Journal of Electroanalytical Chemistry: Hatem M.A. Amin, Yuki Uchida, Enno Kätelhön, Richard G. Compton

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Journal of Electroanalytical Chemistry 880 (2021) 114891

Contents lists available at ScienceDirect

Journal of Electroanalytical Chemistry


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

Determination of standard electrochemical rate constants from semi-circular


sweep voltammetry: A combined theoretical and experimental study
Hatem M.A. Amin a,b, Yuki Uchida a, Enno Kätelhön a,1, Richard G. Compton a,

a
Department of Chemistry, Physical & Theoretical Chemistry Laboratory, University of Oxford, South Parks Road, Oxford, OX1 3QZ, United Kingdom
b
Chemistry Department, Faculty of Science, Cairo University, 12613, Giza, Egypt

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

Article history: Semi-circular potential sweep voltammetry has recently been proposed as a sensitive method to estimate electrochem-
Received 19 October 2020 ical rate constants [J. Electroanal. Chem., 835 (2019) 60–66]. Herein, this approach is experimentally verified using
Received in revised form 24 November 2020 microdisc electrode voltammetry. The reduction of hexaamineruthenium (III) chloride and the oxidation of
Accepted 25 November 2020
ferrocenemethanol, both in aqueous solution, at carbon electrodes were studied. The rate constants obtained from
Available online 28 November 2020
semi-circular voltammetry for both investigated redox systems are in agreement with independent values obtained
Keywords:
from the conventional Nicholson method at macroelectrodes. The method compares favourably in terms of simplicity
Standard electrochemical rate constant and speed as it requires recording only a single voltammogram. This approach provides a novel, fast complementary
Semi-circular sweep voltammetry method for the evaluation of the standard electrochemical rate constant.
[Ru(NH3)6]Cl3 reduction
FcCH2OH oxidation
Kinetics

1. Introduction or nanoelectrodes [8,9]. In their method, ko is determined directly from


two easily accessible experimental parameters, (E1/4 − E1/2) and (E1/2 −
Measuring the rate of electron transfer is crucial in understanding di- E3/4), where E1/2 is the half-wave potential and E1/2 and E3/4 are quartile
verse electrochemical reactions including systems of fundamental, biologi- wave potentials [8]. The use of microelectrodes is advantageous for mea-
cal and medical relevance [1–3]. The standard electrochemical rate suring fast electron transfer kinetics as they increase the mass transport
constant (ko) is the important parameter which quantifies the reversibility rate (so accessing faster rate constants) and minimize the effects of ohmic
of the electron transfer and the kinetics of the reaction [2,4,5]. Linear (IR) drop in solution, the double layer charging and the reactant adsorption
sweep- and cyclic voltammetry at macroelectrodes are commonly used to [10–12]. Linear cyclic voltammetry (CVs) at microelectrodes but at larger
measure ko values [2,4,6]. Specifically, from cyclic voltammetry data at a scan rates together with simulations could also be used to determine ko,
macroelectrode, the standard rate constant of electron transfer is routinely however the attraction of using lower scan rates and accounting for the con-
determined using the method of Nicholson [7]. In the Nicholson procedure, vergent diffusion allows measurements at scan rates essentially free from
ko is calculated based on the peak-to-peak potential separation (ΔE) at var- capacitive currents [13]. Transient methods at microelectrodes have also
ious scan rates (υ) according to the following Eq. [7]: been reported [14–17].
In our previous work, we have presented the theoretical treatment of an
½πDox υðnF=RT Þ1=2 alternative method of voltammetry that employs a semi-circular potential
ko ¼ Ѱ ð1Þ waveform rather than the typical and conventional linear waveform
ðDox =Dred Þα=2
[18,19]. Recently, we have experimentally implemented the semi-circular
sweep voltammetry (SCV) to determine the formal potential of a redox sys-
where Ѱ is the “kinetic parameter”, Dox and Dred are the diffusion co-
tem [20]. This theory was further developed via simulation to give a basis
efficients of the oxidized and reduced species, respectively, υ is the scan
for measuring electrode kinetics [21]. In this paper we present an experi-
rate, F is the Faraday constant, R is the ideal gas constant, T is the temper-
mental verification of the semi-circular sweep voltammetry at a microelec-
ature, n is the number of electrons transferred, and α is the transfer coeffi-
trode and its application to evaluate the ko values of the aqueous redox
cient. Another approach has been introduced by Mirkin and Bard for
couples Ru(NH3)2+/3+ and FcCH2OH0/1+. These redox systems were
calculating the parameter ko based on steady-state voltammetry at micro- 6

⁎ Corresponding author.
E-mail address: richard.compton@chem.ox.ac.uk. (R.G. Compton).
1
Current address: MHP Management- und IT-Beratung GmbH, Office Frankfurt, Edmund-Rumpler-Strasse 3, Alpha Rotex, 60,549 Frankfurt am Main, Germany.

http://dx.doi.org/10.1016/j.jelechem.2020.114891
1572-6657/© 2020 Elsevier B.V. All rights reserved.
H.M.A. Amin et al. Journal of Electroanalytical Chemistry 880 (2021) 114891

selected for study at carbon electrodes as they have relatively fast electron different from the conventional linear wave. As shown in Fig. 1a, the linear
transfer kinetics and likely involve outer-sphere electron transfer [9,11]. potential wave exhibits a constant scan rate over the entire potential sweep,
Satisfactory agreement between the results obtained from this new method whilst the semi-circular wave shows significant changes in the scan rate
at microelectrodes and the traditional Nicholson method at macro- throughout the potential sweep. At the mid-point of the wave (Ecentre), in-
electrodes was seen. In addition, the voltammograms obtained experimen- stantaneously, an infinite scan rate is reached (in principle) while slower
tally show good agreement with the simulated ones, further validating the scan rates are seen towards the limits of the potential window, see
theory given in [21]. Fig. S1. The characteristics of the semi-circular potential wave result in se-
lective amplification of the peak current if Ecentre, corresponding to the max-
2. Semi-circular potential wave sweep voltammetry (SCV) imum scan rate, is set within the close vicinity of the formal potential of the
redox couple of interest, see Fig. 1b. Comparing the two voltammograms in
In this section, we describe briefly the method and the waveform Fig. 1b, the LSV shows a near steady-state response whilst the SCV shows a
employed in our work. Details of the theoretical model can be found in peak feature corresponding to the region of faster than average scan rates.
our previous reports [19,21]. In linear sweep voltammetry, the potential
waveform is expressed as:
3. Experimental
EðtÞ ¼ Eðt ¼ 0Þ  υt ð2Þ
3.1. Materials
where the potential of the working electrode E(t) is swept linearly from an
initial potential E (t = 0) with a constant scan rate (υ). Recently, Uchida Ferrocenemethanol (C11H12FeO, “FcCH2OH”, 97%, ChemCruz®, Santa
et al. have proposed the semi-circular potential waveform as an alternative Cruz Biotechnology, Inc., USA), Hexaammineruthenium(III) chloride ([Ru
wave and can be described as follows [19–21]: (NH3)6]Cl3, Ru 32.1% min, Alfa Aesar, UK), potassium hexacyanoferrate
(K3[Fe(CN)6], 98 + %, Lancaster Synthesis, UK) and potassium chloride
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
  2 (KCl, >99%, Aldrich) were used without further purification. All solutions
t
EðtÞ ¼ Ao j 1− 2 −c j þ Ecentre ð3Þ were prepared using Milli-Q water with a resistivity of 18.2 MΩ cm−1 at
t max 298 K (Millipore Water Systems, UK).

where Ao is the amplitude of the potential wave, t is the time and tmax is the
duration of a single potential sweep between the initial potential Ei and the 3.2. Electrochemical measurements
final potential Ef. Experimentally, the duration of the potential scan tmax de-
fines the applied average scan rate. c = 0 for t ≤ tmax/2 and c = 1 for tmax/ All electrochemical measurements were carried out with an in-house
2 < t < tmax. Ecentre is the average of the minimum and maximum values of made potentiostat. Details of the potentiostat can be found elsewhere
the applied potential (i.e. the mid-point potential of the applied potential [22,23]. This is a low-noise potentiostat with a nominal 60 μV interval of
window in this context) with respect to the reference electrode and it can applied potential on the working electrode. A variable gain sub-femto Am-
be expressed as follows: pere current amplifier DDPCA-300 (FEMTO® Messtechnik GmbH, Berlin,
Germany) was used to measure the current. Two cascaded 2 kHz passive
Ei þ E f RC-filters were used to remove high frequency components from the cur-
Ecentre ¼ ð4Þ
2 rent signal. For data acquisition, a USB-6003 DAQs device (National Instru-
ments, TX, USA) with 32-bit resolution was used. The resulting analogue
Fig. 1a displays the linear and semi-circular potential waveforms inves- signal was digitized at a sampling rate of 100 kHz. Control of the
tigated. So as to compare linear sweep voltammetry (LSV) and semi-circular potentiostat was run through a software written in Python. The data were
sweep voltammetry (SCV) whilst ensuring that the average scan rate over a collected with Python and then were plotted using OriginPro 2017. To
single semi-circular wave is the same as the constant scan rate of the linear record a semi-circular voltammogram, we set the following parameters
wave, the duration and limits of the sweep are kept the same for both in the software: Ecentre, amplitude of 0.514 V, duration of the scan that
waves. The semi-circular wave shows specific features that are obviously is equivalent to the desired average scan rate. Before recording each

Fig. 1. (a) Linear and semi-circular waveforms. The semi-circular wave shown has an amplitude of 0.514 V and a scan duration of 67 s which corresponds to 15 mV s−1 in the
linear wave. (b) Recorded experimental voltammograms corresponding to the two waves shown on (a) obtained at a carbon microdisc electrode (23.5 μm diameter) in
1.0 mM [Ru(NH3)6]Cl3/0.1 M KCl solution with an average scan rate of 15 mV s−1. The arrow shows the direction of potential sweep.

2
H.M.A. Amin et al. Journal of Electroanalytical Chemistry 880 (2021) 114891

voltammogram, the potential of the working electrode was kept at the the value of 59 mV expected at 25 °C for a fully reversible system. The
starting potential for 3 s to attain equilibrium. formal potential can be estimated from the midpoint potential between
A single-compartment, three-electrode glass cell with a carbon elec- the oxidation and reduction peak potentials assuming that the species in
trode as the working electrode, Pt wire (GoodFellow, Cambridge, UK) as the redox couple have identical diffusion coefficients. A value of Eof =
the counter electrode and a saturated calomel electrode (SCE, ALS Co. −0.182 ± 0.002 V was obtained. The error bar represents the uncer-
Ltd., Japan) as the reference electrode was used. Two types of carbon tainty in determining the peak potentials. This formal potential is used
working electrodes from ALS Co. Ltd., Japan, were used: a glassy below for analysing the semi-circular sweep voltammograms. The ob-
carbon (GC) macrodisc electrode with 3.0 mm diameter and a carbon tained Eof is very close to the reported value of −0.178 V for the same
fibre microdisc electrode with a nominal 33 μm diameter. An accurate redox system [26].
radius of the microdisc electrode was determined from the diffusion- The kinetic parameter Ѱ was then calculated from the peak-to-peak sep-
limited current in a steady-state voltammogram recorded in 1.0 mM aration (ΔE is in mV) according to the following empirical Eq. [27]:
[Ru(NH 3 ) 6 ]Cl3 /0.1 M KCl solution, see Fig. S2 in the SI. The radius
was found to be 23.5 ± 0.2 μm, using the known diffusion coefficient ð−0:6288 þ 0:0021 nðΔEÞÞ
Ѱ¼ ð6Þ
[23], which is larger than the nominal value given by the manufacturer. ð1−0:017 nðΔEÞÞ
Before running each experiment, the microelectrode was polished with
MicroPolish alumina slurries with successively smaller sizes of alumina In our calculations, we assumed α = 0.5 and Dox = Dred. A diffusion co-
powder (1, 0.3 and 0.05 μm from Buehler, IL, USA) applied on three sep- efficient D = 8.43 × 10−10 m2 s−1 for [Ru(NH3)6]3+ was used. This is
arate soft lapping pads (Buehler, USA). After each polishing step, the close to but not exactly the same as that of [Ru(NH3)6]2+ (D =
electrode was rinsed with Milli-Q water and finally was dried under a 11.9 × 10−10 m2 s−1) in 0.1 M KCl [23]. Fig. 2c displays the plot of Ѱ
nitrogen flow. versus ΔE for the various scan rates. It is clearly seen that Ѱ decreases as
The electrochemical cell was kept at 25 °C throughout the entire ex- ΔE decreases. The heterogeneous rate constant ko is calculated according
periment by using a water bath with a temperature controller (SCT1, to Nicholson equation (Eq. 1) where the slope of the plot of Ѱ versus
Stuart, UK). The whole setup was placed in a Faraday cage. All solutions [πDnF/(RT)]−1/2v−1/2 gives the value of ko. As shown in Fig. 2d, the
were purged with nitrogen (99.998%, BOC, UK) for at least 30 min prior slope of the straight line for [Ru(NH3)6]2+/3+ system gives ko =
to each experiment and were surrounded by nitrogen during the 0.046 ± 0.001 cm s−1. The error bar represents the standard error in the
measurement. value of the slope.
The simulations in this work were carried out using C++ with OpenMP The [Ru(NH3)6]2+/3+ system was also investigated in a higher
by applying the Alternating Direction Implicit method as previously described supporting electrolyte concentration, namely 1.0 M KCl, as shown in
[13,24,25]. Fig. S4. The potential was swept from +0.302 V to −0.726 at scan rates
from 25 to 300 mV s−1. Eof was found to be −0.212 ± 0.002 V in 1.0 M
4. Results and discussion KCl which is more negative than in 0.1 M KCl reflecting the change of for-
mal potential with the concentration. After analysis of the voltammograms
In this section, we first determine the standard electrochemical rate con- according to the Nicholson method, a value of ko = 0.099 ± 0.006 cm s−1
stant for both the [Ru(NH3)6]2+/3+ and FcCH2OH0/1+ redox couples from was obtained using a literature value of D = 5.5 × 10−10 m2 s−1 [23].
cyclic voltammetry data at GC macroelectodes using the Nicholson ap- The lower D value in 1.0 M KCl than in 0.1 M KCl is probably in part due
proach. We also determine the formal potential of these systems from the to the increased viscosity of the solution with the higher ionic strength
corresponding cyclic voltammograms. Next, we discuss the voltammetric [23]. The ko value in 1.0 M KCl is higher than the value obtained for the
features arising from the application of the semi-circular potential wave- lower supporting electrolyte concentration; The influence of the electrolyte
form at a microelectrode and finally deduce the rate constants for both on the kinetics reflects the nature of the redox couple, as has previously
redox systems based on these experimental data and the theoretical been reported [11,28,29].
model for semi-circular potential sweep voltammetry. Comparisons are
then made. 4.2. Evaluation of the standard electrochemical rate constant at a GC electrode
for ferrocenemethanol from cyclic voltammetry data
4.1. Evaluationof thestandardelectrochemical rateconstant for [Ru(NH3)6]2+/3+
at a GC electrode from macroelectrode cyclic voltammetry data Next, we studied the oxidation of ferrocenemethanol in aqueous solu-
tion, again at a GC (3.0 mm) macroelectrode. In this reaction, FcCH2OH
In this section, we investigate the one-electron reduction of [Ru(NH3)6]3+ is oxidized to FcCH2OH+:
to [Ru(NH3)6]2+ at a planar macrodisc GC electrode:
FcCH 2 OH⇌FcCH 2 OH þ þ e− ð7Þ
3þ − 2þ
RuðNH 3 Þ6 þ e ⇌ RuðNH 3 Þ6 ð5Þ
Fig. 3a shows the voltammetric behaviour obtained at a GC
We first focus on the estimation of the rate constant using the Nich- macroelectrode in 1.04 mM FcCH2OH/ 0.10 M KCl solution over the scan
olson procedure which is based on the peak-to-peak potential separa- rate of 10 to 300 mV s−1. For recording these voltammograms, the poten-
tion in cyclic voltammograms at various scan rates as discussed in the tial was swept from 0.330 V to +0.698 V. The formal potential was esti-
Introduction section [7]. mated from the mid-point potential of the CVs in Fig. 3a and was found
Fig. 2a shows cyclic voltammograms obtained at the GC (3.0 mm) to be Eof = +0.185 V, which is close to the literature value of ~0.200 V
electrode in 1.0 mM [Ru(NH3)6]Cl3/0.1 M KCl at scan rates from 25 to at GC in 0.1 M phosphate buffer solution (PBS) [1]. The voltammograms,
500 mV s−1. Faradaic-only currents are obtained by subtraction of the after correction for the capacitive background, were analysed and the
capacitive currents measured in 0.1 M KCl solution. Fig. S3 shows exam- plots of ΔE vs. log υ and Ѱ vs. ΔE are depicted in Fig. 3b and c, respectively.
ples of the voltammograms obtained at 25 and 500 mV s−1 before and Again, as the scan rate increases, ΔE increases and accordingly Ѱ decreases.
after subtraction of the capacitive currents. The potential was swept The Nicholson method is then applied to estimate ko for ferrocenemethanol
from +0.332 V to −0.696 V and reduction and oxidation peaks are ob- redox reaction. As shown in Fig. 3d, a plot of Ѱ vs. [πDnF/(RT)]−1/2υ−1/2
served. The oxidation and reduction peak potentials slightly shift, but in gave a linear regression with r = 0.999, indicating that the response of this
opposite directions, as the scan rate increases. Accordingly, the peak-to- redox system is governed by the Nicholson eq. A diffusion coefficient D =
peak separation (ΔE) increases with the scan rate increase as shown in 7.8 × 10−10 m2 s−1 for FcCH2OH0/1+ was used [9]. The resulting slope
the plot of ΔE vs. log υ in Fig. 2b. The peak separation is larger than yields ko = 0.057 ± 0.001 cm s−1 for the FcCH2OH0/1+ redox couple.

3
H.M.A. Amin et al. Journal of Electroanalytical Chemistry 880 (2021) 114891

Fig. 2. (a) Cyclic voltammograms recorded at GC (3.0 mm) macroelectrode in 1.0 mM [Ru(NH3)6]Cl3/0.1 M KCl at scan rates from 25 to 500 mV s−1 at 298 K. The plotted
currents are corrected for capacitive current contributions by subtraction of the currents obtained in blank 0.1 M KCl solution. (b) Plot of ΔE vs. log υ for the corresponding
data. (c) Plot of Ѱ versus ΔE. (d) Plot of Ѱ versus [πDnF/(RT)] −1/2v−1/2 for the same data.

4.3. Voltammetric behaviour of linear vs. semi-circular potential waveforms response obtained using linear potential wave is different from that ob-
tained from using the semi-circular wave: The linear sweep voltammetry
We next report the voltammetric behaviour generated from semi- shows almost no peak but a typical plateau corresponding to the near
circular potential waveforms and compare it with linear sweep voltamme- steady-state, diffusion limited, behaviour at potentials more negative to
try. Fig. 4 displays the voltammetric response obtained from the application ca. 0.25 V, whilst the semi-circular sweep voltammogram reveals a sharp
of the linear and semi-circular waveforms in 1.0 mM [Ru(NH3)6]Cl3/ 0.1 M peak. The highest currents measured are significantly larger for the semi-
KCl at carbon microdisc (23.5 μm radius) electrode. circular wave than the traditional linear sweep. Near Ecentre = Eof, the cur-
In the linear sweep voltammogram (dark yellow line), the potential was rent reaches a maximum and rapidly fall off at all other potentials. This is
swept from +0.332 to 0.696 V at a constant scan rate of 15 mV s−1. To run caused by the singularity of the scan rate at the midpoint of the semi-
the equivalent experiment using the semi-circle potential wave (red line) circular wave as shown in Fig. 1 and Fig. S1 and as expected theoretically
within the same potential window, the amplitude was set at 0.514 V and [18–21]. Simulated currents using the semi-circular potential wave are
the duration of the potential scan was set at 67 s such that the average displayed in the black dashed line and are used to calculate ko value
scan rate over the entire sweep equals 15 mV s−1. Semi-circular sweep volt- below. Fig. 5 shows the response of [Ru(NH3)6]Cl3/1.0 M KCl containing
ammetry gives the largest peak current signal when Ecentre = Eof for revers- a higher concentration of supporting electrolyte; a similar response as
ible systems. Therefore, this voltammogram was recorded at Ecentre = Eof = seen in Fig. 4 is observed.
0.182 V (obtained from the midpoint of the CVs in Fig. 2a). The scan rate of
15 mV s−1 corresponds to the dimensionless scan rate of σ = 0.389 used to 4.4. Determination of the standard rate constants for the [Ru(NH3)6]Cl2+/3+
3
generate theoretical simulations as described elsewhere: [13]. and ferrocenemethanol0/1+ redox couples using semi-circular sweep
 2   voltammetry
r F
σ¼ υ ð8Þ
D RT The data of the semi-circular potential sweep voltammetry was
analysed using our previously reported theoretical model [21]. Provided
The Faradaic-only currents (green line) were obtained by subtracting that the parameters D, C, r, Eof are known for the redox system, only a single
the capacitive currents recorded in blank 0.1 M KCl solution (blue line) voltammogram recorded using SCV (red line shown in Fig. 4) with Ecentre =
from the measured currents in [Ru(NH3)6]Cl3/0.1 M KCl solution (red Eof is needed to determine the ko value if a value of the transfer coefficient is
line), as shown in Fig. 4. From Fig. 4, it is clear that the voltammetric assumed. Herein we take for the transfer coefficients α = β = 0.5. The

4
H.M.A. Amin et al. Journal of Electroanalytical Chemistry 880 (2021) 114891

Fig. 3. (a) Cyclic voltammograms recorded at GC (3.0 mm) macroelectrode in 1.04 mM FcCH2OH/0.1 M KCl at scan rates from 10 to 300 mV s−1 at 298 K. The Faradaic
currents are corrected for capacitive current contribution by subtraction of the currents obtained in blank 0.1 M KCl solution. (b) plot of ΔE vs. log υ for the
corresponding data. (c) plot of Ѱ vs. ΔE. (d) plot of Ѱ vs. [πDnF/(RT)]−1/2v−1/2 for the same data.

semi-circular potential sweep voltammograms were analysed to obtain ko. using the semi-circular potential sweep voltammetry and the Nicholson
Simulations using the semi-circular potential wave were carried out to method are summarized and compared to the literature values as shown
find the ko value that generates the best fit of theory to the part of the volt- in Table 1. As can be seen in Table 1, the rate constant obtained for [Ru
ammogram where electrode kinetics dominate (i.e. at potentials more pos- (NH3)6]2+/3+/ 0.1 M KCl is smaller (23 times) than the value obtained
itive to 0.18 V for the 1.0 mM [Ru(NH3)6]Cl3/0.1 M KCl). All simulated for [Ru(NH3)6]2+/3+/ 1.0 M KCl solution where the effect of the
data were obtained using σ = 0.389, α = 0.5 and Ecentre = Eof. The simu- supporting electrolyte has been previously noted [11,28].
lated currents (dashed black line in Fig. 4) for 1.0 mM [Ru(NH3)6]Cl3/ The ko values obtained using semi-circular sweep voltammetry are in
0.1 M KCl solution best fit the experimental results (green line) in the satisfactory agreement with the values obtained independently from the cy-
kinetic-dominated range of +0.22 to 0.18 V when ko = 0.037 ± 0.01 V. clic voltammetric data using the Nicholson method under the same experi-
The error bar reflects the range of ko values that still give a degree of fit mental conditions, see Table 1. As the semi-circular potential sweep
to the experimental curve. Fig. S5 shows the simulated voltammograms ob- voltammetry method needs (in principle) the recording of only one voltam-
tained when a ko = 0.02 cm s−1 was used in the simulation. This mogram (as well as knowing the formal potential of the redox system), it of-
0.02 cm s−1 value very clearly does not show a high quality of fitting con- fers speed and simplicity over the traditional Nicholson method and can
sistent with the estimated error. At potentials beyond the maximum of the usefully complement the latter measurement.
semi-circular voltammograms an incomplete match between experimental The second studied redox system is the oxidation of FcCH2OH to
and simulated data is observed. The discrepancy between theory and exper- FcCH2OH+. Fig. 6 shows the voltammograms obtained using both linear
iment necessarily increases as the potential is swept as the modelling is pro- and semi-circular waves at a microdisc electrode in 1.04 mM FcCH2OH/
gressive, moving step by step through the voltammogram. Thus, tiny 0.10 M KCl solution. In this case the potential was swept from 0.330 to
differences between theory and experiment progressively accumulate. Nev- +0.698 V at an average scan rate of 14 mV s−1 (i.e. sweep duration =
ertheless, fitting of theory to the experimental data in the electrode kinetics 73 s). This average scan rate was selected in order to have the same dimen-
dominated part of the voltammograms provides an effective alternative sionless scan rate of σ = 0.389 as in the case of the [Ru(NH3)6]2+/3+ sys-
method to determine ko values for redox systems. tem. The general voltammetric features arising from the semi-circular
For the solution 1.01 mM [Ru(NH3)6]Cl3/1.0 M KCl, the simulated cur- sweep voltammetry in FcCH2OH0/1+ redox couple are similar to that ob-
rents (dashed black line in Fig. 5) fit very well with the experimental data served in the [Ru(NH3)6]2+/3+ system. Again, a significant amplification
(green line) in the range of +0.2 to 0.2 V, when ko = 0.11 ± of the current is observed at the close vicinity of Ecentre which is set at
0.01 cm s−1. The ko values obtained for the [Ru(NH3)6]2+/3+ redox system Eof = +0.185 V. In the linear sweep voltammogram, a near steady-state

5
H.M.A. Amin et al. Journal of Electroanalytical Chemistry 880 (2021) 114891

Fig. 6. Voltammograms obtained using a semi-circular potential waveform at a


Fig. 4. Voltammograms obtained using a semi-circular potential waveform at a carbon microdisc electrode in 0.10 M KCl (blue line) and in 1.04 mM FcCH2OH/
carbon microdisc electrode in 0.1 M KCl (blue line) and in 1.0 mM [Ru(NH3)6] 0.10 M KCl solution (red line, Faradaic+ capacitive currents) at Ecentre =
Cl3/0.1 M KCl solution (red line, Faradaic+ capacitive currents) at Ecentre = +0.185 V, amplitude = 0.514 V, duration of scan 73 s and an average scan rate
0.182 V, amplitude = 0.514 V, duration of scan 67 s and an average scan rate of of 14 mV s−1. The capacitive current corrected voltammogram for FcCH2OH/KCl
15 mV s−1. The capacitive current corrected voltammogram for [Ru(NH3)6]Cl3/ is shown in the green line. The linear sweep voltammogram in the same solution
KCl is shown in the green line. The linear sweep voltammogram in the same at 14 mV s−1 is displayed as a dark yellow line. The simulated Faradaic currents
solution at 15 mV s−1 is displayed as a dark yellow line. The simulated Faradaic which are obtained using a dimensionless scan rate σ = 0.389 are shown in the
currents obtained using a dimensionless scan rate σ = 0.389 are shown in the dashed black line. The definition of the dimensionless parameters can be found in
dashed black line. The definition of the dimensionless parameters can be found in our previous paper [21]. The arrow shows the direction of potential sweep.
our previous paper [21]. The arrow shows the direction of potential sweep.

behaviour with no peaks is observed. For the FcCH2OH0/1+ redox system,


in the potential range of 0.22 to +0.18 V, the simulated currents for the
semi-circular sweep voltammetry (dashed black line) are in good agree-
ment with the measured Faradaic-only currents (green line) when ko =
0.055 ± 0.01 cm s−1 is used, see Fig. 6. Table 1 summarises the ko values
for both studied systems, [Ru(NH3)6]2+/3+ and FcCH2OH0/1+, determined
using both the Nicholson method and the semi-circular sweep voltammetry
as well as Ko values reported in literature.
Overall Table 1, for all the studied redox systems, the rate constants ob-
tained using the semi-circular sweep voltammetry agree with that obtained
using the Nicholson method and with literature where literature data exist
for the same electrolyte.

5. Conclusions

In this work we experimentally investigated the voltammetric behaviour


of redox systems at microelectrodes while applying the semi-circular poten-
tial wave and compared this response with the response from linear sweep
voltammetry. The peak current obtained using semi-circular sweep wave
Fig. 5. Voltammograms obtained using a semi-circular potential waveform at a
is much larger than that obtained from linear sweep wave when the poten-
carbon microdisc electrode in 1.0 M KCl (blue line) and in 1.01 mM [Ru(NH3)6]
Cl3/1.0 M KCl solution (red line, Faradaic+ capacitive currents) at Ecentre =
tial window is centred within the close vicinity of the formal potential of the
0.212 V, amplitude = 0.514 V, duration of scan 103 s and an average scan rate of redox system. Semi-circular potential voltammetry was exploited to quanti-
10 mV s−1. The capacitive current corrected voltammogram for [Ru(NH3)6]Cl3/ tatively study the kinetics of electron transfer. The [Ru(NH3)6]2+/3+ and
KCl is shown in the green line. The linear sweep voltammogram in the same FcCH2OH0/1+ redox couples were selected as examples and the electro-
solution at 15 mV s−1 is displayed as a dark yellow line. The simulated Faradaic chemical rate constants were determined via SCV and independently from
currents obtained using a dimensionless scan rate σ = 0.389 are shown in the cyclic voltammetry data using Nicholson method. The results showed satis-
dashed black line. The definition of the dimensionless parameters can be found in factory agreement between the two methods. A practical advantage of the
our previous paper [21]. The arrow shows the direction of potential sweep. new method is the need (in principle) of the recording of only a single

Table 1
Summary of the standard heterogeneous rate constants for [Ru(NH3)6]2+/3+ and FcCH2OH0/+1 redox systems obtained at carbon electrode and compared to literature
values. The experiments were done such that σ = 0.389 for both systems.
Redox system Avg. scan rate (mV s−1) ko (cm s−1) Ref.

From semi-circular voltammetry From Nicholson equation From literature @GC

1.00 mM Ru(NH3)6]Cl3/0.1 M KCl 15 0.037 ± 0.01 0.046 ± 0.001 – –


1.01 mM Ru(NH3)6]Cl3/1.0 M KCl 10 0.11 ± 0.01 0.099 ± 0.006 0.13 ± 0.15 (in 1 M KCl) [28]
1.04 mM FcCH2OH/0.1MKCl 14 0.055 ± 0.01 0.057 ± 0.011 0.19 ± 0.05 (in 0.1 M PBS) [1]

6
H.M.A. Amin et al. Journal of Electroanalytical Chemistry 880 (2021) 114891

voltammogram to determine the rate constant. This renders the method a [12] N.V. Rees, O.V. Klymenko, E. Maisonhaute, B.A. Coles, R.G. Compton, The application
of fast scan cyclic voltammetry to the high speed channel electrode, J. Electroanal.
fast and complementary route to the classical procedures. Chem. 542 (2003) 23–32.
[13] R.G. Compton, E. Kätelhön, E. Laborda, K.R. Ward, Understanding Voltammetry: Simu-
Declaration of Competing Interest lation of Electrode Processes, 2nd ed. World Scientific, 2020.
[14] J.O. Howell, R.M. Wightman, Ultrafast voltammetry and voltammetry in highly resis-
tive solutions with microvoltammetric electrodes, Anal. Chem. 56 (1984) 524–529.
The authors declare that they have no known competing financial inter- [15] C. Amatore, C. Lefrou, F. Pflüger, On-line compensation of ohmic drop in
ests or personal relationships that could have appeared to influence the submicrosecond time resolved cyclic voltammetry at ultramicroelectrodes, J.
Electroanal. Chem. Interfacial Electrochem. 270 (1989) 43–59.
work reported in this paper.
[16] C. Amatore, E. Maisonhaute, When voltammetry reaches nanoseconds, Anal. Chem. 77
(2005) 303A–311A.
Acknowledgements [17] X.-S. Zhou, B.-W. Mao, C. Amatore, R.G. Compton, J.-L. Marignier, M. Mostafavi, J.-F.
Nierengarten, E. Maisonhaute, Transient electrochemistry: beyond simply temporal res-
olution, Chem. Commun. (Cambr. Engl.) 52 (2016) 251–263.
H. M. Amin gratefully acknowledges DFG for funding (No. AB 702/1-1). [18] Y. Uchida, E. Kätelhön, R.G. Compton, Linear sweep voltammetry with non-triangular
The authors thank Dr. Christopher Batchelor-McAuley for writing the soft- waveforms: new opportunities in electroanalysis, J. Electroanal. Chem. 818 (2018)
ware to run the semi-circular potential wave. 140–148.
[19] Y. Uchida, E. Kätelhön, R.G. Compton, Linear sweep voltammetry with non-triangular
waveforms at a microdisc electrode, J. Electroanal. Chem. 823 (2018) 465–473.
Appendix A. Supplementary data [20] H.M.A. Amin, Y. Uchida, E. Kätelhön, R.G. Compton, Semi-circular potential sweep volt-
ammetry: experimental verification and determination of the formal potential of a re-
versible redox couple, J. Electroanal. Chem. 836 (2019) 62–67.
Supplementary data to this article can be found online at https://doi. [21] Y. Uchida, E. Kätelhön, R.G. Compton, Sweep voltammetry with a semi-circular poten-
org/10.1016/j.jelechem.2020.114891. tial waveform: electrode kinetics, J. Electroanal. Chem. 835 (2019) 60–66.
[22] C. Batchelor-McAuley, J. Ellison, K. Tschulik, P.L. Hurst, R. Boldt, R.G. Compton, In situ
nanoparticle sizing with zeptomole sensitivity, Analyst 140 (2015) 5048–5054.
References [23] Y. Wang, J.G. Limon-Petersen, R.G. Compton, Measurement of the diffusion coefficients
of [Ru(NH3)6]3+ and [Ru(NH3)6]2+ in aqueous solution using microelectrode dou-
[1] C. Bourdillon, C. Demaille, J. Moiroux, J.-M. Saveant, Catalysis and mass transport in ble potential step chronoamperometry, J. Electroanal. Chem. 652 (2011) 13–17.
spatially ordered enzyme assemblies on electrodes, J. Am. Chem. Soc. (1995) 117. [24] J. Heinze, Diffusion processes at finite (micro) disk electrodes solved by digital simula-
[2] R.G. Compton, C.E. Banks, Understanding Voltammetry, 3rd ed. World Scientific, 2018. tion, J. Electroanal. Chem. Interfacial Electrochem. 124 (1981) 73–86.
[3] R.A. Marcus, N. Sutin, Electron transfers in chemistry and biology, Biochim. Biophys. [25] S. Eloul, R.G. Compton, Implementing high performance voltammetry simulation using
Acta (BBA) – Rev. Bioenerg. 811 (1985) 265–322. the implicit parallel algorithm, J. Electroanal. Chem. 771 (2016) 50–55.
[4] A.J. Bard, L.R. Faulkner, Electrochemical Methods: Fundamentals and Applications, 2nd [26] H.M.A. Amin, Y. Uchida, C. Batchelor-McAuley, E. Kätelhön, R.G. Compton, Non-
ed. Wiley, Hoboken, NJ, 2001. triangular potential sweep cyclic voltammetry of reversible electron transfer: experi-
[5] R.A. Marcus, On the theory of oxidation-reduction reactions involving electron transfer. ment meets theory, J. Electroanal. Chem. 815 (2018) 24–29.
I, J. Chem. Phys. 24 (1956) 966–978. [27] I. Lavagnini, R. Antiochia, F. Magno, An extended method for the practical evaluation of
[6] G.A. Mabbott, An introduction to cyclic voltammetry, J. Chem. Educ. 60 (1983) 697. the standard rate constant from cyclic voltammetric data, Electroanalysis 16 (2004)
[7] R.S. Nicholson, Theory and application of cyclic voltammetry for measurement of elec- 505–506.
trode reaction kinetics, Anal. Chem. 37 (1965) 1351–1355. [28] R. Jarošová, P.M. de Sousa Bezerra, C. Munson, G.M. Swain, Assessment of heteroge-
[8] M.V. Mirkin, A.J. Bard, Simple analysis of quasi-reversible steady-state voltammograms, neous electron-transfer rate constants for soluble redox analytes at tetrahedral amor-
Anal. Chem. 64 (1992) 2293–2302. phous carbon, boron-doped diamond, and glassy carbon electrodes, Phys. Status
[9] P. Sun, M.V. Mirkin, Kinetics of electron-transfer reactions at nanoelectrodes, Anal. Solidi A 213 (2016) 2087–2098.
Chem. 78 (2006) 6526–6534. [29] R. Naegeli, J. Redepenning, F.C. Anson, Influence of supporting electrolyte concentra-
[10] D.O. Wipf, E.W. Kristensen, M.R. Deakin, R.M. Wightman, Fast-scan cyclic voltammetry tion and composition on formal potentials and entropies of redox couples incorporated
as a method to measure rapid heterogeneous electron-transfer kinetics, Anal. Chem. 60 in Nafion coatings on electrodes, J. Phys. Chem. 90 (1986) 6227–6232.
(1988) 306–310.
[11] C. Beriet, D. Pletcher, A further microelectrode study of the influence of electrolyte con-
centration on the kinetics of redox couples, J. Electroanal. Chem. 375 (1994) 213–218.

You might also like