX (32)
X ma x ∂T X ma x
0

Carrying out the integration by parts and the time derivative of the function Q we arrive at:
σe(θ i −σT )
( )
2
− i f T ≤ Tsw



nπ e(θ i −σT ) + d B2



Sn (T) = 

(33)
σe(θ v +σ(T −Ts w ))
 ( )
2
i f T > T



 nπ 2 sw
e(θ v +σ(T −Ts w )) + d B
 

Therefore, if we substitute the expansion of the source term (31) back into equation (27) and
we absorb the time-dependent functions Z n into the constants multiplying the eigenfunctions i.e.
V (X,T) = ∞

n=1 Vn (T) sin(λ n X) we attain:
∞  
 dVn
+ d B λ 2nVn − Sn (T) sin (λ n X) = 0 (34)
n=1
dT
Hence, this relationship yields a first order linear ordinary differential equation for Vn:
dVn
+ d B λ 2nVn − Sn (T) = 0 f or n = 1, 2, 3, . . . (35)
dT
This equation needs an initial condition Vn (0) for each n. From equation (29) and using the orthogo-
nality of the eigenfunctions we find:
 X ma x
2
an = V n (0) = − W (X, 0) sin (λ n X) dX (36)
X ma x 0
087141-6 Samin, Lahti, and Zhang AIP Advances 5, 087141 (2015)

Carrying out the integration by parts we arrive at:


( )
2 1
an = V n (0) = − θ
(37)
nπ e + dB
i

Equation (35) can be easily solved using an integrating factor and the solution is:
 T
e−d B λ n (T −τ) Sn (τ)dτ + e−d B λ nT Vn (0)
2 2
Vn (T) = (38)
0
Hence, the final solution may be written as:
( ) ∞  T 
X 
−d B λ 2n (T −τ) −d B λ 2n T
CB (X,T) = 1 − Q (T) + e Sn (τ)dτ + an e sin (λ n X) (39)
X ma x n=1 0

With an given in equation (37) and Sn given in equation (33).


The non-dimensional flux is given by:

∂CB d Bπ 
( )  T 
d BQ (T)
e−d B λ n (T −τ) Sn (τ)dτ + an e−d B λ nT (40)
2 2
JB = −d B = − n
∂ X X =0 X ma x X ma x n=1 0

The current is related to the non-dimensional flux through the relationship:


I = πϵ F D Ac∗A JB (41)
In equation (41) we used the fact that the area is given by: Area = πϵ 2.
Case 2: The solution methodology is the same. In this case we solve the problem described by
equations (13), (14), (15), (16), and (21).
We again start by linearizing the boundary condition of equation (16), resulting in equa-
tion (23). Thus, employing equation (23) in the linearized form of equation (21) and using the bulk
values in the far field yields the relationship:
e−αθ(T ) Ko X ma x
CB (0,T) = ≡ G(T) (42)
dB + dB o X ma x + X ma x K o e
e−αθ(T )K (1−α)θ(T )

Therefore, and following the procedure outlined in the previous section, we arrive at the
following relationship:
( ) ∞  T 
X 
−d B λ 2n (T −τ) −d B λ 2n T
CB (X,T) = 1 − G (T) + e Yn (τ)dτ + bn e sin(λ n X) (43)
X ma x n=1 0

where:
Xma x (
∂G (T)
)
2 X
Yn (T) = − 1− sin (λ n X) dX
X ma x ∂T X ma x
0

σe Ko X ma x (Ko X ma x e(1−α)(θ i −σT ) + αd B)


−α(θ i −σT )
( )
 2

 − i f T ≤ Tsw
nπ d B + d B e−α(θ i −σT )Ko X ma x + X ma x Ko e(1−α)(θ i −σT )2




= ( )

 (44)
2 σe−α(θ v +σ(T −Ts w ))Ko X ma x (Ko X ma x e(1−α)(θ v +σ(T −Ts w )) + αd B)
>

i f T T

sw

 2
 nπ
d B + d B e−α(θ v +σ(T −Ts w ))Ko X ma x + X ma x Ko e(1−α)(θ v +σ(T −Ts w ))


And:
2 e−αθ i Ko X ma x
bn = − (45)
nπ d B + d B e−αθ i Ko X ma x + X ma x Ko e(1−α)θ i
The non-dimensional flux is given by:

∂CB d Bπ 
( )  T 
dB
e−d B λ n (T −τ)Yn (τ)dτ + bn e−d B λ nT
2 2
JB = −d B = G (T) − n (46)
∂ X X =0 X ma x X ma x n=1 0
087141-7 Samin, Lahti, and Zhang AIP Advances 5, 087141 (2015)

RESULTS AND DISCUSSION


To verify our analytical expressions, the problem for case 1 described by equations (13), (14),
(15) and (18) was solved numerically using the Crank-Nicolson method. In addition, the lineari-
zation and eigenfunction expansion method was used as described by equation (40). The problem
was solved for vertex potential values of θ i = 20, and θ v = −20. These values are reasonable and
roughly correspond to Ei = 0.5 V and Ev = −0.5 V at 298 K. The values chosen for the other
parameters were:
2 (θ v − θ i )
σ = 200, T ma x = , dB = 1 ,
σ
For the numerical solution, ∆X = 0.01 and ∆T = 5 × 10−5 were used. The same time step was used
in implementing the analytical solution. In addition, the first 100 terms of the infinite sum of the
analytical solution were used. The result is shown in figure 1.
In executing the analytical solution, the global quadrature method was used in evaluating the
integrals as implemented in MATLAB’s “integral” function.
It is clear from figure 1 that the analytical solution agrees with the numerical solution. Further-
more, it is found that the relative difference between the two solutions does not exceed 3%. This
difference could be due to the truncation of the infinite series and the numerical approximations in
evaluating the integrals.
Next we verify the derived analytical expressions for case 2 where the kinetics are described
by the Butler-Volmer equation. We first solve the exact problem described by equations (13), (14),
(15), (16) and (21) numerically using the Crank-Nicolson method and then compare the results to
our analytical expression in equation (46).
The problem was solved for the same vertex potential values as case 1. The values chosen for
the other parameters were:
2 (θ v − θ i )
σ = 200, T ma x = , d B = 1 , Ko = 100, α = 0.5
σ
The same mesh size and time step from case 1 were used. Again, we only used the first 100 terms
in the infinite sum of the analytical solution and MATLAB’s built-in “integral” function was used in
evaluating the integrals.
In figure 2 we plot the non-dimensional flux-potential response curve for case 2. Moreover, in
figure 2 we plot the numerical non-dimensional flux-potential curve which we obtain from solv-
ing the problem described by equations (13), (14), (15), (16), and (18) using the Crank-Nicolson
method. Again, in this case, it can be seen that the analytical solution shows good agreement with

FIG. 1. Simulated cyclic voltammetry for Case 1 with the Nernst equation through use of a numerical scheme and through
the derived analytical solutions.
087141-8 Samin, Lahti, and Zhang AIP Advances 5, 087141 (2015)

FIG. 2. Simulated cyclic voltammetry for case 2 with the Butler-Volmer equation through use of a numerical scheme and
through the derived analytical solutions.

the numerical solution. In this case it was found that the relative difference between the two solu-
tions does not exceed 2%. Again, we believe the reason for the observed discrepancy is the small
errors introduced through the numerical approximations in implementing the analytical solution,
and also the truncation of the infinite series in the analytical solution.
It should be noted that our proposed solution method which is based on the linearization
assumption holds only for high reaction rates and for similar oxidant and reductant diffusivities.
For low reaction rates (Ko ≤ 80), the concentration at the electrode surface will not change rapidly
enough to yield a linear concentration profile through the solution as the concentration difference
(the driving force of diffusion) will not be high. The same idea is also true of dissimilar diffusivities:
the oxidant or reductant may not be able to diffuse quickly enough to warrant the assumption of a
linear concentration profile for one of the species. Therefore our solution methodology is limited
to situations of high electron transfer rates and species of equal diffusivities; otherwise the lineari-
zation assumption breaks down. Nevertheless, and despite its limitations, this methodology may be
useful for interpreting experimental data and may also be used for benchmarking purposes.

CONCLUSIONS
In this study we derive analytical solutions to the planar cyclic voltammetry model which
describes mass transport of species through the diffusion equations and the kinetics at the electrode
surface through the Nernst and Butler-Volmer equations. The solution methodology centers on
linearizing the boundary condition at the electrode surface and then using a series expansion. The
methodology only holds for high reaction rates and for similar oxidant and reductant diffusivities.
This method yields analytical expressions and does not utilize the Laplace transform. Furthermore,
the analytical solutions derived through this analysis yield good agreement with the numerical solu-
tion. This methodology is general and can be extended to cases that are more complicated as well
as other voltammetric techniques that rely on the diffusion process such as chronoamperometry.
Additionally, the analytical solutions may be useful for benchmarking purposes and could be used
for interpreting experimental data.

ACKNOWLEDGEMENTS
The authors would like to acknowledge the U.S. Department of Energy DOE-NEUP program,
Project number: 14-6542 for supporting this research.
1 Group Southampton Electrochemistry, Instrumental methods in electrochemistry (Ellis Horwood, Chichester, 2001).
2 D. Britz and B. Kastening, “On the electrochemical observation of a second-order decay of radicals generated by flash
photolysis or pulse radiolysis,” Journal of Electroanalytical Chemistry 56(1), 73-90 (1974).
087141-9 Samin, Lahti, and Zhang AIP Advances 5, 087141 (2015)

3 F. Lantelme and E. Cherrat, “Application of cyclic voltammetry to the study of diffusion and metalliding processes,” Journal
of Electroanalytical Chemistry 244(1-2), 61-68 (1988).
4 N. White and F. Lawson, “Potential sweep voltammetry of metal deposition and dissolution Part I. Theoretical analysis,”

Journal of Electroanalytical Chemistry 25(3), 409-419 (1970).


5 R. S. Nicholson and Irving Shain, “Theory of Stationary Electrode Polarography. Single Scan and Cyclic Methods Applied

to Reversible, Irreversible, and Kinetic Systems,” Anal. Chem. Analytical Chemistry 36(4), 706-723 (1964).
6 Allen J. Bard and Larry R. Faulkner, Electrochemical methods : fundamentals and applications (Wiley, New York, 2001).
7 L. Bortels, B. Van Den Bossche, and J. Deconinck, “Analytical solution for the steady-state diffusion and migration. Ap-

plication to the identification of Butler-Volmer electrode reaction parameters,” JOURNAL OF ELECTROANALYTICAL


CHEMISTRY 422(1/2), 161-167 (1997).
8 A. Molina, E. Torralba, C. Serna, and J. A. Ortuno, “Analytical solution for the facilitated ion transfer at the interface between

two immiscible electrolyte solutions via successive complexation reactions in any voltammetric technique: Application to
square wave voltammetry and cyclic voltammetry,” Electrochim Acta Electrochimica Acta 106, 244-257 (2013).
9 Talivaldis Berzins and Paul Delahay, “Oscillographic Polarographic Waves for the Reversible Deposition of Metals on Solid

Electrodes,” Journal of the American Chemical Society 75(3), 555-559 (1953).


10 Jan C. Myland and Keith B. Oldham, “Quasireversible cyclic voltammetry of a surface confined redox system: a mathemat-

ical treatment,” Electrochemistry Communications 7(3), 282-287 (2005).


11 Keith B. Oldham and Jan C. Myland, “Modelling cyclic voltammetry without digital simulation,” Electrochimica Acta

Electrochimica Acta 56(28), 10612-10625 (2011).


12 A. Eswari and L. Rajendran, “Mathematical modeling of cyclic voltammetry for EC reaction,” Russian Journal of Electro-

chemistry 47(2), 181-190 (2011).


13 A. Eswari and L. Rajendran, “Mathematical modeling of cyclic voltammetry for EC2 reaction,” Russian Journal of Elec-

trochemistry 47(2), 191-199 (2011).


14 R. G. Compton, Eduardo Laborda, and Kristopher R. Ward, Understanding voltammetry : simulation of electrode processes

(2014).

Das könnte Ihnen auch gefallen