fast scan linear sweep voltammetry at a high-speed wall-tube electrode

9
Fast scan linear sweep voltammetry at a high-speed wall-tube electrode Neil V. Rees, Oleksiy V. Klymenko, Barry A. Coles, Richard G. Compton * Physical and Theoretical Chemistry Laboratory, Oxford University, South Parks Road, Oxford OX1 3QZ, UK Received 20 January 2003; received in revised form 5 May 2003; accepted 16 May 2003 Abstract The application of fast-scan linear sweep voltammetry methods to a high-speed wall-tube electrode (HWTE) is reported. Experiments are reported for the oxidation of N,N,N ?,N ?-tetramethyl-p -phenylenediamine (TMPD) in propylene carbonate solution containing 0.10 M tetrabutylammonium perchlorate for a 24 mm radius platinum microdisk electrode housed within the HWTE using a range of scan rates from 200 to 3000 V s 1 and average flow jet velocities from 0.24 to 9.4 m s 1 (corresponding to volume flow rates of 0.003 /0.12 cm 3 s 1 , and centre-line jet velocities from 0.5 to 18.9 m s 1 ). Linear sweep voltammograms (LSVs) are analysed for a simple electron transfer under high volume flow rates, by curve fitting. Analysis of the transient LSVs yielded values for k 0 , a , and E f 0 for TMPD of (5.99 /2.4) /10 2 cm s 1 , 0.469 /0.08 and 0.2179 /0.019 V (vs. Ag), respectively. This is in good agreement with independent experiments conducted using the high-speed channel electrode which yielded the results: k 0 / (6.39 /0.4) /10 2 cm s 1 , a / 0.529 /0.01, and E f 0 / 0.2349 /0.005 V (vs. Ag). The range of applicability of this method for measuring k 0 was also investigated and compared with existing channel electrode techniques. # 2003 Elsevier B.V. All rights reserved. Keywords: Fast scan linear sweep voltammetry; Wall-tube electrode; Heterogeneous rate constant; Oxidation; N,N,N ?,N ?-tetramethyl-p - phenylenediamine; Simulation 1. Introduction The impinging-jet electrode is a specific type of hydrodynamic electrode where a jet of electrolyte is propelled from a nozzle through a filled chamber (usually of the same electrolyte composition) to strike an electrode situated directly opposite the nozzle. The hydrodynamics of this arrangement under laminar conditions are believed to consist of two different flow regimes in the vicinity of the electrode. Directly below the jet, where the flow is axial, there is a stagnant region, which extends a distance of approximately 1 /3 nozzle diameters across the flat surface [1,2]. Outside of this stagnant region is the wall-jet region where the radial velocity begins to decay [3]. In the limiting case where the electrode is entirely within the stagnant region (i.e., the nozzle diameter is greater than that of the electrode), the electrode is uniformly accessible and termed a wall- tube electrode (WTE) [4]. Where the electrode is larger than the nozzle, and extends into the wall-jet region, it is termed a wall-jet electrode (WJE) [5]. Both limiting cases are illustrated in Fig. 1. The WTE and WJE have been the subject of much investigation in terms of their hydrodynamics and applications to mechanistic and analytical work [6]. The WTE has been further developed by Unwin and co- workers utilising microelectrodes to develop a microjet electrode [7]. This has been shown to be effective for investigating electron transfer kinetics [8], coupled solution reactions [9], and flow analysis [10]; and a model has recently been developed to approximate the mass transport within it [11 /13]. In this article we describe a miniaturised WTE for use in a pressurised system previously reported for channel electrode use [14]. This high-speed WTE (‘HWTE’) has been designed to operate at high volume flow rates achievable using the pressurised system, to measure fast * Corresponding author. Tel.: /44-1865-275-413; fax: /44-1865- 275-410. E-mail address: [email protected] (R.G. Compton). Journal of Electroanalytical Chemistry 557 (2003) 99 /107 www.elsevier.com/locate/jelechem 0022-0728/03/$ - see front matter # 2003 Elsevier B.V. All rights reserved. doi:10.1016/S0022-0728(03)00352-8

Upload: neil-v-rees

Post on 02-Jul-2016

221 views

Category:

Documents


1 download

TRANSCRIPT

Page 1: Fast scan linear sweep voltammetry at a high-speed wall-tube electrode

Fast scan linear sweep voltammetry at a high-speed wall-tubeelectrode

Neil V. Rees, Oleksiy V. Klymenko, Barry A. Coles, Richard G. Compton *

Physical and Theoretical Chemistry Laboratory, Oxford University, South Parks Road, Oxford OX1 3QZ, UK

Received 20 January 2003; received in revised form 5 May 2003; accepted 16 May 2003

Journal of Electroanalytical Chemistry 557 (2003) 99�/107

www.elsevier.com/locate/jelechem

Abstract

The application of fast-scan linear sweep voltammetry methods to a high-speed wall-tube electrode (HWTE) is reported.

Experiments are reported for the oxidation of N,N,N ?,N ?-tetramethyl-p -phenylenediamine (TMPD) in propylene carbonate

solution containing 0.10 M tetrabutylammonium perchlorate for a 24 mm radius platinum microdisk electrode housed within the

HWTE using a range of scan rates from 200 to 3000 V s�1 and average flow jet velocities from 0.24 to 9.4 m s�1 (corresponding to

volume flow rates of 0.003�/0.12 cm3 s�1, and centre-line jet velocities from 0.5 to 18.9 m s�1). Linear sweep voltammograms

(LSVs) are analysed for a simple electron transfer under high volume flow rates, by curve fitting. Analysis of the transient LSVs

yielded values for k0, a , and Ef0 for TMPD of (5.99/2.4)�/10�2 cm s�1, 0.469/0.08 and 0.2179/0.019 V (vs. Ag), respectively. This is

in good agreement with independent experiments conducted using the high-speed channel electrode which yielded the results: k0�/

(6.39/0.4)�/10�2 cm s�1, a�/ 0.529/0.01, and Ef0�/ 0.2349/0.005 V (vs. Ag). The range of applicability of this method for

measuring k0 was also investigated and compared with existing channel electrode techniques.

# 2003 Elsevier B.V. All rights reserved.

Keywords: Fast scan linear sweep voltammetry; Wall-tube electrode; Heterogeneous rate constant; Oxidation; N,N,N ?,N ?-tetramethyl-p -

phenylenediamine; Simulation

1. Introduction

The impinging-jet electrode is a specific type of

hydrodynamic electrode where a jet of electrolyte is

propelled from a nozzle through a filled chamber

(usually of the same electrolyte composition) to strike

an electrode situated directly opposite the nozzle. The

hydrodynamics of this arrangement under laminar

conditions are believed to consist of two different flow

regimes in the vicinity of the electrode. Directly below

the jet, where the flow is axial, there is a stagnant region,

which extends a distance of approximately 1�/3 nozzle

diameters across the flat surface [1,2]. Outside of this

stagnant region is the wall-jet region where the radial

velocity begins to decay [3]. In the limiting case where

the electrode is entirely within the stagnant region (i.e.,

the nozzle diameter is greater than that of the electrode),

the electrode is uniformly accessible and termed a wall-

tube electrode (WTE) [4]. Where the electrode is larger

than the nozzle, and extends into the wall-jet region, it is

termed a wall-jet electrode (WJE) [5]. Both limiting cases

are illustrated in Fig. 1.

The WTE and WJE have been the subject of much

investigation in terms of their hydrodynamics and

applications to mechanistic and analytical work [6].

The WTE has been further developed by Unwin and co-

workers utilising microelectrodes to develop a microjet

electrode [7]. This has been shown to be effective for

investigating electron transfer kinetics [8], coupled

solution reactions [9], and flow analysis [10]; and a

model has recently been developed to approximate the

mass transport within it [11�/13].

In this article we describe a miniaturised WTE for use

in a pressurised system previously reported for channel

electrode use [14]. This high-speed WTE (‘HWTE’) has

been designed to operate at high volume flow rates

achievable using the pressurised system, to measure fast

* Corresponding author. Tel.: �/44-1865-275-413; fax: �/44-1865-

275-410.

E-mail address: [email protected] (R.G.

Compton).

0022-0728/03/$ - see front matter # 2003 Elsevier B.V. All rights reserved.

doi:10.1016/S0022-0728(03)00352-8

Page 2: Fast scan linear sweep voltammetry at a high-speed wall-tube electrode

electrochemical kinetics (Fig. 2). In this article we report

the application of fast-scan linear sweep voltammetry to

investigate the heterogeneous electron transfer rate for

the oxidation of N,N,N ?,N ?-tetramethyl-p -phenylene-

diamine (TMPD) in propylene carbonate (PC), together

with high-speed channel electrode experiments [15] to

confirm the kinetic measurements. TMPD has pre-

viously been reported to have a standard heterogeneous

electron transfer rate constant, k0, of 1.5�/10�2 cm s�1

in PC at 293 K [16].To date, the use of WTEs to measure fast hetero-

geneous electron transfer rates has been confined to the

measurement of steady-state voltammograms [7], which

have been analysed using the quarter-potentials method

of Mirkin and Bard [17] assuming uniform accessibility.

As this method is unsuitable for analysis of transient

voltammograms, and has been reported to incur a

relatively large error in the value of the transfer

coefficient, a [7], we have investigated other possible

means of analysing the voltammetry obtained.

In the case of transient linear-sweep voltammograms,we consider the measurement of both peak currents and

peak potentials as a possible means of extracting values

of k0 and a . We also present a method of fitting the

entire voltammogram curve, thereby using all available

voltammetric information, and finding the values of

these kinetic parameters that provide a ‘best-fit’ to the

experimental data.

2. Theory

2.1. Wall-tube experiments

We consider a simple electron transfer reaction at theWTE surface (written as an oxidation):

A�e�?kf

kb

B (1)

where both A and B are kinetically stable on the

timescale of the experiment, and the bulk solution

contains only A before it enters the flow-cell. Theforward and reverse rate constants are given by But-

ler�/Volmer theory:

kf �k0e(1�a)F

RT(E�E0

f); kb�k0e

�aF

RT(E�E0

f)

(2)

where k0 (cm s�1) is the standard electrochemical rate

constant for heterogeneous electron transfer, Ef0 is the

formal potential of the A/B redox couple, a is the

transfer coefficient, and R , T , and F have their usual

significance.

The uniform accessibility property of the WTE (Fig.3) [4,18] allows us to solve the convective-diffusion

equations only in the z direction [18] where the solution

velocity is proportional to z2:

vz��Cz2 (3)

The constant C is given by [19],

Fig. 1. The two limiting forms of the impinging-jet electrode geometry.

Fig. 2. Schematic diagram showing the HWTE (not to scale). Fig. 3. Streamlines and geometry of a WTE.

N.V. Rees et al. / Journal of Electroanalytical Chemistry 557 (2003) 99�/107100

Page 3: Fast scan linear sweep voltammetry at a high-speed wall-tube electrode

C�0:52

�Vf

r3T

�3=2�rT

zT

�0:162

n�1=2 (4)

where Vf is the volume flow rate (cm3 s�1), n is the

kinematic viscosity (cm2 s�1), rT is the nozzle radius

(cm), and zT is the nozzle-electrode separation (cm).

The time-dependent convective-diffusion equationdescribing mass transport of species A to the electrode is

@a

@t�D

@2a

@z2�Cz2 @a

@z(5)

where a�/[A]/[A]bulk.

The boundary condition at the electrode correspond-

ing to Butler�/Volmer kinetics is

D@a

@z jz�0

� (kfa�kbb)z�0 (6)

where kf and kb are given by Eq. (2), and the

concentration tends to its bulk value with increasing

distance from the electrode:

ajz0��1 (7)

The mathematical model Eq. (5) to Eq. (7) was solved

using the finite element technique in Ref. [19].The simulation procedure developed in Ref. [19] for

linear sweep voltammetry was used to generate theore-

tical current�/voltage curves for subsequent comparison

with the experimental ones.

In order to ‘extract’ the values of k0, a and Ef0 from

the experimental voltammograms we used the least

squares method which consists of minimisation of the

following functional:

MSAD(k0; a; E0f )

�1

M

XM

k�1

(Iexp(Ek)�Ith(k0; a; E0f ; Ek))2 (8)

where Iexp(Ek) is the experimental current corresponding

to the potential Ek , Ith(k0, a , Ef0, Ek) is the theoretically

predicted value of the current at Ek , and M is the

number of experimental data points. The values of a , k0

and Ef0 were then found by minimisation of MSAD Eq.

(8) using the modified Newton’s method [15]. The main

iterative formula of the modified Newton’s method of

minimisation of a function f of a vector argument x is

x(k�1)� x(k)�gk[H (k)]�19f (k) (9)

where k is the iteration index, H(k ) is the Hessian matrix

(the matrix of second partial derivatives) of f evaluatedat the point x(k); 9f(k ) is the gradient of f at the same

point, and gk is an adjustable coefficient. All derivatives

were calculated using finite-difference approximations.

2.2. High-speed channel electrode experiments

The methodology used for the simultaneous fitting of

values of k0, a , and Ef0 from the experimental voltam-

mograms has been described in detail elsewhere [15].

3. Experimental

3.1. Reagents

Chemical reagents used were TMPD (Aldrich, 98%),

tetrabutylammonium perchlorate (TBAP) (Fluka Puriss

electrochemical grade, �/99%), and PC (Fluka, purum,

]/99%). These were of the highest grade available and

were used without further purification. Ultra-pure water

was used for cleaning (UHQ grade) and had a resistivity

of not less than 18 MV cm (Elga, High Wycombe,Bucks, UK). All solutions contained 0.10 M TBAP and

were thoroughly degassed with argon (Pureshield Ar-

gon, BOC Gases Ltd, UK). All experiments were

conducted at a temperature of 2929/1 K.

3.2. Instrumentation

The HWTE consists of a platinum microdisk elec-

trode embedded in a miniaturised wall-tube cell con-

nected within the pressurized apparatus described

previously [14]. Fig. 2 shows a schematic representation

of the electrode assembly. High volume flow rates areachieved by pressurizing the chamber containing the

solution and electrode assembly up to 1.5 atm. The

solution flows through the nozzle of the specially

designed wall-tube cell and out to the exit, which is at

ambient atmospheric pressure. The range of volume

flow rates available to the HWTE is determined by the

choice of chamber pressure and also by the solution

viscosity. For example in the present case of PC, volumeflow rates of up to 0.12 cm3 s�1 can be achieved

(corresponding to average flow linear jet speeds ap-

proaching 9.4 m s�1, with centre-line jet velocities of up

to 18.9 m s�1), whereas with acetonitrile the volume

flow rate can reach 0.25 cm3 s�1 (average flow linear jet

speeds of up to 19.8 m s�1, and centre-line jet velocities

of up to 39.5 m s�1).

In considering the solution flow, it is informative toconsider the Reynolds number, Re, given by Ref. [4]

Re�dU

n(10)

where d is the nozzle diameter, n is the kinematic

viscosity, and U is the mean solution velocity. In thepresent case with PC as solvent, the Reynolds number

has a maximum value of 570, which is within the regime

(Re B/2000) where flow has been shown to be laminar

N.V. Rees et al. / Journal of Electroanalytical Chemistry 557 (2003) 99�/107 101

Page 4: Fast scan linear sweep voltammetry at a high-speed wall-tube electrode

[4]. Use of less viscous solvents in the HWTE can easily

result in Reynolds numbers in excess of 2000: the details

of the hydrodynamics of these solvents in the HWTE

will be examined in a following publication.

The high-speed channel electrode (shown schemati-

cally in Fig. 4) and pressurized apparatus have been

described previously [14,15]. Extremely high flow rates

across a microband electrode are achieved by pressuriz-

ing the reaction chamber up to 1.5 atm. The solution

flows through the specially designed flow-cell (with

width, d�/ 0.2 cm and height, 2h�/108 mm) and out

through one of three capillaries of varying internal bore

size, out to the exit which is at ambient atmospheric

pressure. In this way, volume flow rates of 1.85 cm3 s�1

can be achieved with PC, corresponding to linear flow

speeds at the electrode approaching 14 m s�1. The

electrode employed in this work was a Pt microband

with length 12.5 mm and width 0.10 cm.

The ‘fast scan’ potentiostat used was built in-house

and has a scan rate of up to 3�/104 V s�1 and was used

with minimum filtering. This potentiostat is similar to

that described and utilised by Amatore et al. [20�/22]

and is capable of achieving ohmic drop compensation

by means of an internal positive feedback circuit. The

potential was applied with a TTi TG1304 programmable

function generator (Thurlby Thandar Instruments Ltd,

Huntingdon, Cambs, UK) and the current recorded

with a Tektronix TDS 3032 oscilloscope (300 MHz

band-pass, 2.5 GS s�1). Computer programs were

written in FORTRAN 77 and run on a Pentium IV 2

GHz PC, which was also used for data analysis.

3.3. Wall-tube cell and electrodes

The wall-tube cell was constructed from Kel-F (BSL

Ltd, Kidlington, Oxon, UK) and precision bore Pyrex†

tube (Jencons plc, Leighton Buzzard, Beds, UK), and is

shown schematically in Fig. 2. The body of the chamber

is formed from a Pyrex† tube of outer diameter 12.5

mm, wall thickness 1.5 mm, and length 22 mm. The

lead-in tube and nozzle (with nozzle diameter, rT�/127

mm) are formed from hand-drawn Pyrex† and sealed

into a Kel-F plug to form one end to the chamber as

shown in Fig. 2. The microdisk electrode was sealed intoanother Kel-F plug (with an exit hole bored into it) to

complete the cell. A silicone sealant (‘Aquaria Sealant’,

Dow Corning Ltd, Byfleet, Surrey, UK) was used to

ensure a seal between the glass and Kel-F. The Kel-F

plugs were machined to an airtight fit to the inner

diameter of the Pyrex† chamber body, such that the

separation of the nozzle outlet and electrode surface was

approximately 300 mm. Finally, the cell was sheathed ina silicone rubber tube to complete the seal on the cell.

The microdisk electrode was constructed by sealing

platinum wire (�/99.95%, Goodfellow Cambridge Ltd,

Huntingdon, Cambs., UK) into capillary soda glass as

described in the literature [23], and its radius (r1) was

determined to be 24 mm according to reported methods

[24]. The microband electrode was fabricated according

to a previously reported method [14,15].The working electrode was polished in between

experiments using alumina lapping compounds (BDH)

of decreasing size from 3 mm down to 0.25 mm on glass

and soft lapping pads (Kemet Ltd, UK). The polished

electrode was cleaned with ultra pure water and then

immersed in approximately 30% nitric acid and stored

under ultra pure water between each use.

The counter electrode was a smooth, bright platinumwire coil, and a silver wire (99.95%, Johnson Matthey

plc, London, UK) was used as a pseudo-reference

electrode.

4. Results

4.1. Discrimination limits of the curve fitting method

In order to determine the range of k0 which can be

measured experimentally using linear sweep voltamme-

try in the HWTE cell we performed a series of

simulations using the finite element program outlined

above.

We simulated a series of current�/voltage responses

with different values of k0, for fixed values of scan rate,

Vf, D , and v . The resulting voltammograms were thenplotted on the same graph for comparison. It was noted

that at high values of k0, the gradients of the rising part

of the voltammograms converged to a reversible limit.Fig. 4. Schematic diagram of a channel electrode.

N.V. Rees et al. / Journal of Electroanalytical Chemistry 557 (2003) 99�/107102

Page 5: Fast scan linear sweep voltammetry at a high-speed wall-tube electrode

Similarly, at lower values of k0, the gradients of the

rising part of the waves converged to an irreversible

limit. It is necessary to determine the lower limit of k0 in

particular since there is a danger that the fitting

procedure could yield an erroneous result. This is

because the waveshape does not change below a limiting

value of k0, but is shifted to higher potentials (for

oxidation) which will lead to overestimated values of Ef0.

Fig. 5(a) illustrates these two limiting cases. In order to

show the change in the gradients, the waves have been

shifted to coincide at the foot of the wave. Therefore the

abscissa has not been labelled with numerical values, as

these may be different for each voltammogram.

In order to determine the upper and lower measurable

limits of k0, we differentiated the voltammograms for

the range of k0. For each one the highest gradient was

used, corresponding to the point of inflexion of the

rising part of the wave (as shown in Fig. 5(b), where the

abscissa is as in Fig. 5(a)). Next, the difference between

the maximum gradients of reversible and irreversible

limits was measured and the experimental tolerance was

assumed to be 7.5% of this gradient range. Voltammo-

grams were simulated to obtain the minimum and

maximum measurable values of k0, having gradients

7.5% higher than the irreversible limit, and 7.5% less

than the reversible limit respectively. It was found that

the two limits were dependent on the scan rate, and so it

was necessary to find these limits for a range of scan

rates. To do this, we investigated the general current�/

voltage response to the normalised scan rate s and the

normalised heterogeneous rate constant Khet [19], given

by

s�1:546Fv

RTD�1=3 r3

T

Vf

�zT

rT

�0:108

n1=3 (11)

and

Khet�1:602k0D�2=3

�r3

T

Vf

�1=2�zT

rT

�0:054

n1=6 (12)

By considering the variation in voltammetry with

respect to these normalised parameters, a general result

can be obtained independent of the experimental

geometry, solvent viscosity and flow rate. Using theprocedure described above, the upper and lower limits

of Khet were found for a range of s and plotted in Fig. 6.

The lower curve was then fitted with a function, which

gave the following empirical relationship for it

f (s)�0:1559 ln[1�expf3:151(log10 s�0:3643)g]

�0:7721 (13)

For all values of s , it was found that the separation ofthe upper and lower curves was approximately 1.8,

which lead to the following bounds for Khet:

f (s)5 log10 Khet5 f (s)�1:8 (14)

This result is valid for any experimental parameters

which fit in the range of s investigated.

4.2. Steady-state voltammetry of TMPD at the high-

speed channel electrode

We first consider the use of the high-speed channel

electrode to obtain steady-state voltammograms in

Fig. 5. (a) Simulated voltammograms simulated for a range of k0

values, (b) Differential plot of (a), with a�/0.5, D�/10�6 cm2 s�1, v�/

3000 V s�1 and Vf�/0.01 cm3 s�1. Fig. 6. Upper and lower measurable limits of Khet as a function of s .

N.V. Rees et al. / Journal of Electroanalytical Chemistry 557 (2003) 99�/107 103

Page 6: Fast scan linear sweep voltammetry at a high-speed wall-tube electrode

order to measure the heterogeneous rate constant for the

first oxidation wave of TMPD in PC.

A solution containing 6.5 mM TMPD and 0.1 M

TBAP in PC was introduced into the pressure chamberof the apparatus and an initial pressure set. A linear

sweep voltammogram was then recorded at a scan rate

of 0.32 V s�1, which yielded a steady-state response [25]

enabling a limiting current to be measured for the first

oxidation wave. This was repeated at a range of

chamber pressures from 10 to 150 kPa to obtain a

range of volume flow rates (from 0.03 to 1.85 cm3 s�1).

Fig. 7 shows a plot of measured limiting currents againstVf

1/3 according to the Levich equation for channel

electrodes [26]:

ILim�0:925nFw[A]bulk(xeD)2=3(h2d)�1=3V1=3

f (15)

where n , F , D , and [A]bulk have their usual meaning, Vf

is the volume flow rate, and the other terms represent

the geometry of the channel cell and electrode as shown

in Fig. 4. The gradient of this plot can be used to obtain

a value for the diffusion coefficient of TMPD by

application of Eq. (15) as (3.19/0.1)�/10�6 cm2 s�1.Next, the current responses for each voltammogram

were normalized by division by the respective limiting

current and plotted against potential to produce a

‘normalized voltammogram’. These data were then

input into the computer program described previously

for the analysis of steady-state voltammetry [15], and

optimum values of k0, Ef0 and a were obtained

simultaneously. For convenience we display these dataas two separate three-dimensional contour plots. Fig. 8

shows the surface generated by the program as k0 and a

vary independently, and Fig. 9 shows the analogous

surface for the variation of k0 and Ef0. By repeating this

analysis for each data point, we arrive at a value for k0

of (6.39/0.4)�/10�2 cm s�1, a�/0.529/0.01 and Ef0�/

0.2349/0.005 V vs. Ag (quoted as mean9/S.D.). This

compares with a literature value of k0 which has beenreported to be 1.5�/10�2 cm s�1 at 293 K [16].

4.3. Steady-state voltammetry at the HWTE

We next consider the use of the HWTE to obtain

steady-state voltammograms in order to measure diffu-

sion coefficient of TMPD in PC.

A solution containing 6.5 mM TMPD and 0.1 M

TBAP in PC was introduced into the pressure chamber

of the apparatus and an initial pressure set. A linear

sweep voltammogram was then recorded at a scan rate

of 0.5 V s�1, which yielded a steady-state response [25]

enabling a limiting current to be measured for the first

oxidation wave. This was repeated at a range of

chamber pressures to obtain a range of volume flow

rates (from 0.003 to 0.12 cm3 s�1). Fig. 10 shows a plot

of measured limiting currents against Vf1/2 according to

the equation for wall-tube cells [4,18]:Fig. 7. Plot of Ilim vs. Vf

1/3 for 6.5 mM TMPD/0.10 M TBAP in PC

(R2�/0.999).

Fig. 8. Contour plot showing variation in MSAD with k0 and a for

TMPD measured at Vf�/1.75 cm3 s�1, showing a minimum at k0�/

0.063 cm s�1 and a�/0.51.

Fig. 9. Contour plot showing variation in MSAD with k0 and Ef0 for

TMPD measured at Vf�/1.75 cm3 s�1, showing a minimum at k0�/

0.063 cm s�1 and Ef0�/0.236 V (vs. Ag).

N.V. Rees et al. / Journal of Electroanalytical Chemistry 557 (2003) 99�/107104

Page 7: Fast scan linear sweep voltammetry at a high-speed wall-tube electrode

Ilim�0:6244nFp[A]bulkD2=3r21n

�1=6

�Vf

r3T

�1=2�rT

zT

�0:054

(16)

where n , F , D , and [A]bulk have their usual meaning, Vf

is the volume flow rate, n is the solvent kinematic

viscosity, and the other terms represent the geometry of

the channel cell and electrode as shown in Fig. 2. The

diffusion coefficient calculated from the gradient of Fig.

10 using Eq. (16) was found to be 3.1�/10�6 cm2 s�1, in

excellent agreement with that measured above with thechannel electrode.

4.4. Fast linear sweep voltammetry at the HWTE

A solution of 2.86 mM TMPD and 0.10 M TBAP in

PC was placed into the pressure chamber of the

apparatus and the pressure set and capillary selected in

order to achieve the required solution flow rate. A linear

sweep voltammogram was then measured at this flowrate at a scan rate of 200 V s�1 and a transient response

was obtained. Further linear sweep voltammograms

were then recorded over a range of scan rates from

200 to 7000 V s�1 whilst the solution was flowing at

volume flow rates ranging from 0.003 to 0.12 cm3 s�1.

Each recorded voltammogram was averaged a minimum

of 16 times, and negative potentials applied between

scans as a precaution to inhibit electrode passivation. Atno time was any experimental evidence of electrode

passivation or fouling observed.

Next, the voltammograms were amended by the

subtraction of capacitative currents [26], which was

achieved by subtracting the best-fit (least squares) line

through the non-faradaic signal, since the scan was of

sufficient width to enable the non-faradaic response to

become established before the faradaic signal com-menced. A combination of the use of a potentiostat

capable of on-line correction for ohmic drop, and the

low current generated by a microelectrode within the

HWTE made adjustment for ohmic distortion unneces-

sary [27].

The current vs. potential data for the voltammogram

was then input into the computer program describedabove, and optimum values of k0, a , and Ef

0 were

obtained simultaneously by the least squares method.

The simulations were made using parameters set accord-

ing to the experimental geometry, and the flow rate

used. The value used for the diffusion coefficient was

3.1�/10�6 cm2 s�1 as reported above. For convenience

we display this data as two separate contour plots. Fig.

11 shows the surface generated by the program as k0 andEf

0 vary independently, and clearly illustrates the sensi-

tivity of MSAD to the value of Ef0. Fig. 12 shows the

analogous surface for the variation of k0 and a . This

was repeated for over 15 separate voltammograms

corresponding to different scan rates and volume flow

rates. The results obtained were as follows: k0�/(5.99/

2.4)�/10�2 cm s�1, a�/0.469/0.08, and Ef0�/0.2179/

0.019 V (vs. Ag).In all cases these best-fit parameters were used to

simulate a theoretical voltammogram and this was

compared to the experimental result. Fig. 13 shows a

selection of such plots, showing the excellent agreement

observed in all cases.

5. Conclusion

The results obtained for k0 and a using fast-scan

linear sweep voltammetry in the HWTE have been

shown to be in agreement with those measured using

the high-speed channel electrode to record steady-state

voltammetry.

The HWTE has been shown to be an effective method

of measuring a heterogeneous rate constant for TMPD

in PC of k0�/(5.99/2.4)�/10�2 cm s�1. Simulationsindicate that the HWTE may be capable of measuring

Fig. 10. Plot of Ilim vs. Vf1/2 for 6.5 mM TMPD/0.10 M TBAP in PC

(R2�/0.988).

Fig. 11. Contour plot showing variation in MSAD with k0 and Ef0 for

TMPD measured at v�/1920 V s�1 and Vf�/0.065 cm3 s�1, showing a

minimum at k0�/0.051 cm s�1 and Ef0�/0.207 V (vs. Ag).

N.V. Rees et al. / Journal of Electroanalytical Chemistry 557 (2003) 99�/107 105

Page 8: Fast scan linear sweep voltammetry at a high-speed wall-tube electrode

electron transfer rates up to 5 cm s�1. In this regard, it

compares with the high-speed channel electrode,

although the latter has already been used to measure

rates in excess of 3 cm s�1 [29], and is predicted to be

able to reach rates of approximately 10 cm s�1 [15,28].

In terms of practical use, however, the channel

electrode is preferred. First, the hydrodynamics of

channel electrodes have been studied for some years

and are well understood, whereas a comprehensive

model has only recently been proposed for the WTE

and this indicated the existence of unexpected vortices

within improperly designed cell chambers [11�/13].

Second, the impact of the jet on the electrode surface

causes efficient outgassing of the solution forming many

tiny bubbles and possibly compromising the hydrody-

namics causing noise on the voltammograms. Finally,

the assembly and operation of the HWTE is not as

straightforward as with a channel electrode. Alignment

of the jet with the electrode is difficult, and the intra-

chamber pressures caused by the jet places high

demands on the chamber seals. The need for tight seals

also means that the microdisk electrodes cannot be

routinely replaced in the Kel-F blocks making up the

HWTE cell, as each Kel-F plug needs to be machined to

a precise fit with the electrode. This contrasts with the

channel electrode, where a seal is maintained by a

silicone gasket, which allows facile interchange of

electrodes with the flowcell [14]. The channel electrode

is therefore recommended over the HWTE for fast

kinetic measurements for ease of use.

Last we note that the work reported in this article has

utilised the viscous (2.53 cp at 298 K) solvent PC; less

viscous solvents, notably water and especially acetoni-

trile may lead to Reynolds numbers sufficiently large

that the laminar/turbulent threshold may be approached

or exceeded. We return to this, and related hydrody-

namic issues, in a subsequent article.

Fig. 12. Contour plot showing variation in MSAD with k0 and a for

TMPD measured at v�/1920 V s�1 and Vf�/0.065 cm3 s�1, showing a

minimum at k0�/0.051 cm s�1 and a�/0.41.

Fig. 13. Comparison of experimental (*/) and theoretical (k) current�/voltage curves, based on the best-fit parameters obtained, for a selection of

typical voltammograms recorded for a solution of 2.86 mM TMPD/0.10 M TBAP in PC. v�/240 V s�1, Vf�/0.003 cm3 s�1, v�/480 V s�1, Vf�/

0.012 cm3 s�1; v�/1920 V s�1, Vf�/0.065 cm3 s�1, and v�/2880 V s�1, Vf�/0.030 cm3 s�1.

N.V. Rees et al. / Journal of Electroanalytical Chemistry 557 (2003) 99�/107106

Page 9: Fast scan linear sweep voltammetry at a high-speed wall-tube electrode

Acknowledgements

We appreciate the generosity of Professor C. Amatore

in making available to us the designs for the fast-scanpotentiostat apparatus developed in their laboratory.

We thank EPSRC for a studentship for NVR and the

Clarendon Fund of Oxford University for partial

funding for OVK.

References

[1] I. Tani, Y. Komatsu, Proceedings of the 11th International

Congress of Applied Mechanics, H. Gortler (Ed.), Springer-

Verlag, 1966.

[2] T. Strand, AIAA Paper No. 64�/424.

[3] M.B. Glauert, J. Fluid Mech. 1 (1956) 625.

[4] D.-T. Chin, C.-H. Tsang, J. Electrochem. Soc. 125 (1978) 1461.

[5] J. Yamada, H. Matsuda, J. Electroanal. Chem. 44 (1973) 189.

[6] J.V. Macpherson, N. Simjee, P.R. Unwin, Electrochim. Acta 47

(2000) 29.

[7] J.V. Macpherson, S. Marcar, P.R. Unwin, Anal. Chem. 66 (1994)

2175.

[8] J.V. Macpherson, M.A. Beeston, P.R. Unwin, J. Chem. Soc.

Faraday Trans. I 91 (1995) 899.

[9] R.D. Martin, P.R. Unwin, J. Electroanal. Chem. 397 (1995) 325.

[10] J.V. Macpherson, P.R. Unwin, Anal. Chem. 71 (1999) 4642.

[11] J. Melville, N. Simjee, P.R. Unwin, B.A. Coles, R.G. Compton, J.

Phys. Chem. B 106 (2002) 2690.

[12] J. Melville, N. Simjee, P.R. Unwin, B.A. Coles, R.G. Compton, J.

Phys. Chem. B 106 (2002) 10424.

[13] J.L. Melville, B.A. Coles, R.G. Compton, N. Simjee, J.V.

Macpherson, P.R. Unwin, J. Phys. Chem. B 107 (2003) 379.

[14] N.V. Rees, R.A.W. Dryfe, J.A. Cooper, B.A. Coles, R.G.

Compton, S.G. Davies, T.D. McCarthy, J. Phys. Chem. 99

(1995) 7096.

[15] N.V. Rees, O.V. Klymenko, E. Maisonhaute, B.A. Coles, R.G.

Compton, J. Electroanal. Chem. 542 (2003) 23.

[16] H. Fernandez, M.A. Zon, J. Electroanal. Chem. 332 (1992) 237.

[17] M.V. Mirkin, A.J. Bard, Anal. Chem. 64 (1992) 2293.

[18] W.J. Albery, S. Bruckenstein, J. Electroanal. Chem. 144 (1983)

105.

[19] O.V. Klymenko, D.J. Gavaghan, K.E. Harriman, R.G. Compton,

J. Electroanal. Chem. 531 (2002) 25.

[20] C. Amatore, C. Lefrou, F. Pfluger, J. Electroanal. Chem. 270

(1989) 43.

[21] C. Amatore, E. Maisonhaute, G. Simmonneau, Electrochem.

Commun. 2 (2000) 81.

[22] C. Amatore, E. Maisonhaute, G. Simmonneau, J. Electroanal.

Chem. 486 (2000) 141.

[23] R.M. Wightman, D.O. Wipf, in: A.J. Bard (Ed.), Electroanaly-

tical Chemistry, vol. 15, Marcel Dekker, New York, 1989, p. 267.

[24] B.A. Brookes, N.S. Lawrence, R.G. Compton, J. Phys. Chem. B

104 (2000) 11258.

[25] C.M.A. Brett, A.M. Brett, Electrochemistry. Principles, Methods

and Applications, Oxford University Press, Oxford, UK, 1993, p.

115.

[26] V.G. Levich, Physicochemical Hydrodynamics, Prentice-Hall,

Englewood Cliffs, NJ, 1962.

[27] A.J. Bard, L.R. Faulkner, Electrochemical Methods: Fundamen-

tals and Applications, second ed., Wiley, New York, 2001, p. 234.

[28] N.V. Rees, J.A. Alden, R.A.W. Dryfe, B.A. Coles, R.G.

Compton, J. Phys. Chem. 99 (1995) 14813.

[29] N.V. Rees, O.V. Klymenko, B.A. Coles, R.G. Compton, J.

Electroanal. Chem. 534 (2002) 151.

N.V. Rees et al. / Journal of Electroanalytical Chemistry 557 (2003) 99�/107 107