Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Journal of Electroanalytical Chemistry 897 (2021) 115565

Contents lists available at ScienceDirect

Journal of Electroanalytical Chemistry


journal homepage: www.elsevier.com/locate/jelechem

Analytical solutions for the diffusive mass transfer at cylindrical and hollow-
cylindrical electrodes with reflective and transmissive boundary conditions
Tim Tichter a,⇑, Dirk Andrae b, Christina Roth c
a
Technical University of Denmark, Fysiksvej 307, 2800 Kgs Lyngby, Denmark
b
Freie Universität Berlin, Arnimallee 22, 14195 Berlin, Germany
c
Universität Bayreuth, Universitätsstr. 30, 95447 Bayreuth, Germany

A R T I C L E I N F O A B S T R A C T

Keywords: The theoretical treatment of cyclic voltammetry or chronoamperometry at cylindrical electrodes by means of
Cylindrical electrodes convolutive modeling requires an a priori knowledge of the time-dependent mass transfer functions and of the
Chronoamperometry diffusive flux of the electrochemically active species. In this paper, analytical solutions for both of these quan-
Cyclic voltammetry tities are derived for cylindrical and hollow-cylindrical electrodes with reflective and transmissive boundaries
Mass-transfer function
by means of Laplace transformation techniques. Furthermore, explicit equations for the concentration profiles
Convolution
Laplace transformation
of a Cottrellian potential step experiment are provided. All of these expressions are given in terms of infinite
series involving Bessel functions of the first and of the second kind. It is demonstrated that the summation of
only a few terms of these infinite series is usually sufficient to accurately compute the desired time-dependent
mass-transfer function or the current. This renders the numerical inversion of Laplace transformations, utilized
so far, obsolete and represents a mathematical supplement to the recent theory.

1. Introduction ods, since they offer the desired solution at the expense of computational
power.
Cylindrical diffusion plays an essential role in the context of elec- A more sophisticated and more generalized way towards the
trochemical applications – particularly when wire or fiber electrodes, desired solution is given in terms of Laplace transformation techniques
or electrodes with cylindrical pores are employed [1–8]. In a more and convolutive modeling [14,6]. In this context, the diffusion prob-
abstract sense, even porous electrodes like carbon felts can be regarded lem is transferred to – and solved in – the Laplace domain. The desired
as arrays of cylindrical microelectrodes with a statistically fluctuating time-dependent result (the flux of the electrochemically active species
diffusion domain size [9–11]. Unfortunately, the mathematical treat- in case of chronoamperometry and the mass-transfer function in case
ment of cylindrical diffusion poses a formidable challenge. This holds of voltammetry) is obtained subsequently by performing an inverse
true even after a significant simplification where the length of the elec- Laplace transformation on the respective Laplace domain solution.
trode filament/pore under investigation is considered much larger Unfortunately, only a very limited amount of related Laplace transform
than its radius1 such that edge effects are neglected and the radial dif- pairs are known for cylindrical diffusion and the majority of these
fusion component becomes the dominating quantity. This is caused by solutions is not yet described in the context of electrochemistry.
the fact that – unlike for spherical electrodes – the diffusion equation Instead, the time-dependent flux or mass-transfer functions are either
for a cylinder cannot be reduced to a simple form by substitution and a) evaluated as a set of discrete points in time by exploiting algorithms
the solutions usually involve Bessel functions [12]. To avoid these for numerical inversion of Laplace transformation (NILT) e.g. the Tal-
issues, the majority of electrochemical computations in cylindrical sym- bot [9,10,15–17] or the Gaver–Stehfest formulas [18–20] or b) by a
metry are therefore based on a direct and discrete numerical approxima- term-by-term inversion of truncated asymptotic expansions of the
tion of the diffusion equation by means of digital simulation [13]. Laplace domain solutions [8].
Despite the fact that this approach can usually handle the actual situa- Even though method a) provides an almost universal approach, it is
tion to high accuracy, these techniques might be termed brute force meth- not the analytical solution of the actual problem. Likewise, also b) offers

⇑ Corresponding author.
E-mail address: timtic@dtu.dk (T. Tichter).
1
This is a fair approximation in the majority of experimental scenarios.

https://doi.org/10.1016/j.jelechem.2021.115565
Received 4 February 2021; Received in revised form 21 July 2021; Accepted 25 July 2021
Available online 29 July 2021
1572-6657/© 2021 Published by Elsevier B.V.
T. Tichter et al. Journal of Electroanalytical Chemistry 897 (2021) 115565

an (exceptionally precise) approximation to the desired time-depen- the Laplace transformed variable related to the time t. The Laplace
dent result but is not the analytical solution either. transformed redox current – and thus the Laplace transformed function
With this paper we provide the analytical time-domain solutions f i ðsÞ – can be evaluated by taking the spatial derivative of Eq. (2) at the
for the concentration profiles, the flux and the mass-transfer functions electrode surface. According to Fick’s first law it follows
for the following four scenarios (cf. Fig. 1) in axial cylindrical symme-
@pi ðr; sÞ
try. In case (I) a cylindrical electrode with a reactive surface located at I ðsÞ ¼ nFADcB jr¼surface
@r
r ¼ a immersed in a semi-infinite diffusion domain is considered. Prac- pffiffiffiffi 1
tically, this scenario corresponds to a thin rod, a wire, or a fiber elec- ¼ nFA DcB ; ð3Þ
s  mi ðsÞ
trode that is dipped into an electrolyte. For this particular example the
flux (and thus the chronoamperometric current) is already known in where we implicitly defined a new function mi ðsÞ, which is herewith
the field of electrochemistry from the pioneering work of Aoki [1,2]. referred to as the Laplace transformed mass-transfer function. At this
It is included here, for the sake of completeness and generality. The point, it is of utmost importance to note that defining the function
related mass-transfer function in turn is only given in an abstract form mi ðsÞ by Eq. (3) is valid exclusively for limiting chronoamperometry
in the context of heat conduction [21] and is adapted for electrochem- where the surface concentration of the electrochemically active species
istry in this paper. In situation (II) a hollow-cylindrical electrode pro- consumed at the electrode is forced to zero3 at t > 0. In other scenarios
viding an internal finite diffusion domain with a reactive surface the current will change the surface concentrations and hence the defini-
located at r ¼ a is considered. This case basically corresponds to a tion of mi ðsÞ would become ambiguous. As a consequence, all of the fol-
cylindrical pore filled with electrolyte and was treated to a significant lowing derivations are dedicated to limiting chronoamperometry.
extent by Weidner [3–5]. Scenario (III) is dedicated to a cylindrical However, and most importantly, the functions mi ðsÞ can be nevertheless
electrode with a reactive surface located at r ¼ a and an impermeable adapted to any kind of potential perturbation.
boundary at r ¼ b (external finite reflective diffusion domain). This It should be noted that the related time-dependent mass-transfer
particular example was the focus of our recent publications [9,10] function mi ðt Þ possesses the dimension s1=2 . From Eqs. (1)–(3) it fol-
where it was utilized to describe the voltammetric current response lows that
of individual carbon fibers inside a carbon felt electrode. Due to the  pffiffiffiffi 
pffiffiffiffi @Pi ðr; sÞ D @ci ðr; sÞ
lack of an analytical solution it was treated in terms of NILT techniques f i ðsÞ ¼  D  ¼ : ð4Þ
at that time. Finally, situation (IV) describes a cylindrical electrode with @r r¼surface cB @r r¼surface
a reactive surface located at r ¼ a and an ideally permeable wall, corre- From Eq. (3) in turn it immediately follows that
sponding to a transmissive boundary, at r ¼ b. This case can be inter- 
preted as a cylindrical electrode subjected to a homogeneous flow of 1
f i ðt Þ ¼ L1 ðt Þ ð5Þ
electrolyte towards its surface such that a diffusion layer of constant s  mi ðsÞ
thickness is formed2. Taking into account all the four aforementioned sce- and of course
narios, the present paper provides an overview of – and a supplement to – ( )
the analytical solutions for electroanalytical experiments involving cylin- 1
mi ðt Þ ¼ L1 ðt Þ: ð6Þ
drical electrodes and covers almost any experimentally relevant situation. s  f i ðsÞ
Moreover, it intends to satisfy the electrochemist’s mathematical curiosity
by providing detailed derivations of the resulting analytical expressions. Likewise, the time-dependent concentration profile can be accessed
via
  
2. Theory and derivation pi ðr; sÞ 

ci ðr; t Þ ¼ cB 1  L1 ðt Þ ¼ cB 1  L1 Pi ðr; sÞ ðt Þ : ð7Þ
s
The time-dependent redox current of any voltammetric or
In Eqs. (5)–(7), L1 denotes the inverse Laplace transformation
chronoamperometric experiment is given by
operator which is defined for an arbitrary function g by
pffiffiffiffi
I ðt Þ ¼ nFAcB Df i ðt Þ; ð1Þ Z γþi1
1
g ðt Þ ¼ L1 fg ðsÞg ¼ g ðsÞest ds: ð8Þ
where n is the number of electrons transferred, A is the electrode sur- 2πi γi1
face, F is Faraday’s constant, cB and D are the bulk concentration and
In Eq. (8) i is the imaginary unit (i2 ¼ 1) and γ is a constant to be
the diffusion coefficient of the electrochemically active species and
chosen such that all singularities of the integrand – and thus g – are
f i ðt Þ is the specific flux (or current) function with the unit of s1=2 , located to the left of γ in the complex plane.
respectively. Since only f i ðt Þ (specified by the individual experimental An alternative, yet very important way of accessing f i ðt Þ is given by
conditions) defines the shape of the time-dependent current wave, rewriting Eq. (5) in terms of convolution integrals4 which yields
Eq. (1) is somewhat universal. Considering diffusion as the only quan- Z t
tity contributing to the mass transfer in an electrochemical experiment, 1¼ f i ðτÞmi ðt  τÞ dτ: ð9Þ
this particular quantity can be accessed by means of Laplace integral 0
transformation in the following way. Generally, solving the diffusion Since the function mi ðt Þ has a removable singularity at the upper
equation via Laplace transformation techniques yields the Laplace integration limit due to limt!0 mi ðt Þ ! 1 it is more convenient to
transformed concentration profile of the electrochemically active spe- rewrite Eq. (9) by performing an integration by parts. This gives
cies being consumed at the electrode surface as
Z f i ðt Þ
 
pi ðr; sÞ   1 ¼ M i ðt  τÞf i ðτÞjτ¼t
τ¼0 þ M i ðt  τÞ df i ðτÞ; ð10Þ
ci ðr; sÞ ¼ cB ½1  pi ðr; sÞ ¼ cB s1  ¼ cB s1  Pi ðr; sÞ ; ð2Þ f i ð0Þ
s
where M i ðt Þ is the antiderivative of mi ðt Þ. Now, Eqs. (5)–(7) and (10)
where s1  Pi ðr; sÞ is the Laplace transformed dimensionless concentra- include the four functions f i ðt Þ; mi ðt Þ; M i ðt Þ and ci ðr; t Þ, which will be
tion profile, i.e. the quantity depending on electrode geometry. In Eq.
(2), r is the radial coordinate normal to the electrode surface and s is 3
In case of chronoamperometry not operated at the limit, one has to replace cB by the
constant value of cB  cðx ¼ 0; t > 0Þ.
2 4
The thickness of this diffusion layer might be regarded in analogy to a rotating disc This is the usual procedure for simulating electroanalytical experiments with transient
electrode, i.e. it is governed by the flow rate. boundary conditions, e.g. cyclic voltammetry.

2
T. Tichter et al. Journal of Electroanalytical Chemistry 897 (2021) 115565

Fig. 1. Modes of cylindrical diffusion treated in this paper. (I) external semi-infinite, (II) internal finite reflective, (III) external finite reflective, (IV) external finite
transmissive. Boundary conditions are given for the outer point of the diffusion domain opposed to the electrode surface..

determined for the four types of cylindrical electrodes depicted in Fig. 1 Performing the spatial derivative of Eq. (4) in Eq. (15) and evaluat-
at next. ing it at the electrode surface located at r ¼ a defines f I ðsÞ as
 pffiffiffi
2.1. General solution of the diffusion equation in axial-cylindrical K1 a Ds
f I ðsÞ ¼ pffiffi  pffiffiffi : ð16Þ
symmetry
s K0 a Ds

The diffusion equation in axial-cylindrical symmetry can be stated Likewise, the function mI ðsÞ is defined by
as  pffiffiffi
 2  K0 a Ds
@cðr; t Þ @ cðr; t Þ 1 @cðr; t Þ mI ðsÞ ¼ pffiffi  pffiffiffi : ð17Þ
¼D þ : ð11Þ
@t @r 2 r @r s K1 a Ds

Performing a Laplace transformation on Eq. (11) gives Finally, using the Laplace transformation property for an integra-
 2  tion in the time domain gives M I ðsÞ as
@ cðr; t Þ 1 @cðr; t Þ
cðr; t ¼ 0Þ þ scðr; sÞ ¼ D þ : ð12Þ  pffiffiffi
@r 2 r @r
K0 a Ds
Assuming that the concentration of the electrochemically active M I ðsÞ ¼  pffiffiffi : ð18Þ
s3=2 K1 a Ds
species equals the bulk concentration cB at t ¼ 0 at each and every r,
it follows
  2.2.1. Situation I: The inverse Laplace transformation
D 2 @ 2 cðr; t Þ @cðr; t Þ
0¼ r þr  r 2 cðr; sÞ þ r 2 cB : ð13Þ The normalized concentration profile related to Eq. (15) is derived
s @r 2 @r
by applying the complex inversion formula (Eq. 8). It follows
The general solution of Eq. (13) can be written in terms of zeroth-  
order modified Bessel functions of the first (I0 ) and second (K0 ) kind Z γþi1 K0 r pffiffisffi
cI ðr; t Þ 1 D
¼1   est ds: ð19Þ
as cB 2πi γi1 s K0 apffiffisffi
rffiffiffiffi rffiffiffiffi D
s s
cðr; sÞ ¼ Cþ I0 r þ C K0 r þ cB : ð14Þ Since the integrand of Eq. (19) has a branch point at s ¼ 0 and no
D D
singularities in the left half-plane [21] we use the contour of Fig. 2(C).5
The coefficients Cþ and C need to be adapted by the boundary The respective solution to Eq. (19) was provided in the context of heat
conditions of the experiment and will be discussed in detail for situa- conduction from a cylindrical pipe by Carslaw and Jaeger [21]. It is
tions I-IV now. slightly rearranged to fit the concentration profile of the educt species as
Z
cI ðr; t Þ 2 1 Dtu2 J0 ður ÞY0 ðuaÞ  Y0 ður ÞJ0 ðuaÞ
2.2. Situation I: The semi-infinite cylindrical diffusion space ¼ e   du: ð20Þ
cB π 0 u J20 ðuaÞ þ Y20 ðuaÞ
For a semi-infinite external cylindrical diffusion domain, Eq. (14) In Eq. (20), J0 and Y0 are zeroth order Bessel functions of the first
can be simplified extensively. It immediately follows that Cþ ¼ 0 as and of the second kind, respectively. Considering Eq. (20), the related
I0 ð1Þ ¼ 1, but cI ð1; sÞ ¼ cB . Since a Cottrellian potential step exper- f I ðt Þ can be found readily by performing the spatial derivative and
iment forces the surface concentration (at r ¼ a) of the educt species to evaluating it at r ¼ a and applying the relation of Eq. (4). In this man-
zero, it follows that cða; sÞ ¼ 0 as well. This defines the constant C ner we obtain
and thus the function cðr; sÞ. Normalizing by cB gives
 pffiffiffi
cI ðr; sÞ K0 r Ds
¼ s1   pffiffiffi : ð15Þ
cB s K0 a Ds
5
Fig. 2 (A) and (B) depict the complex map of the function mI ðsÞ, which is most
illustrative for this scenario.

3
T. Tichter et al. Journal of Electroanalytical Chemistry 897 (2021) 115565

Fig. 2. (A) Real part of the function mI ðsÞ. (B) Imaginary part of the function mI ðsÞ. (C) contour for the inverse Laplace transformation with a branch point at s ¼ 0.

pffiffiffiffi Z 2 2
 pffiffiffi
4 D 1 eDta v I0 a Ds
f I ðt Þ ¼  2  dv; ð21Þ
aπ 2
0 v J0 ðvÞ þ Y20 ðvÞ mII ðsÞ ¼ pffiffi  pffiffiffi : ð26Þ
s I1 a Ds
where the relation (A.9), given in the appendix, was used and where we
substituted v ¼ ua. In the context of electrochemistry, Eq. (21) is known  pffiffiffi
I0 a Ds
from the early work of Aoki [1,2], as well as from a more recent pub- M II ðsÞ ¼  pffiffiffi : ð27Þ
lication [22] and is therefore not discussed further here. Nevertheless, s3=2 I1 a Ds
it was listed here for the sake of completeness. In contrast to Eq. (21),
the inversion of mI ðsÞ is not given in any of these references. However,
it can be derived in an analogue way to Eq. (20) (cf. Carslaw and Jaeger 2.3.1. Situation II: The inverse Laplace transformation
[21] p. 335). A much faster way of accessing mI ðt Þ would be, however, Calculating the inverse Laplace transformation of Eqs. (24)–(27)
a differentiation of M I ðt Þ with respect to t, once M I ðt Þ is known. Fortu- can be performed by exploiting the residue theorem (cf. appendix
nately, an abstract version of M I ðt Þ is already provided by Carslaw and (A.10)). In case of Eq. (24), the inverse Laplace transformation is
Jaeger in the context of heat conduction ([21], p. 338). Evaluating this defined by
 
particular formula at the surface of the cylinder (electrode surface), Z γþi1 I0 r pffiffisffi
substituting the specific constants a and D, again exploiting the relation cII ðr; t Þ 1 D
¼1   est ds: ð28Þ
(A.9) given in the appendix, and substituting v ¼ au we get cB 2πi γi1 s I0 apffiffisffi
D
Z 1 2 2
4a 1  eDta v
M I ðt Þ ¼ pffiffiffiffi  2  dv; ð22Þ The respective solution is given by Carslaw and Jaeger in the con-
Dπ 2 0 v J1 ðvÞ þ Y21 ðvÞ
3
text of heat conduction [21, p. 328]. By examining the integrand of
where J1 and Y1 are first-order Bessel functions of the first and of the [28], it is seen that all singularities (simple poles) are located along
second kind. Since the integration in Eq. (22) is performed with respect the negative real axis (cf. Fig. 3).6 More precisely at s ¼ Dα2n , where
to the dummy variable v, but accessing mI ðt Þ from M I ðt Þ requires for a αn represents the n-th positive zero of J0 ðaαÞ. The Bessel function of
differentiation with respect to t, the operations of differentiation and the first kind is introduced by exploiting identity (A.6), given in the
pffiffi pffiffiffiffi
integration can by interchanged. This defines mI ðt Þ as appendix, since s ¼ i Dαn . Calculating the n-th residue is performed
pffiffiffiffi Z 2 2 by first using
4 D 1 eDta v  pffiffi
mI ðt Þ ¼   dv: ð23Þ
aπ 2 0 v J21 ðvÞ þ Y21 ðvÞ @I0 paffiffiDffi s aαn
s js¼Dα2n ¼  J ðaαn Þ ð29Þ
@s 2 1
2.3. Situation II: The internal finite reflective cylindrical diffusion space on the denominator of (28). There, identity (A.6) was used. A subse-
quent combination with the numerator yields
In case of an internal finite reflective diffusion domain (situation
2 J0 ðrαn ÞeDαn t
2

II), the Laplace transformed concentration profile (cII ðr; sÞ), the Laplace Resn ¼  : ð30Þ
a αn J1 ðaαn Þ
transformed flux (f II ðsÞ) and the Laplace transformed mass-transfer
functions (mII ðsÞ and M II ðsÞ) are defined in a similar way than in situa- The residue at zero is 1, since I0 ð0Þ ¼ 1 in numerator and denom-
tion I. Since K0 ð0Þ ! 1 but 0 ⩽ cII ð0; sÞ ⩽ cB it follows that C ¼ 0 in inator of Eq. (28). Since J0 has infinitely many zeros, the sum of all
Eq. (14). In analogy to the Cottrellian potential step in Eq. (15) the residues is an infinite series. Therefore, it follows that
normalized concentration profile is found as
cII ðr; t Þ 2 1 J0 ðrαn ÞeDαn t
2

 pffiffiffi ¼ ∑ : ð31Þ
cB a n¼1 αn J1 ðaαn Þ
cII ðr; sÞ I0 r Ds
¼ s1   pffiffiffi ; ð24Þ In analogy to the semi-infinite diffusion, the function f II ðt Þ can be
cB s I0 a Ds generated readily from Eq. (31), by using Eq. (4). Thus, a differentia-
tion with respect to r, evaluating the derivative at r ¼ a and multiply-
where the region inside of the cylinder is bound by 0 ⩽ r ⩽ a. Since in pffiffiffiffi
ing with  D (this time the negative sign is introduced since r is
this case, r ⩽ a, the Laplace domain solutions for f II ðsÞ; mII ðsÞ and M II ðsÞ
bounded by a) gives
are defined in analogy to the previous example (though with an oppo-
site sign) by
 pffiffiffi
I1 a Ds
f II ðsÞ ¼ pffiffi  pffiffiffi : ð25Þ 6
This applies to all the following scenarios with finite diffusion. Fig. 3(A) and (B) depict
s I0 a Ds
the real and imaginary part of the function mII ðt Þ.

4
T. Tichter et al. Journal of Electroanalytical Chemistry 897 (2021) 115565

Fig. 3. (A) Real part of the function mII ðsÞ. (B) Imaginary part of the function mII ðsÞ. (C) Contour for the inverse Laplace transformation of a function with simple
poles located along the negative real axis (at s ¼ Dα2n ).

pffiffiffiffi
2 D 1 Dα2n t From Eq. (35), again the function f III ðsÞ can be accessed by perform-
f II ðt Þ ¼ ∑e ; ð32Þ
a n¼1 ing the spatial derivative with respect to r and evaluating it at r ¼ a
and by using Eq. (4). This gives
again with αn being the n-th positive real zero of J0 ðaαn Þ. The normal-
 pffiffiffi  pffiffiffi  pffiffiffi  pffiffiffi
ized flux Eq. (32) is neither given in the context of heat conduction by I1 b Ds K1 a Ds  K1 b Ds I1 a Ds
Carslaw and Jaeger, nor for a Cottrellian potential step in the context of f III ðr; sÞ ¼ pffiffi h  pffiffiffi  pffiffiffi  pffiffiffi  pffiffiffii : ð36Þ
electrochemistry. s I0 a Ds K1 b Ds þ K0 a Ds I1 b Ds
In analogy to Eq. (32), the inverse Laplace transformation to mII ðsÞ
of Eq. (27) can be found. Since this quantity is given by others [3–5,8], The Laplace domain solutions for mIII ðsÞ and M III ðsÞ are given in our
the derivation is omitted here and only the result is listed for the sake previous publications as
of completeness as  pffiffiffi  pffiffiffi  pffiffiffi  pffiffiffi
I0 a Ds K1 b Ds þ K0 a Ds I1 b Ds
pffiffiffiffi
2 D 1 mIII ðr; sÞ ¼ pffiffi h  pffiffiffi  pffiffiffi  pffiffiffi  pffiffiffii : ð37Þ
1 þ ∑ eDαn t ;
2
mII ðt Þ ¼ ð33Þ s I1 b Ds K1 a Ds  K1 b Ds I1 a Ds
a n¼1

where this time αn > 0 is the n-th positive zero of the Bessel function of and
the first kind and of order one (J1 ðaαn Þ ¼ 0).  pffiffiffi  pffiffiffi  pffiffiffi  pffiffiffi
I0 a Ds K1 b Ds þ K0 a Ds I1 b Ds
Since M II ðt Þ is the antiderivative of mII ðt Þ with respect to t, a term M III ðr; sÞ ¼ h  pffiffiffi  pffiffiffi  pffiffiffi  pffiffiffii : ð38Þ
by term integration of Eq. (33) gives s3=2 I1 b Ds K1 a Ds  K1 b Ds I1 a Ds
pffiffiffiffi
2 Dt 2 1 1 
þ pffiffiffiffi ∑ 2 1  eDαn t ;
2
M II ðt Þ ¼ ð34Þ
a a D n¼1 αn 2.4.1. Situation III: The inverse Laplace transformation
where again αn > 0 is the n-th positive root of J1 ðaαn Þ. The concentration profile of the electrochemically active species
formed at the electrode surface in situation III during a Cottrellian
2.4. Situation III: The external finite reflective cylindrical diffusion space potential step experiment can be regarded as analogue to the general
solution for heat conduction in an internally heated hollow cylinder
The derivation of the Laplace domain solution for the mass- provided by Carslaw and Jaeger [21, pp. 332–335]. In analogy, the
transfer function of a cylindrical electrode in an external cylindrical concentration profile of the species being consumed at the electrode
diffusion space was the focus of one of our recent works [9] where surface is found as
we approximated the diffusion domain inside a carbon felt elec-
cIII ðr; t Þ 1 2 J ðbαn Þ½J0 ðrαn ÞY0 ðaαn Þ  Y0 ðrαn ÞJ0 ðaαn Þ
2
trode as an array of cylindrical microelectrodes. The respective ¼ π ∑ eDtαn 1 :
cB n¼1 J20 ðaαn Þ  J21 ðbαn Þ
time-domain solution was, however, generated by utilizing a mod-
ified Talbot contour for the numerical inversion of Laplace trans- ð39Þ
forms which was somewhat unsatisfying but necessary, since no The αn in Eq. (39) are defined as the n-th positive zero of the equa-
analytical solutions were accessible at that time. These will be, tion 0 ¼ J0 ðaαn ÞY1 ðbαn Þ  Y0 ðaαn ÞJ1 ðbαn Þ. The function f III ðt Þ is gen-
however, provided here. For the sake of completeness, the respec- erated again by differentiating Eq. (39) with respect to r, evaluating
tive concentration profiles and the normalized Cottrellian current pffiffiffiffi
the result at r ¼ a and multiplying with D according to Eq. (4). This
will be provided as well. gives
The Laplace domain solution of the concentration profile of the pffiffiffiffi
electrochemically active species formed at the electrode surface during 2 D 1 Dtα2n J21 ðbαn Þ
f III ðt Þ ¼ ∑e ; ð40Þ
a Cottrellian potential step experiment can be obtained from Eq. (14). a n¼1 J0 ðaαn Þ  J21 ðbαn Þ
2

Unlike in situations I) and II), none of the constants Cþ and C will be


where we used (A.9). As in the case of f III ðt Þ, analytical solutions for
zero now, since 0 < a ⩽ r ⩽ b < 1. However, similar to case I)
mIII ðt Þ and M III ðt Þ were not reported in the literature. However, they
cða; t Þ ¼ 0 as t > 0. Furthermore, the no-flux boundary at r ¼ b needs
can be obtained similar to situation II) by applying the complex inver-
to be considered, meaning that at @c@rðr;sÞ
jr¼b ¼ 0. Combining these
sion formula. For mIII ðt Þ we get
boundary conditions and determining the constants Cþ and C gives Z γþi1
the concentration profile as 1
mIII ðt Þ ¼ mIII ðsÞest ds: ð41Þ
 pffiffiffi  pffiffiffi  pffiffiffi  pffiffiffi 2πi γi1
cIII ðr; sÞ K1 b Ds I0 r Ds þ I1 b Ds K0 r Ds
1
¼ s  h  pffiffiffi  pffiffiffi  pffiffiffi  pffiffiffii : ð35Þ By carefully examining the integrand in (41), we notice that mIII ðsÞ
cB s I0 a s K1 b s þ K0 a s I1 b s
D D D D has an infinite number of simple poles located along the negative real
axis (cf. Fig. 3). Therefore, we can obtain the inverse Laplace

5
T. Tichter et al. Journal of Electroanalytical Chemistry 897 (2021) 115565

transformation as the sum of the residues at all of these poles. Setting chemically active species – similar to the concept of a rotating
pffiffi pffiffiffiffi
s ¼ Dα2n such that s ¼ i Dαn , we find the location of the poles7 as disc electrode. Therefore the respective concentration profile will
the n-th positive roots of the equation eventually approach a steady state. In the context of heat conduc-
0 ¼ Y1 ðaαn ÞJ1 ðbαn Þ  Y1 ðbαn ÞJ1 ðaαn Þ, where we used the Bessel func- tion, a similar problem was solved by Carslaw and Jaeger for a
tion identities (A.6) and (A.8), given in the appendix. Calculating the hollow cylinder which is heated at the inner boundary (r ¼ a)
residues (except the one at s ¼ 0) is performed as follows. In the denom- and kept at a constant temperature at the outer (r ¼ b) boundary
inator we put [21, pp. 332–335]. This can be translated to the context of electro-
rffiffiffiffi rffiffiffiffi rffiffiffiffi rffiffiffiffi chemistry in a straightforward way if the respective temperature
pffiffi @ s s s s
 s K1 b I1 a I1 b K1 a jpsffi ¼ipffiffiDffiαn profile is associated with the concentration profile of the species
@s D D D D
formed at the electrode surface. The respective Laplace domain

¼ pffiffiffiffifb½J0 ðbαn ÞY1 ðaαn ÞþJ1 ðaαn ÞY2 ðbαn Þa½J0 ðaαn ÞY1 ðbαn ÞþJ1 ðbαn ÞY2 ðaαn Þg: solution for the species consumed at the electrode surface can be
4 D
ð42Þ derived from this result in a straightforward manner and can be
formulated as
In Eq. (42), Y2 is the Bessel function of the second kind and of
pffiffi pffiffiffiffi  pffiffiffi  pffiffiffi  pffiffiffi  pffiffiffi
order two. Evaluating the numerator of (41) at s ¼ i Dαn we get K0 b Ds I0 r Ds  I0 b Ds K0 r Ds
cIV ðr; sÞ
rffiffiffiffi rffiffiffiffi rffiffiffiffi rffiffiffiffi ¼ s1  h  pffiffiffi  pffiffiffi  pffiffiffi  pffiffiffii : ð50Þ
s s s s cB s I0 a s K0 b s  K0 a s I0 b s
K0 a I1 b þ I0 a K1 b est jpsffi ¼ipffiffiDffiαn D D D D
D D D D
iπeDtαn
2
The function f IV ðsÞ is again defined by differentiating Eq. (50) with
¼ fJ0 ðaαn ÞY1 ðbαn Þ  J1 ðbαn ÞY0 ðaαn Þg; ð43Þ pffiffiffiffi
2 respect to r, evaluating the result at r ¼ a and multiplying by D. This
gives
where we used the Bessel function identities (A.5) and (A.7). Combin-  pffiffiffi  pffiffiffi  pffiffiffi  pffiffiffi
ing (42) and (43) we get the n-th residue. The desired inverse Laplace K0 b Ds I1 a Ds þ I0 b Ds K1 a Ds
transformation is obtained subsequently as the sum of all these residues f IV ðsÞ ¼  pffiffi h  pffiffiffi  pffiffiffi  pffiffiffi  pffiffiffii : ð51Þ
and hence s I0 a Ds K0 b Ds  K0 a Ds I0 b Ds
pffiffiffiffi 1
mIII ðt Þ ¼ ResIII;0 þ 2 D ∑ ResIII;n ð44Þ From Eq. (6) the definition of mIV ðt Þ follows as
n¼1
with

eDtαn fJ0 ðaαn ÞY1 ðbαn Þ  J1 ðbαn ÞY0 ðaαn Þg


2

ResIII;n ¼ Bða; b; αn Þ eDtαn ¼


2
; ð45Þ
b½J0 ðbαn ÞY1 ðaαn Þ þ J1 ðaαn ÞY2 ðbαn Þ  a½J0 ðaαn ÞY1 ðbαn Þ þ J1 ðbαn ÞY2 ðaαn Þ

 pffiffiffi  pffiffiffi  pffiffiffi  pffiffiffi


I0 a Ds K0 b Ds  K0 a Ds I0 b Ds
defining the function Bða; b; αn Þ as the term containing the Bessel func- mIV ðsÞ ¼  pffiffi h  pffiffiffi  pffiffiffi  pffiffiffi  pffiffiffii ; ð52Þ
tions. The residue at s ¼ 0 is found by evaluating the following limit s K0 b Ds I1 a Ds þ I0 b Ds K1 a Ds
8  pffiffiffi  pffiffiffi  pffiffiffi  pffiffiffi 9
< I0 a Ds K1 b Ds þ K0 a Ds I1 b Ds = which also defines M IV ðsÞ as
lim  pffiffi h  pffiffiffi  pffiffiffi  pffiffiffi  pffiffiffii ð46Þ  pffiffiffi  pffiffiffi  pffiffiffi  pffiffiffi
s!0 : @ ;
@s
 s K1 b Ds I1 a Ds  I1 b Ds K1 a Ds I0 a Ds K0 b Ds  K0 a Ds I0 b Ds
M IV ðsÞ ¼  h  pffiffiffi  pffiffiffi  pffiffiffi  pffiffiffii : ð53Þ
Doing so, we find s3=2 K0 b Ds I1 a Ds þ I0 b Ds K1 a Ds
pffiffiffiffi
2a D
ResIII;0 ¼ 2 ; ð47Þ
b  a2 2.5.1. Situation IV: The inverse Laplace transformation
The inverse Laplace transformation of Eq. (50) is given by the gen-
which defines
eral result provided by Carslaw and Jaeger [21, pp. 332–335]. It can
pffiffiffiffi a 1
be formulated as
þ ∑ Bða; b; αn Þ eDtαn :
2
mIII ðt Þ ¼ 2 D 2 ð48Þ
b  a2 n¼1 cIV ðr; t Þ lnðbÞ  lnðr Þ
¼1
From Eq. (48), the respective M III ðt Þ is obtained as cB lnðbÞ  lnðaÞ
pffiffiffiffi J20 ðbαn Þ½J0 ðrαn ÞY0 ðaαn Þ  Y0 ðrαn ÞJ0 ðaαn Þ
2 1 Bða; b; αn Þ h i 1
 π ∑ eDtαn
2
2 Dat ;
1  eDtαn ;
2
M III ðt Þ ¼ 2 þ pffiffiffiffi ∑ ð49Þ J20 ðaαn Þ  J20 ðbαn Þ
b  a2 D n¼1 α2n n¼1

ð54Þ
which completes the derivation for scenario (III).
where αn are the n-th positive root of
2.5. Situation IV: The external finite transmissive cylindrical diffusion space Y0 ðaαn ÞJ0 ðbαn Þ  J0 ðaαn ÞY0 ðbαn Þ ¼ 0. Consequently, the function
f IV ðt Þ can be accessed by differentiating Eq. 54 with respect to r, eval-
pffiffiffiffi
An external finite transmissive cylindrical diffusion space might uating the derivative at r ¼ a and multiplying the result by D. This
be assumed whenever the diffusive mass transfer towards a cylin- gives
drical electrode is supported by a convective flux of the electro- pffiffiffiffi " #
D 1 1
Dtα2n J20 ðbαn Þ
f IV ðt Þ ¼ þ 2 ∑e ; ð55Þ
7
a lnðbÞ  lnðaÞ n¼1 J20 ðaαn Þ  J20 ðbαn Þ
The term in the square brackets in the denominator of Eq. (37) (the one containing the
Bessel functions) has to be zero.

6
T. Tichter et al. Journal of Electroanalytical Chemistry 897 (2021) 115565

where we have used identity (A.9). The inverse Laplace transformation electrode radii. Likewise, the radial diffusion contribution of a cylin-
of Eq. (52) is found by applying the residue theorem. The poles of drical electrode can be spotted best in panel (A), for f I ðt Þ at an elec-
mIV ðsÞ are found at the roots of the denominator. These are located trode with a small diameter (top curve). Indeed, it can be seen that
along the negative real axis and are all simple. Setting s ¼ Dα2n such the diffusive mass transfer increases as the electrode radius becomes
pffiffi pffiffiffiffi
that s ¼ i Dαn , the αn are defined by smaller. In analogy to panel (A), panel (B) depicts the function mI ðt Þ
0 ¼ J0 ðbαn ÞY1 ðaαn Þ  Y0 ðbαn ÞJ1 ðaαn Þ. Calculating the residues is per- for different values of a. It can be seen that mI ðt Þ decreases sooner as
formed by setting a gets smaller. Likewise M I ðt Þ will be smaller than (or at most equal
rffiffiffiffi rffiffiffiffi rffiffiffiffi rffiffiffiffi to) the planar semi-infinite M ðt Þ, since it is the antiderivative of mI ðt Þ.
pffiffi @ s s s s
s K0 b I1 a þ I0 b K1 a jpsffi¼ipffiffiDffiαn Panels (D)–(F) of Fig. 4 correspond to an internal cylindrical finite
@s D D D D reflective diffusion domain (e.g. to a cylindrical pore). In panel (D) it
π
¼ pffiffiffiffi fJ0 ðbαn Þ½aY0 ðaαn Þ  Y1 ðaαn Þ þ J1 ðaαn Þ½bY1 ðbαn ÞY0 ðbαn Þ can be seen that if a decreases, the flux function f II ðt Þ will deplete
4 D sooner in time. In contrast, panel (E) shows that in case of a finite dif-
aJ0 ðaαn ÞY0 ðbαn Þ  bY1 ðaαn ÞJ1 ðbαn Þg fusion domain the function mðt Þ (here mII ðt Þ) approaches a limit as
π
¼ pffiffiffiffi N IV ðαn ; a; bÞ ð56Þ time increases. This limit depends on the diffusion domain size and
4 D is achieved faster, the smaller the diffusion domain gets. From Eq.
in the denominator and (33), this particular limit can be quantified as
 rffiffiffiffi rffiffiffiffi rffiffiffiffi rffiffiffiffi mII ðt ! 1Þ ¼ 2D1=2 a1 . Since the function M II ðt Þ is the antiderivative
s s s s
est K0 b I0 a  I0 b K0 a jpffis¼iDαn of mII ðt Þ, this particular limit will enforce a constant slope in M II ðt Þ,
D D D D which is indeed seen in Fig. 4(F).
π
¼ ½Y0 ðaαn ÞJ0 ðbαn Þ  J0 ðaαn ÞY0 ðbαn ÞeDtαn In general, panels (A), (B), (D) and (E) of Fig. 4 illustrate that f i ðt Þ
2

2
π and mi ðt Þ will behave somewhat complementary. This implies that if
¼ Z IV ðαn ; a; bÞeDtαn
2
ð57Þ f i ðt Þ depletes faster, mi ðt Þ might approach a limit and vice versa. Fur-
2
thermore, it is worth to mention that all the series solutions derived in
in the numerator. Thus, combining Eqs. (56) and (57) gives the n-th the theory section converge quite rapidly for large values of t such that
residue. Since there are infinitely many poles located at the negative a summation of only few terms (5 to 50) is already sufficient to approx-
real axis, the desired inverse Laplace transformation mIV ðt Þ is found imate the desired time-dependent result. However, particularly the
as the following infinite series series solutions for the functions f i ðt Þ and mi ðt Þ will be divergent if t
pffiffiffiffi 1 Z IV ðαn ; a; bÞ approaches zero, since ðπt Þ1=2 – the function describing the flux and
mIV ðt Þ ¼ 2 D ∑ eDtαn : ð58Þ
n¼1 N IV ðαn ; a; bÞ the mass transfer at a planar electrode – is singular at t ¼ 0. Conse-
quently, more terms are required in the summation for small values
In analogy to situation (III), the M IV ðt Þ can be derived from Eq. (58)
of t. Consequently, all 500 roots were used during the summation in
by a term by term integration. This finally gives
all the calculations.
2 1   Z IV ðαn ; a; bÞ In contrast to Fig. 4, Fig. 5 depicts the functions f i ðt Þ; mi ðt Þ and
M IV ðt Þ ¼  pffiffiffiffi ∑ 1  eDtαn 2 ð59Þ
D n¼1 αn N IV ðαn ; a; bÞ M i ðt Þ for cylindrical electrodes with an external finite diffusion space
(scenario III and IV of Fig. 1). Among them, scenario III – the external
and concludes the theory section of this paper.
cylindrical finite reflective diffusion space – is the most complicated
case. This is caused by the fact that the diffusive flux towards the elec-
3. Results and discussion trode can occur in one, two or three different modes. In the case that b
is only slightly larger than a (a very narrow gap between electrode and
Utilizing the analytical expressions derived in the theory section, wall in Fig. 1(III), the diffusive flux can be described accurately in
we obtain the functions f i ðt Þ; mi ðt Þ and M i ðt Þ at any desired point in terms of planar diffusion only. Thereby, the diffusion will behave pla-
time. Since the concentration profile itself is usually less important nar semi-infinite at small time instances since the concentration profile
in electrochemical applications, the respective solutions are not fur- has not reached the outer boundary. However, as soon as the zone of
ther considered here. The major challenge in computing f i ðt Þ; mi ðt Þ dilution hits the impermeable wall located at r ¼ b, the diffusive flux
and M i ðt Þ lies in the determination of the roots of the related Bessel will deplete significantly (cf. Fig. 5(A), curve on the left side). How-
function expressions. These calculations were performed numerically ever, it is exceptionally worth to note that in the case of small values
by exploiting the fsolve() function of the python module scipy.optimize. of b, no radial diffusion contribution will ever dominate the flux.
In this manner the first 500 roots were calculated for each individual Therefore, the flux can be described as planar finite.8
case. In case of moderately sized values of b it can be seen that the diffu-
Since the functions f I ðt Þ; mI ðt Þ and M I ðt Þ are the only ones which sive flux will at first behave planar semi-infinite at small values of t.
are not given in terms of an infinite series, but in the form of an impro- Subsequently, it will increase owing to the radial diffusion contribu-
per integral, one might approximate the desired solution by perform- tion and follow the external cylindrical semi-infinite diffusion model.
ing a numerical integration i.e. by trapezoidal rule. Despite the fact Finally, once the concentration profile hits the impermeable wall
that this procedure is somewhat awkward since it supports an analytic located at r ¼ b, the diffusive flux will decay rapidly owing to a lack
result numerically, the computations usually provide fairly accurate of active substance. Consequently, the flux function f III ðt Þ will intersect
results. From Fig. 4(A)–(C) it can be seen that a computation of the with its planar semi-infinite analogue. Finally, in the case of very large
functions f I ðt Þ; mI ðt Þ and M I ðt Þ from the analytic expression (solid values of b, the diffusion can be described as external cylindrical semi-
lines) produces the same result as a computation by using the Talbot infinite and in the case of large values of a and b as planar semi-infinite
method for numerical inversion of Laplace transforms (dots). Further- at any experimentally relevant time instance.
more, it can be seen that for small values of t, and for large values of In contrast to the flux function, panel (B) of Fig. 5 depicts the func-
the electrode radius a, the functions f I ðt Þ; mI ðt Þ and M I ðt Þ merge with tion mIII ðt Þ. It can be seen that in a finite diffusion domain mIII ðt Þ will
their respective planar semi-infinite analoga (grey curves). This can be
explained by the fact that the curvature of the electrode is negligible in
direct proximity to its surface and that the zone of dilution has not 8
Actually, the point at which the diffusion might be described as planar finite reflective
expanded too much into the electrolyte at small values of t. Further- depends on both, a and b, since the radial diffusion contribution gets more pronounced if a
more, the curvature of the electrode becomes insignificant for large decreases.

7
T. Tichter et al. Journal of Electroanalytical Chemistry 897 (2021) 115565

Fig. 4. (A), (B), (D), (E) Double-logarithmic and (C), (F) normal representation of the functions f I ðt Þ; mI ðt Þ; f II ðt Þ; mII ðt Þ and M I ðt Þ and M II ðt Þ corresponding to an
external cylindrical semi-infinite diffusion domain (index I) and an internal cylindrical finite reflective diffusion domain (index II). Solid lines represent the
analytic results whereas dots correspond to numerical computations via Talbot’s method for numerical inversion of Laplace transformation which are shown as a
comparison. The grey line represents the analytic functions f ðt Þ; mðt Þ and M ðt Þ for a planar electrode in a semi-infinite diffusion domain. All simulations were
performed for ten different electrode radii, a, spanning a logarithmic grid from 10 μm to 1 mm, respectively. The diffusion coefficient was set to
D ¼ 1  105 cm2 s1 . In all cases, the functions merge with the result of a planar semi-infinite diffusion domain as t ! 0.

Fig. 5. (A), (B), (D), (E) Double-logarithmic and (C), (F) normal representation of the functions f III ðt Þ; mIII ðt Þ; f IV ðt Þ; mIV ðt Þ and M III ðt Þ and M IV ðt Þ corresponding to
an external cylindrical finite reflective (index III) and an external cylindrical finite transmissive (index IV) diffusion domain. Solid lines represent the analytic
results whereas dots correspond to numerical computations via Talbot’s method for numerical inversion of Laplace transformation, which are shown as a
comparison. The grey line represents the analytic functions f ðt Þ; mðt Þ and M ðt Þ for a planar electrode in a semi-infinite diffusion domain. All simulations were
performed for a constant electrode radius of a ¼ 40 μm and ten different diffusion domain sizes b spanning a logarithmic grid from a þ 10 μm to a þ 1 mm,
respectively. The diffusion coefficient was set to D ¼ 1  105 cm2 s1 . In all cases, the functions would merge with the result of a planar semi-infinite diffusion
domain as t ! 0 and a and b become very large and a ≪ b. However, in the present case, as b becomes very large, all functions merge at the scenario of an external
cylindrical semi-infinite diffusion space with an electrode of radius a ¼ 40 μm.

8
T. Tichter et al. Journal of Electroanalytical Chemistry 897 (2021) 115565

Fig. 6. Normalized CV curves simulated at a potential sweep rate of ν ¼ 2 mVs1 such that the experiment lasts a total time of t ¼ 1000 s. CVs were simulated on
the base of (A) M I ðt Þ, (B) M II ðt Þ, (C) M III ðt Þ and (D) M IV ðt Þ. In case of (A) and (B), ten different electrode radii, spanning a logarithmic grid from a ¼ 10 μm to
a ¼ 1 mm were employed, whereas in case of (C) and (D) an electrode radius of a ¼ 40 μm was used together with ten values of b spanning a logarithmic grid
which ranges from b ¼ a þ 10 μm to b ¼ a þ 1 mm. All CV simulations were performed on the base of the recurrence relation A.11, given in the appendix..

first deplete faster than the planar semi-infinite mðt Þ and will approach lized. In order to selectively illustrate the diffusion domain effect on
a limit as t increases. According to Eq. (48), this limit can be quantified the shape of the CV, the related current/voltage curves are calculated
 1 for reversible electrode kinetics and are normalized by their individual
as mIII ðt ! 1Þ ¼ 2aD1=2 b2  a2 . Since again M III ðt Þ is the
antiderivative of mIII ðt Þ, this limit will be the slope of M III ðt Þ. This is peak current. This results in the CV curves depicted in Fig. 6.
depicted in panel (C). In panel (A), corresponding to a semi-infinite diffusion domain, it can
The last scenario, the case of an external finite transmissive cylin- be seen that as a increases, the diffusion tail in the CV will decrease fas-
drical diffusion domain, is depicted in panels (D)–(F) of Fig. 5. In panel ter, since the radial diffusion contribution vanishes. In case of very large
(D), it can be seen that the flux function f IV ðt Þ again follows the planar a, the CV will be equivalent to the planar semi-infinite case.
semi-infinite f ðt Þ at t close to zero and the semi-infinite cylindrical f I ðt Þ Likewise, panel (B) depicts the CV curves under internal finite dif-
for large values of b. For small values of b, in turn f IV ðt Þ can be fusion conditions. It can be seen that the diffusion tail in the CV
described as planar finite transmissive diffusion. Finally, in case of decreases faster as a becomes smaller. This is caused by the fact that
intermediately sized b, the diffusion will follow the order of planar the diffusion zone inside the pore will get more finite and thus the lack
semi-infinite, external cylindrical semi-infinite and in the end of electrochemically active species becomes more dominant.
approach a limit owing to the transsmissive diffusion space. According Panel (C), corresponding to an external cylindrical finite reflective
diffusion domain, again represents the most complicated scenario –
to Eq. (55), this limit is quantified by mIV ðt ! 1Þ ¼ D1=2 ½a lnðb=aÞ1 .
particularly in case of the peak associated with the backward reaction.
Vice versa, mIV ðt Þ, depicted in panel (E) will deplete from the external
In case of very small and very large values b, however, the results are
cylindrical semi-infinite mI ðt Þ as t increases and b is in an intermediate
unambiguous and can be described by a planar finite reflective or an
range. Likewise, the related M IV ðt Þ will always be below (or at most
external cylindrical semi-infinite diffusion domain, respectively. In
equal) to its external cylindrical semi-infinite analogue as b increases
case of intermediate values of b, the depleting part of the CV first
and will approach the planar semi-infinite M ðt Þ if a and b are both cho-
becomes less steep, owing to the increasing diffusion domain size.
sen sufficiently large.
However, as b increases further, the magnitude of the backwards peak
Since panels (A) and (D) of Figs. 4 and 5 basically correspond to the
becomes smaller since the electrochemically active species formed at
normalized flux during a chronoamperometric experiment with sta-
the electrode surface diffuses away much faster owing to the radial
tionary boundary conditions, the next step will be the computation
diffusion.
of a related cyclic voltammetric experiment with a transient boundary.
Finally, in panel (D), corresponding to an external cylindrical finite
For this purpose, the respective functions M I ðt Þ to M IV ðt Þ will be uti-
transmissive diffusion domain, the results can be interpreted in a

9
T. Tichter et al. Journal of Electroanalytical Chemistry 897 (2021) 115565

π
straightforward way again. As b decreases, the depleting part in the CV K0 ðixÞ ¼  ½Y0 ðxÞ þ iJ0 ðxÞ ðA:7Þ
will become significantly less steep. This is caused by a constant mass 2
transport towards the electrode surface. Of course, as b becomes very π
large, again the state of an external cylindrical semi-infinite diffusion K1 ðixÞ ¼  ½J1 ðxÞ  iY1 ðxÞ ðA:8Þ
2
will be reached.
Identities (A.5) and (A.6) are graphically illustrated in OMS on
4. Summary and conclusions page 514 and identities (A.7) and (A.8) on pages 534 and 535, respec-
tively. Another, yet very important identity is given in AS9.1.16 in
This paper reports on analytical solutions for the chronoampero- terms of the Wronskian relation
metric current and the mass-transfer functions of cylindrical and hol- 2
low-cylindrical electrodes with semi-infinite and finite transmissive Yν ðxÞJνþ1 ðxÞ  Yνþ1 ðxÞJν ðxÞ ¼ : ðA:9Þ
πx
or reflective spatial boundary conditions. In case of finite diffusion
An important tool for contour integration is the residue theorem
domains, all solutions were given in the form of infinite series involv-
[25]. It states that the result of a complex integral is only determined
ing Bessel functions. In the case of semi-infinite diffusion domains, the
by the sum of the residues at the poles inside the closed and simple
solutions are given in terms of improper integrals. Until recently, many
contour C.
of these solutions were either not reported in the context of electro-
I
chemistry or not known in general. Limiting expressions for the mass
transfer functions or the diffusive flux are also provided, once they f ðzÞ dz ¼ 2πi ∑Res f ðzÞjz¼zk ðA:10Þ
C k
exist. Based on the analytical solutions presentend in this paper, it is
finally demonstrated that the NILT Talbot approach, which is used This relation is of utmost importance in this paper for scenarios II,
so far, provides an accurate and practical approach for approximating III and IV.
the time-dependent flux or mass transfer functions at cylindrical Once the mass transfer functions of a certain electrode geometry
electrodes. are derived, computing cyclic voltammetry responses at any degree
of reversibility can be done by exploiting the following approximation
pffiffiffiffi  
CRediT authorship contribution statement nFAcR;B D  1 þ eξðkΔt Þ  S
I ðkΔt Þ≈ pffiffiDffi eαξðkΔt Þ ; ðA:11Þ
k0
þ M i ðΔt Þ  ½1 þ eξðkΔt Þ 
Tim Tichter: Conceptualization, Methodology, Software, Valida-
tion, Visualization, Writing - original draft, Writing - review & editing. where S is defined by
Dirk Andrae: Formal analysis, Writing - original draft, Writing -
k1
review & editing. Christina Roth: Supervision, Writing - original
S ¼ ∑ I ð jΔt Þ½M i ððk  j þ 1ÞΔt Þ  M i ððk  jÞΔt Þ: ðA:12Þ
draft, Writing - review & editing, Funding acquisition. j¼1

Basically, relation (A.11) is a simplified version of the much more


Declaration of Competing Interest
general approximation given in Ref. [9] (cf. for derivation) and the
tool which was used for simulation of the CV curves in Fig. 6.
The authors declare that they have no known competing financial
interests or personal relationships that could have appeared to influ-
ence the work reported in this paper. References

[1] K. Aoki, K. Honda, K. Tokuda, H. Matsuda, Voltammetry at microcylinder


Appendix A electrodes: Part I. Linear sweep voltammetry, Journal of Electroanalytical
Chemistry and Interfacial Electrochemistry 182 (1985) 267–279, https://doi.
org/10.1016/0368-1874(85)87005-2.
This appendix contains fundamental mathematical definitions and
[2] K. Aoki, H. Kaneko, Theory of irreversible cyclic voltammograms at microcylinder
identities which are extensively used throughout the main manuscript. electrodes, Journal of Electroanalytical Chemistry and Interfacial Electrochemistry
These are basically extracted from the monographs of Oldham, Myland 247 (1988) 17–27, https://doi.org/10.1016/0022-0728(88)80127-X.
[3] J.W. Weidner, P.S. Fedkiw, Effect of ohmic, mass-transfer, and kinetic resistances
and Spanier [23] (referred to as OMS) and Abramowitz and Stegun
on linear-sweep voltammetry in a cylindrical-pore electrode, Journal of The
[24] (referred to as AS). Electrochemical Society 138 (1991) 2514–2526, https://doi.org/10.1149/
The Bessel functions Jn ðxÞ; Yn ðxÞ as well as the modified Bessel 1.2086011.
functions In ðxÞ and Kn ðxÞ are defined in AS9.1.10, AS9.6.10, [4] J.W. Weidner, Linear-sweep voltammetry of a soluble redox couple in a cylindrical
electrode, Journal of The Electrochemical Society 138 (1991) 258C–264C, https://
AS9.6.2 and AS9.1.2 as follows: doi.org/10.1149/1.2085824.
 x n 1 ð1Þj x2j
[5] J.W. Weidner, P.S. Fedkiw, Linear-sweep voltammetry in a cylindrical-pore
Jn ðxÞ ¼ ∑ ðA:1Þ electrode, Analytical Chemistry 164 (1992) 449–453, https://doi.org/10.1021/
2 j¼0 j! ðn þ jÞ! 2 ac00028a021.
[6] P.J. Mahon, K.B. Oldham, Incorporating electrode kinetics into the convolutive
 x n 1 1 x2j modeling of reactions at planar, cylindrical and spherical electrodes,
In ðxÞ ¼ ∑ ðA:2Þ Electrochimica Acta 46 (2001) 953–965, https://doi.org/10.1016/S0013-4686
2 j¼0 j! ðn þ j Þ! 2 (00)00680-0.
[7] L.K. Bieniasz, Automatic solution of integral equations pertinent to diffusion with
first order homogeneous reactions at cylindrical wire electrodes, Journal of
π Iν ðxÞ  Iν ðxÞ Electroanalytical Chemistry 674 (2012) 38–47, https://doi.org/10.1016/
Kn ðxÞ ¼ lim n ¼ 0; 1; 2; . . . ðA:3Þ
2 ν!n sinðνxÞ j.jelechem.2012.04.003.
[8] L.K. Bieniasz, Automatic solution of integral equations describing electrochemical
transients under conditions of internal cylindrical diffusion, Journal of
cosðπνÞJν ðxÞ  Jν ðxÞ
Yn ðxÞ ¼ lim n ¼ 0; 1; 2; . . . ðA:4Þ Electroanalytical Chemistry 700 (2013) 30–39, https://doi.org/10.1016/
ν!n sinðνxÞ j.jelechem.2013.04.010.
[9] T. Tichter, D. Andrae, J. Mayer, J. Schneider, M. Gebhard, C. Roth, Theory of cyclic
Some very important Bessel function identities, which have been voltammetry in random arrays of cylindrical microelectrodes applied to carbon felt
used in the derivations presented in the main manuscript, are: electrodes for vanadium redox flow batteries, Physical Chemistry Chemical Physics
21 (2019) 9061–9068, https://doi.org/10.1039/C9CP00548J.
I0 ðixÞ ¼ J0 ðxÞ ðA:5Þ [10] T. Tichter, J. Schneider, D. Andrae, M. Gebhard, C. Roth, Universal algorithm for
simulating and evaluating cyclic voltammetry at macroporous electrodes by
considering random arrays of microelectrodes, ChemPhysChem 21 (2019)
I1 ðixÞ ¼ iJ1 ðxÞ ðA:6Þ 428–441, https://doi.org/10.1002/cphc.201901113.

10
T. Tichter et al. Journal of Electroanalytical Chemistry 897 (2021) 115565

[11] L. Landon-Lane, A.J. Downard, A.T. Marshall, Single fibre electrode measurements Electroanalytical Chemistry 614 (2007) 121–130, https://doi.org/10.1016/
- A versatile strategy for assessing the non-uniform kinetics at carbon felt j.jelechem.2007.11.010.
electrodes, Electrochimica Acta 354 (2020), https://doi.org/10.1016/ [19] C. Montella, J.P. Diard, New approach of electrochemical systems dynamics in the
j.electacta.2020.136709 136709. time-domain under small-signal conditions. I.A family of algorithms based on
[12] A. Lasia, Electrochemical Impedance Spectroscopy and its Applications, Springer, numerical inversion of Laplace transforms, Journal of Electroanalytical Chemistry
2014, https://doi.org/10.1007/978-1-4614-8933-7. 623 (2008) 29–40, https://doi.org/10.1016/j.jelechem.2007.11.010.
[13] M. Störzbach, J. Heinze, The Crank Nicolson Technique – an efficient algorithm for [20] C. Montella, Re-examination of the potential-step chronoamperometry method
the simulation of electrode processes at macro- and microelectrodes, Journal of through numerical inversion of Laplace transforms. I. General formulation and
Electroanalytical Chemistry 346 (1993) 1–27, https://doi.org/10.1016/0022- numerical solution, Journal of Electroanalytical Chemistry 633 (2009) 35–44,
0728(93)85001-W. https://doi.org/10.1016/j.jelechem.2009.04.019.
[14] K.B. Oldham, Convolution: a general electrochemical procedure implemented by a [21] H.S. Carslaw, J.C. Jaeger, Conduction of Heat in Solids, Oxford University Press,
universal algorithm, Analytical Chemistry 58 (1986) 2296–2300, https://doi.org/ 1959.
10.1021/ac00124a040. [22] E. Laborda, J. González, F. Martínez-Ortiz, A. Molina, Voltammetry at
[15] A. Talbot, The accurate numerical inversion of Laplace transforms, Journal of microelectrodes of reversible electrode reactions with complex stoichiometry: A
Applied Mathematics 23 (1979) 97–120, https://doi.org/10.1093/imamat/ general analytical theoretical framework, Journal of Electroanalytical Chemistry
23.1.97. 872 (2020), https://doi.org/10.1016/j.jelechem.2020.113932 113932.
[16] J.A.C. Weideman, N.L. Trefethen, Parabolic and hyperbolic contours for [23] K. Oldham, J. Myland, J. Spanier, An Atlas of Functions, Springer US, 2009,
computing the Bromwich integral, Mathematics of Computation 76 (2007) https://doi.org/10.1007/978-0-387-48807-3.
1341–1356, https://doi.org/10.1090/S0025-5718-07-01945-X. [24] M. Abramowitz, I.A. Stegun, Handbook of Mathematical Functions with Formulas,
[17] B. Dingfelder, J.A.C. Weideman, An improved Talbot method for numerical Graphs, and Mathematical Tables, 10th printing., Dover Publications, New York,
Laplace transform inversion, Numerical Algorithms 68 (2015) 167–183, https:// 1972.
doi.org/10.1007/s11075-014-9895-z. [25] G. Stephenson, P.M. Radmore, Advanced Mathematical Methods for Engineering
[18] C. Montella, LSV modelling of electrochemical systems through numerical and Science Students, Cambridge University Press, 1990, https://doi.org/
inversion of Laplace transforms. I - The GS-LSV algorithm, Journal of 10.1017/cbo9781139168120.

11

You might also like