Next Article in Journal
Optimization of Anodized-Aluminum Pressure-Sensitive Paint by Controlling Luminophore Concentration
Previous Article in Journal
Intelligent Chiral Sensing Based on Supramolecular and Interfacial Concepts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Voltammetry under a Controlled Temperature Gradient

BVT Technologies, a.s., Hudcova 533/78c, 612 00 Brno, Czech Republic
*
Author to whom correspondence should be addressed.
Sensors 2010, 10(7), 6821-6835; https://doi.org/10.3390/s100706821
Submission received: 30 May 2010 / Revised: 20 June 2010 / Accepted: 1 July 2010 / Published: 14 July 2010
(This article belongs to the Section Chemical Sensors)

Abstract

:
Electrochemical measurements are generally done under isothermal conditions. Here we report on the application of a controlled temperature gradient between the working electrode surface and the solution. Using electrochemical sensors prepared on ceramic materials with extremely high specific heat conductivity, the temperature gradient between the electrode and solution was applied here as a second driving force. This application of the Soret phenomenon increases the mass transfer in the Nernst layer and enables more accurate control of the electrode response enhancement by a combination of diffusion and thermal diffusion. We have thus studied the effect of Soret phenomenon by cyclic voltammetry measurements in ferro/ferricyanide. The time dependence of sensor response disappears when applying the Soret phenomenon, and the complicated shape of the cyclic voltammogram is replaced by a simple exponential curve. We have derived the Cotrell-Soret equation describing the steady-state response with an applied temperature difference.

Graphical Abstract

1. Introduction

The theory of electrode processes is abundantly described in the literature [1,2]. The electrode response is controlled by electroactive compound transport between the electrode surface and bulk solution. The main mass transport driving forces are electromigration, convection and diffusion. We focus here on convection and diffusion under non-isothermal conditions, where the electrochemical sensor response is controlled by mass transfer between the bulk solution and the layer adhering to electrode surface as well as by the electron transfer kinetics in the electrode reaction. The main process controlling the transfer in the Nernst layer is diffusion. However, diffusion is a relatively slow process. It is much slower than the electrode reaction. The closest neighborhood of an electrode surface is depleted of electroactive compounds. Therefore, the electrode response decreases, as described by the Cotrell Equation [2]. The thickness of the layer where the electrochemical reaction takes place is 0.1 nm to 10 nm. Due to liquid viscosity, it is nearly impossible to use hydrodynamic forces (convection) to accelerate the mass transfer in a layer of this thickness [3].
If very strong hydrodynamic forces are used, i.e., the effective Reynold′s number is extremely high, big fluctuations will cause hydrodynamic noise in the proximity of this layer due to turbulent flow [4,5]. In a layer of a few nanometers adhering to the electrode, pure hydrodynamics are also nearly impossible to use as a tool for the improvement of mass transport between the bulk and the electrode surface. Hence, the question whether another driving force can be applied to increase the mass transport. We reasoned that applying a temperature gradient between the electrode and the solution could serve as a second driving force because thermodiffusion mass transport is independent of the concentration gradient.
The use of temperature to influence the electrode reaction was previously used by Gründler [6], who used a Pt wire heated by Joule heat. Laser pulses have been applied to heat the electrode surface [7]. In another method, the electrode was placed on an insulating layer on metal, which was heated by Joule heat [8,9].
Over the past two decades, attention has been devoted to the improvement of heated electrodes. Grundler and Zerihun have previously tested the functionality of a heated electrode in different arrangements for oxygen measurement, reaching similar cyclic voltammograms as obtained by us [10].
The mass transfer and microfluidics in the proximity of the heated electrode were solved by Frischmuth et al., while the thermodiffusion was not involved in their calculations [11]. Green et al. described the basic theory of electrothermally induced fluid flow. The study contains the simulations of the velocity field above the heated electrodes [12]. Oduoza performed electrochemical reaction simulations on the heated platinum wire. He described the mass transfer in terms of convection and hydrodynamics by means of the theory of similarity [13]. The work of Jian-Jun Sun et al. describes the graphite heated cylinder electrodes. The mass transport was solved by similarity theory using the similarity numbers, but it does not consider thermodiffusion [14]. Temperature field digital simulations around the wire heated electrodes are presented in another study [15].
Many results increase understanding of the phenomena on heated electrodes, but in many of the studies the methods of non-equilibrium thermodynamics were used. There are also many examples of the analytical use of heated electrodes, which prove the potential of this electrochemical research.
Heated electrodes have been used in many applications including lead detection [16], formaldehyde, methanol and formic acid oxidation [17], anodic stripping voltammetry on heated mercury film electrode [18], electrochemical behavior of cytochrome C [19], identification of DNA damage [20], interaction between DNA and metal complexes [21], electrochemistry of nicotinamide adenine dinucleotide [22], improvement of glucose and maltose sensor specificity [23], electrochemistry of ascorbic acid [24], rutin detection in the nanomolar range [25], thermal stabilization of glucose heated electrodes [26], tool to prevent biochemical fouling on electrodes [27], capillary electrophoresis detectors [28], flow detectors [29], and disposable electrodes [30].
Different concepts were reported regarding heated electrode preparation and use [3136]. Heated electrodes were combined with electroluminescence in several studies [14,37,38]. A very important application of heated electrodes lies in the emerging field of ionic liquids [39,40]. The comparison of the analytical efficiency of heated electrodes with other methods for detecting lead traces was introduced by Yonghong Li [41].
The heated electrodes seem to be a promising tool in electrochemistry. However, no simple and comprehensive theory of their electrochemical response has been presented yet. All published simulations neglect the fact that a thermal gradient starts thermodiffusion, which is a second mass transport mechanism that can be controlled independently of the concentration by changes in the temperature. This control offers a new parameter for electrochemical measurements, which is not sufficiently appreciated.
The question addresses the technology, which assures the possibility of exact control of temperature gradients at the electrode independently of the concentration. The microelectronic technologies enable the preparation of electrodes on ceramic materials with extremely high specific heat conductivity λ Compare the values provided in Table 1.
The heating resistor can be integrated at a distance of more than ten microns from the electrode. If we realize that the specific heat conductivity of a BeO ceramic is more than two orders higher than that of water, then the temperature of the working electrode surface can be controlled in a very precise manner. The temperature gradient is concentrated in the Nernst layer.
The equation describing the diffusion and temperature are formally the same (parabolic partial equations). Only the parameter α in the temperature equations is about four orders higher than D in the diffusion equations. On the other hand, the transport phenomena connected with non-diagonal elements of the (non-equilibrium thermodynamics) flux matrix are about two orders lower than the diagonal ones.

Theory

The derivation of the electrode response under isothermal conditions is widely described [1,2]. If a thermal gradient is induced, the system depends on non-equilibrium thermodynamics. The mass transport connected to the thermal gradient was first observed by Ludwig Soret in 1856, which represents a typical example of the coupling of two gradient-cross phenomena. If we suppose that the reaction on the electrode surface is sufficiently fast, then it is possible to assume that the concentration on the electrode surface is zero because the analyte is immediately consumed.
The entropy production σ can be expressed by the following equation (for the general case of several diffusing substances in a continuous, non-isothermal system where no chemical reaction takes place) [43]:
σ = J q grad 1 T + i = 1 n J i grad ( μ i T )
where:
σ
is the local entropy production,
Jq
is the flow of heat,
T
is the absolute temperature,
Ji
is the flow of component i in moles per unit area per unit time, and
μi
is the chemical potential of component i.
After some simplification, using the Onsager relations of reciprocity [43,3], the final equation for the flow in a binary system can be deduced:
J s d = L 11 μ ss grad ( C s ) L 1 q gradT T
where:
J s d
is the flow of the solute relative to the solvent in binary solutions,
Lik
is a phenomenological coefficient relating to the kth driving force (L11, L1q where q is heat),
μss
is the chemical potential of the solute in a binary system, and
Cs
is the molar concentration of the solute in a binary solution.
The total flow of solute depends on two terms. The first term describes the classical or ordinary diffusion proportional to the concentration gradient. The second term describes the thermal diffusion flow induced by the temperature gradient. The coefficient L11μss is a classical diffusion coefficient. The coefficient L1q/T is proportional to the cross non-diagonal coefficient in the classical phenomenological flow matrix in non-equilibrium thermodynamics. The coefficient L1q is proportional to the solute concentration Cs. Therefore, it is usual to define the thermal diffusion coefficient DT by the relation:
L 1 q T = C s D T
Equation (2) can then be rewritten in the form:
J s d = D grad ( C s ) C s D T grad ( T )
where D is the diffusion coefficient. Thermal diffusion is often characterized by the Soret coefficient sT, which is the ratio of thermal diffusion coefficient to the ordinary diffusion coefficient:
s T = D T D = L 1 q DC s T
Comparing this definition with Equation (4), it is clear that the Soret coefficient describes the equilibrium where the magnitude of flow caused by the gradient of the temperature is exactly equal to the flow caused by the gradient of the concentration [42]. Under these conditions (3), Soret coefficients can be expressed as:
s T = grad ( ln C s ) gradT
It was found that the thermal diffusion coefficient is smaller by a factor from 100 to 1,000 than the ordinary diffusion coefficient for electrolytes, non electrolytes and gases [43]. The concentration gradient is relatively small unless the thermal gradient is very large. However, a completely different situation is solved here. Due to the very high thermal conductivity of ceramics, the thermal gradient in the liquid adjacent to the electrode surface reaches very high values of the temperature gradient, and in addition, the temperature gradient can be adjusted independently of the solute concentration Cs. The thermodiffusion can play an important role in the enhancement of the mass transfer in the proximity of the surface of the electrode. Other new results can be found in the literature [44,45].

Theory of the potential steps at an planar electrode including thermo diffusion mass transport

Consider the reaction [1] O + neR, which is started by a potential step of any magnitude. An experiment begins at t = 0 and at a potential at which no current flows. The potential E is instantaneously changed to a value anywhere on the reduction wave. Very rapid charge transfer kinetics are assumed here, so O and R are always in equilibrium at the electrode surface, with the concentration ratio given by the Nernst equation:
θ = C O ( 0 , t ) C R ( 0 , t ) = exp nF ( E E ) RT 1
where:
CO(x,t)
is the concentration of the oxidized compound (reducible),
CR(x,t)
is the concentration of the reduced compound (oxidable),
n
is the number of electrons in the electrode reaction,
F
is the Faraday constant,
E
is the electrode potential,
E0
is the standard electrode reaction potential,
R
is the gas constant,
T1
is the temperature of the electrode surface, and
t
is the time.
Not only a voltage but also a temperature gradient is applied there. The planar electrode (Figure 1) is maintained at temperature T1, and the solution surrounding the electrode is maintained at temperature T2 by an external thermostat.
The problem can be specified as follows:
  • - The electrode is planar and has a surface area A. Only diffusion along the x-axis perpendicular to the electrode surface needs to be considered. The electrode is sufficiently large so that the edge effects can be neglected.
  • - The solution is initially homogenous. Specifically, the initial concentration of O and R are CO(x,0) = CO and CR(x,0) = 0 for all values of x at time t = 0.
  • - The electrolysis cell is sufficiently large that the bulk concentrations of O and R are unchanged from the initial values even after electrolysis has been running for a certain time. In other words, CO(x,t)→CO, CR(x,t)→0 while x→∞.
  • - For every O molecule consumed, an R molecule is formed; in other words, the fluxes JO(0,t) and JR(0,t) of O and R at the electrode surface are equal and opposite in sign: JO(0,t) = −JR(0,t).
  • - The electron-transfer reaction is very fast, such that O and R are always in equilibrium at the electrode surface with the concentration ratio given by the Nernst Equation (7).
  • - The system can be treated as two binary systems of (solvent and O) and (solvent and R). There is no interaction between O and R.
  • - The reaction takes place at the electrode surface, the temperature of which is T1.
There is only one difference from the classical solution. The flux depends on the temperature gradient and on the concentration gradient:
J ( x , t ) = D C ( x , t ) x D . C ( x , t ) . s T . T ( x , t ) x
We suppose that the system is in thermal equilibrium, which means:
T ( x , t ) x = α . Δ T = α ( T 2 T 1 )
where α is a proportionality constant. The electrode is heated (T1 > T2). The final diffusion equations are derived using Fick’s laws:
t C Ox ( t , x ) = D Ox ( 2 x 2 C Ox ( t , x ) ) + σ Ox ( x C Ox ( t , x ) )
t C Re d ( t , x ) = D Re d ( 2 x 2 C Re d ( t , x ) ) + σ Re d ( x C Re d ( t , x ) )
where σOx = DOx.sT,Ox.α(T1−T2) and σRed = DRed.sT,Red.α(T1−T2). DOx and DRed are diffusion coefficients for O and R, respectively. sT, Ox and sT,Red are the Soret coefficients for O and R, respectively.
The definition of the problem implies the boundary conditions at infinity:
lim x C Ox ( t , x ) = C 0 , lim x C Re d ( t , x ) = 0 .
The boundary conditions at the surface of electrode are expressed through the flux balance, which means that the flux of the oxidized compound is in equilibrium with the flux of the reduced compound. The second condition states that all oxidized compounds are immediately changed to the reduced form at the electrode surface:
D Ox ( ( x C Ox ( t , x ) ) x = 0 ) + σ Ox C Ox ( t , x ) = D Re d ( ( x C Re d ( t , x ) ) x = 0 ) + σ Re d C Re d ( t , x )
C Ox ( t , 0 ) C Re d ( t , 0 ) = e ( nF ( E E 0 ) RT 1 )
We can denote the exponential term as θ:
θ ( T 1 ) = e ( n F ( E E 0 ) R T 1 )
The initial conditions are: COx(0,x) = C0, CRed(0,x) = 0.
In the most simple case it is possible to consider DOx = DRed = D and σOx = σRed = σ.
The application of the Laplace transform to equations (10) and (11) in consideration of conditions (12) yields:
i = i 0 ( e Bt π t + B . erf ( Bt ) )
where:
i 0 = nFAC 0 D 1 + θ ( T 1 ) ,
B = D . [ s T . α ( T 1 T 2 ] 2 ,
and i0 is steady state current.
Equation (16) has very interesting impacts. At zero temperature gradient at the electrode, T1 = T2, the coefficient B = 0, erf (0) = 0, and it becomes the classical Cotrell equation.
If T1T2 then B0 and e Bt π t decreases quickly, while the second term exponentially tends to B, which is time independent. If the temperature gradient is present, the current approaches a steady state value given by equation:
i = n . F . A . C 0 D 1 + θ ( T 1 ) . s T . α . ( T 1 T 2 )
where θ(T1) stresses the fact that due to high temperature conductivity of ceramic substrate of the sensor the T1 temperature of the electrode surface is known.

2. Experimental Section

Measurements were performed in a device consisting of a glass cell TC1, conic stirrer and connector KSA1 and electrochemical sensor AC1.W2.RS (H,T) (BVT Technologies, Czech Republic). The AC1.W2.RS (H) electrochemical sensor bears platinum working and auxiliary electrodes, a pseudo-reference silver electrode and a heating circuit. The cell TC1 was placed in a small thermostat TK-1 (KEVA, Czech Republic). The whole system schematic is shown in Figure 2.

3. Results and Discussion

Formula (16) plays a key role in this new method. If the temperature difference is zero, then the parameter B is also zero. Then, the exponential term vanishes, and the second term vanishes as well. The formula changes to the normal Cottrell equation. However, if the temperature difference is nonzero, then B is also nonzero. The influence on Equation (16) is dramatic.
The classical Cottrell term decreases with an exponential rate. The second term, containing the error function, converges very fast to the value of the square root of B. This value is constant and time independent. The situation is shown in Figure 3.
The experimental exponent −0.47 is close to the theoretical value −0.5 in the case of no temperature gradient; however, if the temperature is implied the response stabilizes to a steady state current i0.sT. (T1−T2). The current is stabilized in 2 s in comparison with pure diffusion where the current is not stabilized after 60 s. As the Cottrell term vanishes with an exponential rate, after a short time the response of the sensor is controlled by the thermo diffusion term only.
In the case of cyclic voltammetry measurements the response is changed dramatically. The higher the temperature difference between electrode and solution is, the narrower oxidation and reduction wave is until it completely disappears. This situation is visible in Figure 4.
At a temperature difference of 35 °C, the reduction and oxidation curves are the same. The thermodiffusion under such conditions is sufficiently strong that the first term in (16) can be omitted. The cyclic voltammetry can be described by (19), where the voltage is in the term θ (T1) = exp (nF(E−E0)/RT1).
The complicated shape of the cycling voltammetry disappears and at a temperature difference of 35 °C only a very simple curve remains, which can be analyzed in a simple manner using formula (19). The agreement between experiment and theory is again very good. The same or very similar experimental results were also obtained in [21,22,37,35].
It is also possible to see that the curves shift to the left. The slope of curve is also increased. This phenomenon can be understood if we realize that the reaction takes place on the electrode surface, the temperature of which is growing. The solvent is maintained at the temperature T2, and the temperature of the electrode T1 is increased.
Liquid flow in the vicinity of or above the heated surface can be solved using a different point of view. Lorenz [46] studied a hydrodynamic system that was an approximation of flow above the earth surface. He found out that deterministic differential equations may provide a solution that seems to be random noise with nondeterministic behavior. His work led to the “strange attractor” concept [47]. A set of convective flows originates in the layer above the heated surface (see Figure 5) [48].
Regarding the electrochemical measurements, the thermally induced microconvection is focused in a Nernst layer of a few microns in thickness. Therefore, the temperature gradient is applicable for micro to nano mixing, which is localized in the very vicinity of the working electrode. The experiments performed did not answer whether the prevailing mechanism is microconvection or pure thermodiffusion.
Bringuier [44] published the kinetic theory of colloid thermal diffusion. Figure 6 presents his very instructive illustration of flow behavior under a temperature gradient. This figure shows that the application of a temperature gradient significantly changes the flow at the surface of an electrode. Even though these results were developed for monatomic gases it can give a picture of nature of phenomena that can occur at electrode surface as well.

4. Conclusions

The time dependence of the sensor response disappears when applying the Soret phenomenon. The sensor current stabilizes to constant value, and the complicated shape of cyclic voltammogram is replaced by simple exponential curve. The Cotrell-Soret equation describing the steady-state response with applied temperature difference is then:
i = n . F . A . C 0 . D . s T . α , ( T 1 T 2 ) 1 1 + e nF ( E E 0 ) RT 1
where T1 is the temperature of the sensor, and T2 is the temperature of the liquid.
If a voltammetric scan is performed in a broad potential window, one can identify a current value where thermodiffusion is in equilibrium with the electrode reaction: i0 = −n.F.A.C0.sT.α.(T1−T2).
Then Equation (20) allows computation of E0 at temperature T1:
i i 0 = 1 1 + e nF ( E E 0 ) RT 1

List of Symbols

A
area of electrode surface
CO(x,t)
the concentration of oxidized compound (reducible)
CR(x,t)
the concentration of reduced compound (oxidable)
Cs
molar concentration of solute in a binary solution
C0
initial molar concentration
DOx, DRed
diffusion coefficients for Ox and Red, respectively
E
cell potential
E0
standard redox potential
F
Faraday constant
i
electric current
i0
steady state current
J
molar flux density
Ji
flow of component i in moles per unit area per unit time
Jq
is a flow of heat
Jsd
flow of solute relative to solvent in binary solutions
Lik
phenomenological coefficient relating to the kth driving force (L11, L1q where q is heat)
λ
specific heat conductivity
n
number of electrons in electrode reaction
R
gas constant
sT, Ox and sT,Red
Soret coefficients for Ox and Red, respectively
t
time
T
absolute temperature
T1
temperature of electrode surface
T2
temperature of liquid
α
proportionality constant
μi
chemical potential of component i
μss
chemical potential of solute in a binary system

Acknowledgments

This work was supported by BVT Technologies, a.s. This work was partly funded by the Czech national project IBIS FT-TA/089 and NIMS—KAN200520702 and 7RP EU project InFuLoc 230749. We would like to acknowledge the valuable comments of Emil Paleček (Institute of Biophysics, Academy of Sciences of the Czech Republic, Brno), Paul Duckworth and Boris Schlensky (eDAQ Pty Ltd, Australia).

References

  1. Bard, AJ; Faulkner, LR. Electrochemical Methods–Fundamentals and Applications; John Wiley & Sons: New York, NY, USA, 1980. [Google Scholar]
  2. Rieger, PH. Electrochemistry; Prentice Hall: Engelwood Cliffs, NJ, USA, 1987. [Google Scholar]
  3. Leal, LG. Advanced Transport Phenomena: Fluid Mechanics and Convective Transport Processes; Cambridge University Press: Cambridge, UK, 2007. [Google Scholar]
  4. Bird, RB; Steward, WE; Lightfoot, EN. Transfer Phenomena, 2nd ed; John Wiley & Sons: Toronto, Canada, 2001; p. 912. [Google Scholar]
  5. Frost, W; Moulden, TH (Eds.) Handbook of Turbulence Volume 1 Fundamentals and Applications; Plenum Press: New York, NY, USA, 1977; p. 498.
  6. Gründler, P; Zerihun, T; Kirbs, A; Grabow, H. Simultaneous joule heating and potential cycling of cylindrical microelectrodes. Anal. Chim. Acta 1995, 305, 232–240. [Google Scholar]
  7. Smalley, JF; Geng, L; Feldberg, SW; Rogers, LC; Leddy, J. Evidence for adsorption of Fe(CN)63−/4− on gold using the indirect laser-induced temperature-jump method. J. Electroanal. Chem 1993, 356, 181–200. [Google Scholar]
  8. Harima, Y; Aoyagui, S. Electrode potential relaxation following a rapid change of temperature: Part II. Theory. J. Electroanal. Chem 1977, 81, 47–52. [Google Scholar]
  9. Harima, Y; Aoyagui, S. Electrode potential relaxation following a rapid change of temperature. J. Electroanal. Chem 1976, 69, 419–422. [Google Scholar]
  10. Gründler, P; Zerihun, T. Electrically heated cylindrical microelectrodes. The reduction of dissolved oxygen on Pt. J. Electroanal. Chem 1996, 404, 234–238. [Google Scholar]
  11. Frischmuth, K; Visocky, P; Gründler, P. On modeling heat transfer in chemical microsensors. Int. J. Eng. Sci 1996, 34, 523–530. [Google Scholar]
  12. Green, NG; Ramos, A; González, A; Castellanos, A; Morgan, H. Electrothermally induced fluid flow on microelectrodes. J. Electrostat 2001, 53, 1–87. [Google Scholar]
  13. Oduoza, CF. Electrochemical mass transfer at a heated electrode in a vertical annular flow cell. Chem. Eng. Process 2004, 43, 921–928. [Google Scholar]
  14. Sun, JJ; Guo, L; Zhang, DF; Yin, WH; Chen, GN. Heated graphite cylinder electrodes. Electrochem. Commun 2007, 9, 283–288. [Google Scholar]
  15. Mahnke, N; Markovic, A; Duwensee, H; Wachholz, F; Flechsig, GU; van Rienen, U. Numerically optimized shape of directly heated electrodes for minimal temperature gradients. Sens. Actuat. B 2009, 137, 363–369. [Google Scholar]
  16. Zerihun, T; Gründler, P. Electrically heated cylindrical microelectrodes. Determination of lead on Pt by cyclic voltammetry and cathodic stripping analysis. J. Electroanal. Chem 1996, 415, 85–88. [Google Scholar]
  17. Zerihun, T; Gründler, P. Oxidation of formaldehyde, methanol, formic acid and glucose at ac heated cylindrical Pt microelectrodes. J. Electroanal. Chem 1998, 441, 57–63. [Google Scholar]
  18. Jasinski, M; Kirbs, A; Schmehl, M; Gründler, P. Heated mercury film electrode for anodic stripping voltammetry. Electrochem. Commun 1999, 1, 26–28. [Google Scholar]
  19. Voss, T; Gründler, P; Brett, CMA; Oliveira Brett, AM. Electrochemical behavior of cytochrome c at electrically heated microelectrodes. J. Pharm. Biomed. Anal 1999, 19, 127–133. [Google Scholar]
  20. Korbut, O; Buckova, M; Tarapcik, P; Labuda, J; Gründler, P. Damage to DNA indicated by an electrically heated DNA—modified carbon paste electrode. J. Electroanal. Chem 2001, 506, 143–148. [Google Scholar]
  21. Wang, J; Gründler, P. The influence of temperature on the interaction between DNA and metal complec at a heated gold-wire microelectrode. J. Electroanal. Chem 2003, 540, 153–157. [Google Scholar]
  22. Lau, C; Flechsig, GU; Gründler, P; Wang, J. Electrochemistry of nicotinamide adenine dinucleotide (reduced) at heated platinum electrodes. Anal. Chim. Acta 2005, 554, 74–78. [Google Scholar]
  23. Lau, C; Borgmann, S; Maciejewska, M; Ngounou, B; Gründler, P; Schuhmann, W. Improved specificity of reagentless amperometric PQQ-sGDH glucose biosensors by using indirectly heated electrodes. Biosens. Bioelectron 2007, 22, 3014–3020. [Google Scholar]
  24. Ke, JH; Tseng, HJ; Hsu, ChT; Chen, JCh; Muthuraman, G; Zen, JM. Flow injection analysis of ascorbic acid based on its thermoelectrochemistry at disposable screen-printed carbon electrodes. Sens. Actuators, B 2008, 130, 614–619. [Google Scholar]
  25. Wu, SH; Sun, JJ; Zhang, DF; Lin, ZB; Nie, FH; Qiu, HY; Chen, GN. Nanomolar detection of rutin based on adsorptive stripping analysis at single-sided heated graphite cylindrical electrodes with direct current heating. Electrochim. Acta 2008, 53, 6596–6601. [Google Scholar]
  26. Tseng, TF; Yang, YL; Chuang, MC; Lou, SL; Galik, M; Flechsig, GU; Wang, J. Thermally stable improved first-generation glucose biosensors based on Nafion/glucose-oxidase modified heated electrodes. Electrochem. Commun 2009, 11, 1819–1822. [Google Scholar]
  27. Duwensee, H; Vázquez-Alvarez, T; Flechsig, GU; Wang, J. Thermally induced electrode protection against biofouling. Talanta 2009, 77, 1757–1760. [Google Scholar]
  28. Chen, Y; Lin, Z; Chen, J; Sun, J; Zhang, L; Chen, G. New capillary electrophoresis-electrochemiluminescence detection system equipped with an electrically heated Ru(bpy)32+/multi-wall-carbon-nanotube paste electrode. J. Chromatogr. A 2007, 1172, 84–91. [Google Scholar]
  29. Wang, J; Jasinski, M; Flechsig, GU; Grundler, P; Tian, B. Hot-wire amperometric monitoring of flowing streams. Talanta 2000, 50, 1205–1210. [Google Scholar]
  30. Jenkins, DM; Song, C; Fares, S; Cheng, H; Barrettino, D. Disposable thermostated electrode system for temperature dependent electrochemical measurements. Sens. Actuator. B 2009, 137, 222–229. [Google Scholar]
  31. Voß, T; Gründler, P; Kirbs, A; Flechsig, GU. Temperature pulse voltammetry: hot layer electrodes made by LTCC technology. Electrochem. Commun 1999, 1, 383–388. [Google Scholar]
  32. Wang, J; Gründler, P; Flechsig, GU; Jasinski, M; Lu, J; Wang, J; Zhao, Z; Tian, B. Hot-wire stripping potentiometric measurements of trace mercury. Anal. Chim. Acta 1999, 396, 33–37. [Google Scholar]
  33. Qiu, F; Compton, RG; Coles, BA; Marken, F. Thermal activation of electrochemical processes in a Rf-heated channel flow cell: experiment and finite element simulation. J. Electroanal. Chem 2000, 492, 150–155. [Google Scholar]
  34. Coles, BA; Moorcroft, MJ; Compton, RG. Non – aqueous electrochemical studies at a high temperature channel flow cell heated by radio frequency radiation. J. Electroanal. Chem 2001, 513, 87–93. [Google Scholar]
  35. Wachholz, F; Biała, K; Piekarz, M; Flechsig, GU. Temperature pulse modulated amperometry at compact electrochemical sensors. Electrochem. Commun 2007, 9, 2346–2352. [Google Scholar]
  36. Xu, H; Xing, S; Zeng, L; Xian, Y; Shi, G; Jin, L. Microwave-enhanced voltammetric detection of copper(II) at gold nanoparticles-modified platinum microelectrodes. J. Electroanal. Chem 2009, 625, 53–59. [Google Scholar]
  37. Lin, Z; Sun, J; Chen, J; Guo, L; Chen, G. A new electrochemiluminescent detection system equipped with an electrically controlled heating cylindrical microelectrode. Anal. Chim. Acta 2006, 564, 226–230. [Google Scholar]
  38. Lin, Z; Sun, J; Chen, J; Guo, Li; Chen, G. The electrochemiluminescent behavior of luminol on an electrically heating controlled microelectrode at cathodic potential. Electrochim. Acta 2007, 53, 1708–1712. [Google Scholar]
  39. Chen, Y; Chen, X; Lin, Z; Dai, H; Qiu, B; Sun, J; Zhang, L; Chen, G. An electrically heated ionic-liquid/multi-wall carbon nanotube composite electrode and its application to electrochemiluminescent detection of ascorbic acid. Electrochem. Commun 2009, 11, 1142–1145. [Google Scholar]
  40. Lin, Z; Chen, Xi; Chen, H; Qiu, B; Chen, G. Electrochemiluminescent behavior of N6-isopentenyl-adenine/Ru(bpy)32+ system on an electrically heated ionic liquid/carbon paste electrode. Electrochem. Commun 2009, 11, 2056–2059. [Google Scholar]
  41. Li, Y; Liu, X; Zeng, X; Liu, Y; Liu, Xi; Wei, W; Luo, S. Simultaneous determination of ultra-trace lead and cadmium at a hydroxyapatite-modified carbon ionic liquid electrode by square-wave stripping voltammetry. Sens. Actuat. B 2009, 139, 604–610. [Google Scholar]
  42. Katchalsky, A; Curran, PF. Nonequilibrium Thermodynamics in Biophysics; Harvard University Press: Cambridge, MA, USA, 1967. [Google Scholar]
  43. Jost, J. Diffusion in Solids, Liquid and Gases; Academic Press: New York, NY, USA, 1960. [Google Scholar]
  44. Bringuier, E; Bourdon, A. Kinetic theory of colloid thermodiffusion. Physica A 2007, 385, 9–24. [Google Scholar]
  45. Bringuier, E. Kinetic theory of inhomogeneous diffusion. Physica A 2009, 388, 2588–2599. [Google Scholar]
  46. Lorenz, EN. Deterministic Nonperiodic Flow. J. Atmos. Sci 1963, 20, 130–141. [Google Scholar]
  47. Sparrow, C. Applied mathematical sciences. In The Lorenz Equations: Bifurcations, Chaos and Strange Attractors; Springer-Verlag: New York, NY, USA, 1982. [Google Scholar]
  48. Holodniok, M; Klic, A; Kubicek, M; Marek, M. Metody analyzy nelinearnich dynamickych modelu; Academia: Prague, Czechoslovakia, 1986. [Google Scholar]
Figure 1. Schematics of temperature gradient application on planar electrode realized on ceramics and immersed in the solution analyzed.
Figure 1. Schematics of temperature gradient application on planar electrode realized on ceramics and immersed in the solution analyzed.
Sensors 10 06821f1
Figure 2. Schema of the Soret system (the gap between the cone and electrode surface is 1 mm).
Figure 2. Schema of the Soret system (the gap between the cone and electrode surface is 1 mm).
Sensors 10 06821f2
Figure 3. The time dependence of sensor output current; comparison of Cottrell response (▪ T1 = T2) and response with thermo diffusion (▴T1T2 = 35 °C)
Figure 3. The time dependence of sensor output current; comparison of Cottrell response (▪ T1 = T2) and response with thermo diffusion (▴T1T2 = 35 °C)
Sensors 10 06821f3
Figure 4. The change of cyclic voltammetry with application of thermal gradient. The ▪ points are experimental points, and the continuous line is the theoretical run of function 1/(1 + exp((nF.(E−E0)/RT)).
Figure 4. The change of cyclic voltammetry with application of thermal gradient. The ▪ points are experimental points, and the continuous line is the theoretical run of function 1/(1 + exp((nF.(E−E0)/RT)).
Sensors 10 06821f4
Figure 5. Convective flow above the heated surface.
Figure 5. Convective flow above the heated surface.
Sensors 10 06821f5
Figure 6. Particle and heat flows in a monoatomic gas subjected to a temperature gradient at steady state. The spectral particle-current density J2(E) shows that “cold” molecules (E < 5 kT/2) go to the hot region (J2(E) > 0), and vice versa; the net particle-current density ∫J2(E)dE vanishes. The heat-current density jQ is ∫EJ2 (E) dE. Reprinted from Physica A, 385, E. Bringuier, A. Bourdon, Kinetic theory of colloid thermodiffusion, 9–24, Copyright (2007), with permission from Elsevier.
Figure 6. Particle and heat flows in a monoatomic gas subjected to a temperature gradient at steady state. The spectral particle-current density J2(E) shows that “cold” molecules (E < 5 kT/2) go to the hot region (J2(E) > 0), and vice versa; the net particle-current density ∫J2(E)dE vanishes. The heat-current density jQ is ∫EJ2 (E) dE. Reprinted from Physica A, 385, E. Bringuier, A. Bourdon, Kinetic theory of colloid thermodiffusion, 9–24, Copyright (2007), with permission from Elsevier.
Sensors 10 06821f6
Table 1. Values for thermal conductivity (λ), thermal diffusion (α) and diffusion coefficient for small molecules like water.
Table 1. Values for thermal conductivity (λ), thermal diffusion (α) and diffusion coefficient for small molecules like water.
MaterialλαD
W/m·Km2/sm2/s
Water at 25 °C0.60.14 × 10−610−10–10−12*
Al2O3 ceramic358.8 × 10−6
BeO18089 × 10−6
Ag420172 × 10−6
*Low molecular weight compounds.

Share and Cite

MDPI and ACS Style

Krejci, J.; Sajdlova, Z.; Krejci, J., Jr.; Marvanek, T. Voltammetry under a Controlled Temperature Gradient. Sensors 2010, 10, 6821-6835. https://doi.org/10.3390/s100706821

AMA Style

Krejci J, Sajdlova Z, Krejci J Jr., Marvanek T. Voltammetry under a Controlled Temperature Gradient. Sensors. 2010; 10(7):6821-6835. https://doi.org/10.3390/s100706821

Chicago/Turabian Style

Krejci, Jan, Zuzana Sajdlova, Jan Krejci, Jr., and Tomas Marvanek. 2010. "Voltammetry under a Controlled Temperature Gradient" Sensors 10, no. 7: 6821-6835. https://doi.org/10.3390/s100706821

Article Metrics

Back to TopTop