Chapter 8

 

Temperature modulated open circuit potential spectroscopy

 

8.1. Introduction

 

Kinetic studies of fast chemical reactions, e.g. the neutralisation in aqueous solution of strong acid by strong base that has a half-life of 10-6 s-1 or less, under ordinary conditions are difficult to characterise using conventional kinetic techniques. For fast processes it is difficult to deconvolute the chemical and mixing kinetics. These technical problems have been overcome by the development of a group of methods known as relaxation methods. The pioneer worker in this field was Manfred Eigen 1. Relaxation methods differ fundamentally from the conventional kinetic methods in that the system is initially at equilibrium under a given set of conditions. These conditions are suddenly changed; the system is no longer at equilibrium, and it relaxes to a new state of equilibrium. The speed with which it relaxes can be measured and the rate constants can be obtained.

 

There are various ways in which the conditions are disturbed. One possibility is changing the hydrostatic pressure. Another approach is to increase the temperature suddenly, usually by the rapid discharge of a capacitor. This method is called the temperature jump or T-jump method. It is possible to raise the temperature of a tiny cell containing a reaction mixture by a few degrees in less than 10-7 s, sufficiently rapid to permit the study of fast chemical processes 1-3.

 

A few groups studying electrochemical process, e.g., mass transport, charge transfer reactions and adsorption processes, have employed the temperature jump method at electrodes. Gründler et al. 4-29 used dc electrical heating to perform electrochemical experiments. Many assumptions were taken in these studies the main one being that the heat is only perturbing the temperature of the electrode, the rest of the system remaining constant. In the experimental conditions, the heating up of the solution created a gradient of temperature from the warm areas to the cold areas of the electrolyte. The gradient of temperature originated a gradient of concentration close to the source of heating making non-homogeneous conditions in the solution. Other significant effects to be considered were the convective flow that was originated when in stationary conditions two parts of a liquid were at different temperatures. Heating system generated noise in the electrochemical measurements of the electrode. Results were hardly dependent on convection and gradients of temperature and concentration. Pure temperature effects cannot be deconvoluted easily from the results suggesting that the temperature jump of the electrode was not very precise. Valdes and Miller 30-32 described a quantitative analysis of thermal modulation voltammetry, modifying the experimental methodology using a laser as heating source. A laser beam heated the absorptive back of a think platinum disk mounted in a glass cell. The beam was chopped between 5 Hz and 10 Hz. The periodic change in the temperature produced a current modulation that was recorded during a controlled potential scan. From the rotation speed of the disk and the modulated current data it was possible to obtain the temperature coefficients necessary to describe the changes in temperature. The Valdes and Miller method that quantifies the temperature perturbation in terms of engineering parameters (flux of solution, Reynolds’s number, etc). Valdes and Miller in their experimental set-up heated the electrode but at the same time the electrochemical solution. They did not consider the Soret effect (the change in the distribution of ions due to the temperature induced formation of a free energy gradient) which is a critical parameter when the solution is heated. The heat lost by the backing medium (in this case the spindle used to rotate the platinum disk) was not considered. These facts suggest the results obtained are not very precise. 

 

Olivier et al. 33 employed a temperature jump method to perform electrochemical impedance measurements. Following the Valdes idea, an infrared diode was used as heating source.  This infrared diode was modulated in a sine wave. The infrared beam irradiated the working electrode through a prismatic glass cell. From the electrochemical impedance measurements the thermoelectrochemical transfer functions were obtained to describe the system.

 

Gabrielli et al. 34, 35 investigated electrochemical impedance measurements using the temperature jump method for heterogeneous reaction mechanisms. Gabrielli suggested that many processes were controlled by the transport of the charged particles. The studies considered the Soret effect, assuming low temperature gradients. It was argued that in dilute solutions the low temperature gradient result in a very low concentration gradient and hence a negligible Soret effect. Under these conditions the temperature distribution could be assumed to be controlled by the thermal diffusion equation. The electrochemical system could be described by the thermal diffusion equation. This method assumed a constant temperature gradient during the diffusion process that is not very realistic approximation. Another misconception of the theory was that the temperature coefficient of the reference electrode potential, which is important in the kinetic region, was under estimated. Again the background medium of the electrode was neglected.

 

Rotenburg 36, 37 overcame Gabrielli’s problem with respect to the reference electrode and developed a new method to perform thermoelectrochemical impedance. Again in these studies a laser was chosen as the source of heat. The electrode potential was controlled and swept by a potentiostat and the ac response recorded by a frequency response analyser. Considering the phase angle and the ac current responses, the method describes in detail the system. A problem of this methodology was that the electrode was illuminated through the solution.

 

The aim of this chapter is to develop temperature modulated open circuit potential spectroscopy. This new technique overcomes the problems of the solution heating. A diamond heat sink is introduced as electrode backing medium to adsorb all the heat from the electrode and not allow the heating of the solution, hence reducing the Soret effect to negligible values. The backing medium is now considered as a relevant element in the method. The sine wave modulation of the heating source (infrared diode) allows better control and more precise analysis of the electrochemical system. The backing medium (diamond) is transparent to infrared illumination avoiding the need to pass the beam through the solution. It is demonstrated how this method allows impedance studies to be performed over a wide range of frequencies because there is no influence of the electrical elements in this new technique. Normally impedance depends on the bandwidth of the potentiostat and frequency response analysers used in the experiment.

 

8.2. Theory

 

The electrochemical system employed in these studies has not been described before and requires a new theoretical development to establish how the changes in the temperature of the electrode modify the total potential of the cell. This section is divided in two to simplify the theoretical approach. Firstly it is considered how the light intensity influences the change in the temperature of electrode; secondly the change in temperature is related to the change in the total potential of the electrochemical cell.

 

8.2.1. Relationship between light intensity and the electrode temperature

 

The diagram below shows the system to be considered:

Figure 8.1. Schematic diagram of the system diamond-gold / electrolyte

 

It is assumed that the gold layer is infinitely thin, i.e. it can be assumed that temperature (T) is uniform on the electrode surface.

 

The Fick’s laws of heat transport are:   

 

                                                       J is heat flux     [equation 8.1]

 

                                                                            [equation 8.2]

If at the gold surface (x=0) it is assumed that all the incident light is converted to heat:

 

                                                                       

                                                                                                             [equation 8.3]

where I(t) is the photon flux at time t and e is the energy of a single photon.

 

To continue it is necessary to make a further assumption corcerning the system. It is assumed that diamond is heated infinitely quickly on the experimental time scale. Hence the gold/diamond system is of uniform temperature.

 

Hence

   Jx=0 = 0    Þ                                               [equation 8.4]

 

Equation 8.1 may be solved for the semi-infinite boundary conditions:

 

At t=0                                        T= Ta                                             for all x            

 [equation 8.5]

At all t                                       T= Ta                                             for x®¥       

 [equation 8.6]

 

For positive x , the Laplace transform of equation 8.2 with respect to t is

 

                                            [equation 8.7]

 

Applying the boundary condition from equation 8.5 yields:

                                                   [equation 8.8]

 

 

which rearranges to:

            [equation 8.9]

 

Equation 8.9 is of the form:

 

                    [equation 8.10]

for which the Laplace transform is:

 

                                        [equation 8.11]

or

 

                      [equation 8.12]

 

Setting s=0 indicates:

                        [equation 8.13]

 and therefore

                  [equation 8.14]

 

The inverse of equation 8.14 is:

 

                                                       [equation 8.15]

 

Hence equation 8.9 may be written

 

                                               [equation 8.16]

 

 

The Laplace transform of the boundary condition from equation 8.6 is:

 

                                                   at                        x ®¥

           [equation 8.17]

 

So

B=0                                        [equation 8.18]

 

 

 

And hence

       [equation 8.19]

 

Differentiating equation 8.19 with respect to x yields:

 

           [equation 8.20] 

 

which at x=0 is

            [equation 8.21]

 

Laplace transform of the boundary conditions from the equation 8.4 gives:

 

                        [equation 8.22]

 

Comparing equations 8.21 and 8.22 results in

 

                                                            [equation 8.23]

 

which when substituted in the equation 8.19 at x=0 gives

 

                    [equation 8.24]

 

Laplace inversion using convolution theory results in:

 

              [equation 8.25]

For a system subject to a modulation

 

                          [equation 8.26]

 

The steady-state AC response is:

 

        [equation 8.27]

 

If  g = U2             Þ                   dg = 2UdU                                       

                                                          ß

 

                                              [equation 8.28]

 

 

 

Remembering that

 

                             [equation 8.29]

gives

        [equation 8.30]

 

Reordering, this final expression yields:

 

                        [equation 8.31]

 

i.e. the modulated temperature lags the modulated intensity by 45o.

 

 

8.2.2. Relationship between the temperature and the open circuit potential

 

Considering the equivalent circuit shown below:

Figure 8.2. Equivalent circuit diagram

 

DVj is defined as the  junction potential between the (hot) electrode and the (cold) contact wire

 

                                                     [equation 8.32]

 

For small T changes bj is constant. Therefore:

 

                                       [equation 8.33]

 

DVs is the change in potential due to the Soret effect

 

                                                       [equation 8.34]

 

For small T changes bs is constant. Therefore:

 

                                     [equation 8.35]

 

The Soret effect looks after the change in the distribution of ions in solution due to the temperature induced formation of a free energy gradient. The Soret potential is analogous to a  junction potential. It should be  noted that the ions responsible for the Soret effect are those of the background electrolyte and not the electroactive species. J.F. Smalley et al. 38 indicate that the Soret potential is established in nanoseconds, i.e. DVs follows any temperature change. 

 

DVf  is the change in potential of the redox couple due to the change of temperature

 

                                                                [equation 8.36]

 

For small T changes bf is constant. Therefore:

 

                                                              [equation 8.37]

 

 

Now considering that

                                       [equation 8.38]

 

where C’ is the integrated capacitance between Vi and the open circuit potential or potential of zero charge (pzc), for a small changes in T:

 

                   [equation 8.39]

 

           [equation 8.40]

 

Differentiating equation 8.38 with respect to Vi gives:

 

                               [equation 8.41]

 

as                                                where C is the differential capacitance

 

this can be written as

                                 [equation 8.42]

Reordering equation 8.42 yields

 

                                                             [equation 8.43]

 

and substituting in the denominator of equation 8.40 gives

 

 

                [equation 8.44]

Substituting the equation 8.38 into equation 8.44

 

                    [equation 8.45]

 

If the change of temperature is small:

 

                              [equation 8.46]

     

where                                   [equation 8.47]

 

The measured potential change is:

 

                               [equation 8.48]

 

Substituting equations 8.33, 8.35 and 8.46 into the equation 8.48 gives:

 

                                 [equation 8.49]

and if on the time scale of the experiment bi is independent of time differenciating results in:

 

                [equation 8.50]

 

where  jDL is the current across the capacitor. As the net current must be zero under open circuit potential condition, Kirchoff’s law indicates jDL = -jF , where jF is the Faradaic current flowing. If the Faradaic impedance, Z, is independent of T (electron transfer at OHP) then the voltage drop across the Faradaic component is DVF = jF Z + DVf. As the Faradaic component is parallel to the double layer     DVF = DVi  then  DVi = - jDL Z + DVf.

 

Hence:

              [equation 8.51]

 

Substituting for Vi using equation 8.48:

 

                [equation 8.52]

 

and rearranging equation 8.52 yields

 

     [equation 8.53]

 

Equation 8.53 shows the relationship between DVT and the temperature.

 

 

8.3. Experimental set-up

 

Electrochemical experiments were performed using a two electrode system. This section descibes the apparatus used.

 

8.3.1. Electrolyte solutions

 

18.3 MW cm ultrapure deionised water (Millipore) was used as solvent and 0.1 mol dm-3 sulfuric acid (H2SO4) as supporting electrolyte.

 

Different solutions were prepared contaning:

 

·        1´10-3 mol dm-3 of potasium ferricyanide (K3Fe(CN)6) / 1´10-3 mol dm-3 of potasium ferrocyanide (K4Fe(CN)6)  in 0.5 mol dm-3 H2SO4 in H2O.

 

·        0.1´10-3 mol dm-3 of potasium ferricyanide (K3Fe(CN)6) / 0.1´10-3 mol dm-3 of potasium ferrocyanide (K4Fe(CN)6)  in 0.5 mol dm-3 H2SO4 in H2O.

 

·        1´10-3 mol dm-3 of ferrosulfate (Fe2SO4) / 1´10-3 mol dm-3 of ferricsulfate (Fe2(SO4)3)  in 0.5 mol dm-3 H2SO4 in H2O.

 

·        0.1´10-3 mol dm-3 ferrosulfate (Fe2SO4) / 0.1´10-3 mol dm-3 of ferricsulfate (Fe2(SO4)3) in 0.5 mol dm-3 H2SO4 in H2O.

 

Glassware was cleaned with a five step process (for further details see section 5.3.1)

 

All the solutions were purged with standard laboratoy grade oxygen-free argon (Ar) for aproximmately 30 minutes before any electrochemical experiment was performed.

 

8.3.2. Characteristics of the substrates

 

The substrate used for these studies was quartz. Samples were already cut and supplied in pieces of 1 cm2 (1 cm ´ 1 cm) with 1 mm thickness. Normal pre-treatment procedure was applied before the growth stage (see chapter 2 for further details). Quartz was used as a substrate because it allows the growth of diamond in such conditions that the final samples were transparent a requirement essential for the electrodes in these studies.

 

8.3.3. Growing characteristics of the samples

 

Diamond deposition was performed for seven hours, giving a film thickness of approximately 5 mm. Films were cooled in the chamber under a hydrogen atmosphere. Films were transparent. The quality of the diamond was assessed using Raman spectroscopy and scanning electron microscopy, sem (check chapter 2 for further details).

 

8.3.4. Electrode construction and cell assembly

 

Approximately 40 nm of chromium were deposited in a 0.4 cm diameter circle on top of a 1 cm ´ 1 cm diamond sample. This was followed by approximately      300 nm of gold (see figure 8.3.a). The sample was then attached to a glass tube using the following method. A piece of a single wire (approximately six centimetres long) was attached to the edge of the sample with silver conducting paint (see figure 8.3.b). After drying the wire was coated with araldite to protect the electrical contact (see figure 8.3.c). A glass tube whose end was of the same diameter (external one, see figure 8.4) as the gold area was sealed to the diamond sample using silicon rubber sealant (see figure 8.3.d and 8.3.e). The electrode was left to dry for 24 hours.

 

The cell was completed with a counter electrode. It was secured through a plastic moulding that rested on top of the glass tubing. In figure 8.3 a schematic diagram of the contact is shown.

Figure 8.3. Schematic diagram of the different stages in the construction of the electrical contact and the cell assembly (note that the diagram is not to scale). a) After gold deposition; b) attachment of the single wire; c) protection of the contact with araldite resin d) placement of the glass tube; e) sealed of the glass tube to the diamond sample using silicon rubber sealant.

 

Figure 8.4. Detailed diagram of the glass cell used in these studies

 

8.3.5. Counter Electrodes

 

The counter electrodes have been described in section 5.3.3.

 

8.3.6. Reference electrode

 

A platinum wire was used as a pseudo reference electrode in these studies when a three-electrode mode configuration was required.

 

8.3.7.   Faraday Cages

 

Faraday Cages have been described in section 7.2.6.

 

8.3.8. Electronic equipment used in these studies

 

A light beam from 5mW infra red diode source was focused on the back of the gold diamond electrode. The light beam was modulated at variable frequencies using a lock in amplifier (Standford Research Systems Model SR830 DSP). The lock in amplifier fed an ac (0.75 V) and dc (4.25V) sinusoidal component to the infrared diode. The fraction modulated in the light was 10%. After the beam irradiated the gold diamond electrode the electrochemical response in ac and dc terms (in voltage) was analysed by the lock in amplifier. For each frequency used in the experiment, phase angle and voltage responses at open circuit potential conditions were recorded. The electrochemical glass cell operated in two electrode mode (gold diamond electrode as working electrode and a platinum counter electrode as reference electrode/counter electrode) (see figure 8.3). The cell assembly was housed in a Faraday cage to avoid any stray light or electrical pickup from external sources.

 

Cyclic voltammograms measurements were performing using EG&G Princeton Applied Research Potentiostat/Galvanostat Model 273; controlled using Research Electrochemistry software (version 4.3).

 

Impedance measurements were performed using a Solartron 1286 potentiostat and a Solartron 1250 Frequency Response Analyser (FRA). The instruments were controlled using Z-plot software. In all ac experiments modulation amplitude of 10 mV was applied.

Figure 8.5. Experimental temocps display

 

8.4. Cyclic voltammograms

 

The cell was operated in three-electrode mode configuration to record the cyclic voltammograms (adding a platinum wire electrode to the usual two-electrode mode described in other sections).

 

An i-E curve for a gold (on diamond substrate) electrode immersed in the   1´10-3 mol dm-3 of potassium ferrocyanide (K4Fe(CN)6)/ 1´10-3 mol dm-3 of ferricyanide (K3Fe(CN)6) in 0.5 mol dm-3 H2SO4 aqueous solution is shown in the figure 8.6.

Figure 8.6. A cyclic voltammogram recorded at gold (on diamond substrate) electrode immersed in 1´10-3 mol dm-3 of potassium ferricyanide (K3Fe(CN)6)/ 1´10-3 mol dm-3 of potassium ferrocyanide (K4Fe(CN)6) in 0.5 mol dm-3 H2SO4 aqueous solution. The i-E curve was recorded at scan rate of 0.1 V s-1.

 

The curve shows a reversible reduction peak at -0.12 V and a reversible oxidation peak at 0.09 V that correspond with the formation of the species Fe(CN)64- / Fe(CN)63- respectively. The behaviour observed is the typical for the reduction of K3Fe(CN)6/K4Fe(CN)6 in aqueous solvent. The magnitude in the differences between peaks (0.21 V) can be explained by the fact that the cyclic voltammogram was recorded in absence of iR compensation.

 

 An i-E curve for a gold (on diamond substrate) electrode immersed in the     0.1´10-3 mol dm-3 of ferricsulfate (Fe2(SO4)3)/ 0.1´10-3 mol dm-3 of ferrosulfate (FeSO4 ) in 0.5 mol dm-3 H2SO4 aqueous solution is shown in the figure 8.7.

Figure 8.7. A cyclic voltammogram recorded at gold (on diamond substrate) electrode immersed in 0.1´10-3 mol dm-3 of ferricsulfate (Fe2(SO4 )3)/ 0.1´10-3 mol dm-3 of ferrosulfate (FeSO4 ) in 0.5 mol dm-3 H2SO4 aqueous solution. The i-E curve was recorded at scan rate of 0.1 V s-1.

 

The curve shows a reversible reduction peak at -0.09 V and a reversible oxidation peak at 0.06 V that correspond with the formation of the species Fe+2/ Fe+3 respectively. The behaviour observed is the typical for the reduction of FeSO4/Fe2(SO4)3 in aqueous solvent. The magnitude in the differences between peaks (0.15 V) can be explained by the fact that the cyclic voltammogram was recorded in absence of iR compensation.

 

An i-E curve for a gold (on diamond substrate) electrode immersed in the 1´10-3 mol dm-3 of ferricsulfate (Fe2(SO4 )3)/ 1´10-3 mol dm-3 of ferrosulfate (FeSO4 ) in 0.5 mol dm-3 H2SO4 aqueous solution is shown in the figure 8.8.

Figure 8.8. A cyclic voltammogram recorded at a gold (on diamond substrate) electrode  immersed in 1´10-3 mol dm-3 of ferricsulfate (Fe2(SO4 )3)/ 1´10-3 mol dm-3 of ferrosulfate (FeSO4 ) in 0.5 mol dm-3 H2SO4 aqueous solution. The i-E curve was recorded at scan rate of 0.1 V s-1.

 

The curve shows a reversible reduction peak at -0.10 V and a reversible oxidation peak at 0.07 V that correspond with the formation of the species Fe+2/ Fe+3 respectively. The behaviour observed is the typical for the reduction of FeSO4/Fe2(SO4)3 in aqueous solvent. The magnitude in the differences between peaks (0.17 V) can be explained by the fact that the cyclic voltammogram was recorded in absence of iR compensation.

 

These cyclic voltammograms were recorded before any ac experiments or temopcs experiments to ensure that the electrodes were functioning correctly.

 

 

8.4. Ac impedance experiments

 

All ac impedance experiments were recorded using a 10 mV ac modulation. The frequency range was between 65 KHz and 0.01 Hz using a scale of 7 steps per decade. The cell was operated in three-electrode mode.

 

Figures 8.9, 8.10 and 8.11 describe the ac impedance behaviour of                   1´10-3 mol dm-3 of potasium ferricyanide (K3Fe(CN)6) / 1´10-3 mol dm-3 of potasium ferrocyanide (K4Fe(CN)6)  in 0.5 mol dm-3 H2SO4 in H2O.

Figure 8.9. Imaginary component (Z²) against real part (Z¢) of the impedance for 1´10-3 mol dm-3 of potasium ferricyanide (K3Fe(CN)6) / 1´10-3 mol dm-3 of potasium ferrocyanide (K4Fe(CN)6)  in 0.5 mol dm-3 H2SO4 in H2O. The ac modulation was 10 mV. Frequency range was between 65 KHz and 0.01 Hz. In the inset plot it can be appreciated better the shape of the semicircle at high frequencies.

 

A semicircle for high frequencies is observed. In medium and low frequencies the basic behaviour of the impedance is linear. This plot is in agreement with the Randles equivalent circuit. The semicircle region at high frequencies corresponds with the frequency region where all the non-charging current may be reported by an ohmic resistance. This behaviour is described by the equation:

 

 

                                                                                                          

           [equation 8.54]

 

where Z¢ an Z² are the real and the imaginary component of the impedance respectively; RW is the ohmic solution resistance; Rct is the charge transfer resistance at the electrode. The equation 8.54 allows by extrapolation of the semicircle in the plot the calculation of the value of the Rct and RW. The maximum in the semicircle occurs at                   , hence full analysis of the semicircle allows the interfacial capacitance to be determined.

At low frequencies there is a diffusion controlled electrode process. In this process the main characteristic is the linear correlation of Z¢ and Z². This linear behaviour is described by the following equation:

                                                               

                                                                                                           [equation 8.55]

 

where Cd is the differential capacitance of the double layer; s is an experimental parameter that can be calculated from the following equation:

 

 

                                                                                                           [equation 8.56]

 

where T is the temperature; R the gas constant; n is the number of electrons exchanged in the electrochemical process; A is the exposed area of the electrode; DO and DR are the diffusion coefficients for the oxidise and reduced form of the redox couple respectively; CO* and CR* are the concentrations of the oxidise and reduced form of the redox couple at the equilibrium stage.

Figure 8.10. Phase component (DEG) against frequency  (Hz) for 1´10-3 mol dm-3 of potasium ferricyanide (K3Fe(CN)6) / 1´10-3 mol dm-3 of potasium ferrocyanide (K4Fe(CN)6)  in 0.5 mol dm-3 H2SO4 in H2O. The ac modulation was 10 mV. Frequency range was between 65 KHz and        0.01 Hz. The frequency is in a logaritmic scale.

 

There is a change in phase in the region between 100 Hz and 1000 Hz. This region relates the exchange of an electron in the redox couple.

Figure 8.11. Magnitude against frequency (Hz) for 1´10-3 mol dm-3 of potasium ferricyanide (K3Fe(CN)6) / 1´10-3 mol dm-3 of potasium ferrocyanide (K4Fe(CN)6)  in 0.5 mol dm-3 H2SO4 in H2O. The ac modulation was 10 mV. Frequency range was between 65 KHz and 0.01 Hz. The frequency and the magnitude axes are in a logaritmic scale.

 

Figures 8.12, 8.13 and 8.14 describes the ac impedance behaviour of 0.1´10-3 mol dm-3 of potasium ferricyanide (K3Fe(CN)6) / 0.1´10-3 mol dm-3 of potasium ferrocyanide (K4Fe(CN)6)  in 0.5 mol dm-3 H2SO4 in H2O.

Figure 8.12. Imaginary component (Z²) against real part (Z¢) of the impedance for 0.1´10-3 mol dm-3 of potasium ferricyanide (K3Fe(CN)6) / 0.1´10-3 mol dm-3 of potasium ferrocyanide (K4Fe(CN)6)  in 0.5 mol dm-3 H2SO4 in H2O. The ac modulation was 10 mV. Frequency range was between 65 KHz and 0.01 Hz. In the inset can be appreciated better the shape of the semicircule at high frequencies.

 

A very similar behaviour is observed in figure 8.12 as figure 8.9.

Figure 8.13. Phase component (DEG) against frequency  (Hz) for 0.1´10-3 mol dm-3 of potasium ferricyanide (K3Fe(CN)6) / 0.1´10-3 mol dm-3 of potasium ferrocyanide (K4Fe(CN)6)  in 0.5 mol dm-3 H2SO4 in H2O. The ac modulation was 10 mV. Frequency range was between 65 KHz and 0.01 Hz. The frequency is in a logaritmic scale.

Figure 8.14. Magnitude against frequency (Hz) for 0.1´10-3 mol dm-3 of potasium ferricyanide (K3Fe(CN)6) / 0.1´10-3 mol dm-3 of potasium ferrocyanide (K4Fe(CN)6)  in 0.5 mol dm-3 H2SO4 in H2O. The ac modulation was 10 mV. Frequency range was between 65 KHz and 0.01 Hz.  Frequency and magnitude axes are in a logaritmic scale.

 

The change in phase behaviour between figures 8.10 and 8.13 can be analysed in terms of change in charge transfer resistant. Rct is given by

 


                                                                                                           [equation 8.57]

 

where i0 is defined by

 


                                                                                                           [equation 8.58]

 

n is the number of electrons involved in the electron transfer (in this case just one), A is the area of the working electrode, k0 is the standard rate constant that indicates the kinetic facility of the redox couple and a is the transfer coefficient that varies between one and zero. As the concentration of the redox couple is the same for both ions, equation 8.58 becomes

 

                                                                                                           [equation 8.59]

 

If the concentration of the redox couple is reduced by one order of magnitude, the charge transfer resistance is increased as it depends on the inverse of the concentration. A higher charge transfer resistance leads to the maximum in the semicircle occurring at lower frequency.  The change in the phase behaviour displayed in figures 8.10 and 8.13 is therefore explained.

 

Little difference is observed between the date in figure 8.11 and in figure 8.14.

 

Figures 8.15, 8.16 and 8.17 describe the ac impedance behaviour of 0.1´10-3 mol dm-3 of ferricsulfate (Fe2(SO4 )3)/ 0.1´10-3 mol dm-3 of ferrosulfate (FeSO4 ) in 0.5 mol dm-3 H2SO4 aqueous solution.

Figure 8.15. Imaginary component (Z²) against real part (Z¢) of the impedance for 0.1´10-3 mol dm-3 of ferricsulfate (Fe2(SO4 )3)/ 0.1´10-3 mol dm-3 of ferrosulfate (FeSO4 ) in 0.5 mol dm-3 H2SO4 aqueous solution. The ac modulation was 10 mV. Frequency range was between 65 KHz and 0.01 Hz.

 

A well defined semicircle is shown in the high frequency region in accordance with the theory. In the diffusion controlled region there is a not very well defined straight line.

 

Figure 8.16. Phase component (DEG) against frequency (Hz) for 0.1´10-3 mol dm-3 of ferricsulfate (Fe2(SO4 )3)/ 0.1´10-3 mol dm-3 of ferrosulfate (FeSO4 ) in 0.5 mol dm-3 H2SO4 aqueous solution. The ac modulation was 10 mV. Frequency range was between 65 KHz and   0.01 Hz. Frequency axe is in a logaritmic scale

 

There is a change in phase in the region between 1 Hz to 1000 Hz. This region relates the exchange of an electron in the redox couple.

Figure 8.17. Magnitude against frequency (Hz) for 0.1´10-3 mol dm-3 of ferricsulfate (Fe2(SO4 )3)/ 0.1´10-3 mol dm-3 of ferrosulfate (FeSO4 ) in 0.5 mol dm-3 H2SO4 aqueous solution. The ac modulation was 10 mV. Frequency range was between 65 KHz and 0.01 Hz. Magnitude and frequency axes are in a logaritmic scale.

 

8.5. Temperature modulation open circuit potential experiments

 

Pre-eliminary experiments indicate that the optimal conditions to commission temperature modulation open circuit potential experiments using the set up explained in the section 8.3.7 is to utilise the less concentrated ferrosulfate/ferricsulfate redox couple.

 

The experimental set up allows a study of a small range of frequency, approximately between 0.01 Hz and 4 KHz. When bigger values to 4 KHz were reached the signal was so noisy that signal could not be recovered. This limitation was related with the power of the infra red diode. Due to the limitation in the frequency region to be studied the couple to be chosen should have a change in phase in the region that can be measured. After checking all the plots related to the phase change as a function of the frequency (see figures 8.10, 8.13 and 8.16) the conclusion is that only ferrosulfate / ferricsulfate redox couple is in the range of the set up. The reason to use the less concentrated solution is that when the concentration is low the quality of the signal (less noise in the measurements) from the cell is optimal and much better defined plots can be obtained.

 

Figures 8.18, 8.19 and 8.20 describe the temperature modulation open circuit potential behaviour of 0.1´10-3 mol dm-3 of ferricsulfate (Fe2(SO4)3)/0.1´10-3 mol dm-3 of ferrosulfate (FeSO4) in 0.5 mol dm-3 H2SO4 aqueous solution.

Figure 8.18. Imaginary component (Z²) against real part (Z¢) of the temperature modulation open circuit experiment for 0.1´10-3 mol dm-3 of ferricsulfate (Fe2(SO4 )3)/ 0.1´10-3 mol dm-3 of ferrosulfate (FeSO4 ) in 0.5 mol dm-3 H2SO4 aqueous solution.  Frequency range was between     4 KHz and 0.01 Hz. The inset details the shape of the semicircle at high frequencies.

 

Figure 8.19. Phase component (DEG) against frequency (Hz) for 0.1´10-3 mol dm-3 of ferricsulfate (Fe2(SO4 )3)/ 0.1´10-3 mol dm-3 of ferrosufhate (FeSO4 ) in 0.5 mol dm-3 H2SO4 aqueous solution. Frequency range was between 4 KHz and 0.01 Hz. Frequency axe is in a logarithmic scale.

 

Figure 8.20. Magnitude (mV) against frequency (Hz) for 0.1´10-3 mol dm-3 of ferricsulfate (Fe2(SO4 )3)/ 0.1´10-3 mol dm-3 of ferrosulfate (FeSO4 ) in 0.5 mol dm-3 H2SO4 aqueous solution. Frequency range was between 4 KHz and 0.01 Hz. Magnitude and frequency axes are in a logarithmic scale.

 

Figure 8.18 shows a straight line very close to 45 degrees that suggests a diffusion process or composition of two similar electrochemical stages. Figure 8.19 suggests a possible change in phase between 1 Hz and 1000 Hz.

 

For further analysis of the temocps results adaptation in the mathematical developed done in section 8.2 to the experimental set-up is required.

Figure 8.21. Diagram of the experimental set up that details the ac and dc components in the system.

 

From equation 8.31:

 

Considering that the diamond samples used as substrates were estimated to be about 5 mm thick and the fact that the diamond samples were polycrystalline rather than single crystal all constants to be included where placed in a new experimentally adjustable parameter rD :

 

                                  [equation 8.60]

 

Differentiating the above equation

 

                                                      [equation 8.61]

Substituting equations 8.60 and 8.61 into equation 8.53:

 

           [equation 8.62]

 

To solve equation 8.62 only the ac component is considered. The aim is to link the total voltage change with the applied frequency in the experimental conditions:

 

Supposing that the total voltage change is due to pure ac component then

 

                where B is an arbitrary parameter      [equation 8.63]   

Differentiating equation 8.63 with respect to t yields:     

 

                                                                  [equation 8.64]

Substituting the equations 8.63 and 8.64 into equation 8.62 gives

 

                                                                                                           [equation 8.65]

Rearranging yields

 

                                                                                                           [equation 8.66]

Therefore:

 

                                                                                                           [equation 8.67]

The lock in measures                hence:

 

 

                                                                                                           [equation 8.68]

If bj and bs are small:

 

                         [equation 8.69] 

 

If no diffusion is observed

 

            Z = Rct                         Rct: charge transfer resistance               [equation 8.70]

 

 

 

If diffusion is observed

 

                                                                    [equation 8.71]

 

where s is defined in equation 8.56.

 

Mathcad 8 Profesional (MathSoft, Inc.) was operated to fit the equation 8.69 in diffusion conditions with the experimental data (figures 8.18, 8.19 and 8.20) Parameters for the equation 8.69 were taken from different sources.

 

Parameters for the gold (on diamond substrate) electrode:

 

·        Values of C (capacitance) and RF (faradaic resistance) are from the average of experimental impedance. 

 

·        Values of constants like F (faraday), DR and DO (diffusion coefficients), R (constant of gases) are from CRC handbook 39; bi is from Smalley et al. 38.

 

·        Values of T (temperature), CO and CR (concentration of the redox species) and Agold (gold area on diamond electrode) are from the experimental conditions.

 

·        Parameter, bf, is adjusted to give phase shift in frequency plot at correct place.

 

·        Parameter, r, acts as a scalar to fit the theoretical data with the experimental one.

 

All parameters are in SI units.

 

Calculating s

Plots:

Figure 8.22. Imaginary component (Z²) against real part (Z¢) of the temperature modulation open circuit for 0.1´10-3 mol dm-3 of ferricsulfate (Fe2(SO4 )3)/ 0.1´10-3 mol dm-3 of ferrosulfate (FeSO4) in 0.5 mol dm-3 H2SO4 aqueous solution. Black plus (+) and red cross (´) correspond with the experimental and the theoretical data respectively. Frequency range was between 4 KHz and 0.01 Hz. The inset details the shape of the semicircle at high frequencies.

 

Figure 8.23. Phase component (DEG) against frequency (Hz) for 0.1´10-3 mol dm-3 of ferricsulfate (Fe2(SO4 )3)/ 0.1´10-3 mol dm-3 of ferrosulfate (FeSO4 ) in 0.5 mol dm-3 H2SO4 aqueous solution. Black plus (+) and red cross (´) correspond with the experimental and the theoretical data respectively. Frequency range was between 4 KHz and 0.01 Hz. Frequency axis is on a logarithmic scale.

 

Figure 8.24. Magnitude (mV) against frequency (Hz) for 0.1´10-3 mol dm-3 of ferricsulfate (Fe2(SO4 )3)/ 0.1´10-3 mol dm-3 of ferrosulfate (FeSO4) in 0.5 mol dm-3 H2SO4 aqueous solution. Black plus (+) and red cross (´) correspond with the experimental and the theoretical data respectively. Frequency range was between 4 KHz and 0.01 Hz. Magnitude and frequency axes are on a logarithmic scale.

 

Figure 8.22 shows a good agreement between the theory and the experimental data. In the first plot the general shape, magnitude and slope are equivalent for theoretical and experimental data. In the second plot there is a reasonable fit between the experimental and the theoretical data. The dispersion of the experimental data can be attributed to the fact that when high frequencies are reached the level of noise in the experimental measurements raise. This increment of noise is explained in terms of the limits of the set up. As the limits of the set up are approached the magnitude and the phase are measured with much higher levels of noise. This disturbance in the magnitude and the phase clearly affects the impedance data in the region of the high frequencies.

 

Figure 8.23 shows a good agreement between the experimental phase and the theoretical one. However some dispersion of the data can be observed as the high frequencies region is reached. This fact can be linked again with the limits of the experimental set up and confirm the explanation of the data dispersion in the impedance measurements at high frequencies. The dispersion of the phase at high frequencies could be explained in terms of the relaxation phenomenon. When the frequency is increased the relaxation time for the cooling process of the electrode is reduced. As higher frequencies are reached lower values for the temperature perturbation are obtained increasing the noise in the measurements.

 

Figure 8.24 shows a nice fit between the theoretical and the experimental data. Again as was detailed for the impedance and phase plots some dispersion data is observed at the high frequencies region. The same arguments used before can be used again to explain this.

 

Some extra information from the experimental data needs to be emphasised. The value of rD relates the change of the temperature when the frequency tends to one. From equation 8.57 it can be calculated that the change of temperature at the gold surface at 1 Hz is 1.4´10-2 K. This small change of temperature at the electrode perturbs the electrochemical system in controlled conditions.

 

There is no doubt that the technique can be improved. Looking at the equation 8.31 in detail it is apparent that the power of the heating source is increased higher frequencies can be measured. If the domain of higher frequencies is to be investigated a more powerful heating source will be required. A laser with the appropriate wavelength will be an ideal heating source to perform successfully temocps at very high frequencies.

 

It has to be remarked that temocps impedance is measured independently and without the use of a potentiostat. Just giving the appropriate constants in the equation 8.69 precise impedance measurements can be performed. This fact opens the possibility of studying further range of frequencies where potentiostats and frequency response analysers are limited for the disruption that their electrical components originate at very high frequencies. Using this technique less investment will be necessary to perform impedance measurements.

 

8.6. Conclusions

 

A new method of impedance measurements has been developed using relative simple equipment. This simplicity avoids the limitations of more sophisticated apparatus allowing impedance studies over wide range of frequencies.

 

8.7. References

 

1              M. Eigen, Disc. Farad. Soc., 1954, 17, 194.

2              M. J. Pilling, 'Reaction Kinetics', ed. P. W. Atkins, Oxford University Press, 1975.

3              K. J. Laidler, 'Chemical Kinetics (Third Edition)', ed. L. S. Berger, Harper Collins Publishers, 1987.

4              P. Grundler, T. Zerihun, A. Kirbs, and H. Grabow, Anal. Chim. Acta, 1995, 305, 232.

5              K. Frischmuth, P. Visocky, and P. Grundler, Internat. J. Engin. Sci., 1996, 34, 523.

6              P. Grundler, A. Kirbs, and T. Zerihun, Analyst, 1996, 121, 1805.

7              T. Zerihun and P. Grundler, J. Electroanal. Chem., 1996, 415, 85.

8              T. Zerihun and P. Grundler, J. Electroanal. Chem., 1996, 404, 243.

9              P. Grundler, Fresenius J. Anal. Chem., 1998, 362, 180.

10             P. Grundler and G. U. Flechsig, Electrochim. Acta, 1998, 43, 3451.

11             T. Zerihun and P. Grundler, J. Electroanal. Chem., 1998, 441, 57.

12             A. Beckmann, A. Schneider, and P. Grundler, Electrochem. Commun., 1999, 1, 46.

13             P. Grundler and A. Kirbs, Electroanalysis, 1999, 11, 223.

14             M. Jasinski, A. Kirbs, M. Schmehl, and P. Grundler, Electrochem. Commun., 1999, 1, 26.

15             T. Voss, P. Grundler, A. Kirbs, and G. U. Flechsig, Electrochem. Commun., 1999, 1, 383.

16             T. Voss, P. Grundler, C. M. A. Brett, and A. M. O. Brett, J. Pharmac. Biomed. Anal., 1999, 19, 127.

17             J. Wang, P. Grundler, G. U. Flechsig, M. Jasinski, J. M. Lu, J. Y. Wang, Z. Q. Zhao, and B. M. Tian, Anal. Chim. Acta, 1999, 396, 33.

18             A. Beckmann, B. A. Coles, R. G. Compton, P. Grundler, F. Marken, and A. Neudeck, J. Phys. Chem. B, 2000, 104, 764.

19             P. Grundler, Fresenius J. Anal. Chem., 2000, 367, 324.

20             A. Schneider, G. U. Flechsig, and P. Grundler, J. Electrochem. Soc., 2000, 147, 3768.

21             T. Voss, A. Kirbs, and P. Grundler, Fresenius J. Anal. Chem., 2000, 367, 320.

22             J. Wang, P. Grundler, G. U. Flechsig, M. Jasinski, G. Rivas, E. Sahlin, and J. L. L. Paz, Anal. Chem., 2000, 72, 3752.

23             J. Wang, M. Jasinski, G. U. Flechsig, P. Grundler, and B. M. Tian, Talanta, 2000, 50, 1205.

24             G. U. Flechsig, O. Korbut, and P. Grundler, Electroanalysis, 2001, 13, 786.

25             P. Grundler and D. Degenring, J. Electroanal. Chem., 2001, 512, 74.

26             P. Grundler and D. Degenring, Electroanalysis, 2001, 13, 755.

27             M. Jasinski, P. Grundler, G. U. Flechsig, and J. Wang, Electroanalysis, 2001, 13, 34.

28             O. Korbut, M. Buckova, P. Tarapcik, J. Labuda, and P. Grundler, J. Electroanal. Chem., 2001, 506, 143.

29             C. Prado, G. U. Flechsig, P. Grundler, J. S. Foord, F. Marken, and R. G. Compton, Analyst, 2002, 127, 329.

30             J. L. Valdes and B. Miller, J. Phys. Chem., 1988, 92, 4483.

31             J. L. Valdes and B. Miller, J. Phys. Chem., 1988, 92, 525.

32             J. L. Valdes and B. Miller, J. Phys. Chem., 1989, 93, 7275.

33             A. Olivier, E. Merienne, J. P. Chopart, and O. Aaboubi, Electrochim. Acta, 1992, 37, 1945.

34             C. Gabrielli and B. Tribollet, J. Electrochem. Soc., 1994, 141, 1147.

35             C. Gabrielli, M. Keddam, and J. F. Lizee, J. Electroanal. Chem., 1993, 359, 1.

36             Z. A. Rotenberg, Russ. J. Electrochem., 1998, 34, 170.

37             Z. A. Rotenberg, Electrochim. Acta, 1997, 42, 793.

38             J. F. Smalley, C. V. Krishnan, M. Goldman, S. W. Feldberg, and I. Ruzic, J. Electroanal. Chem., 1988, 248, 255.

39          D. R. Linde, in ‘CRC Handbook of Chemistry and Physics’, ed. D. R. Linde, 1994