Вы находитесь на странице: 1из 7

519

Feature Article

Polythiophene Derivative Conducting Polymer Modified Electrodes and Microelectrodes for Determination of Ascorbic Acid. Effect of Possible Interferents
Stelian Lupu,a Adele Mucci,b Laura Pigani,b Renato Seeber,b * and Chiara Zanardib
a

Department of Analytical Chemistry and Instrumental Analysis, Faculty of Industrial Chemistry, University Politehnica of Bucharest, Calea Grivitei 132, 78122 Bucharest, Romania degli Studi di Modena e Reggio Emilia, via G. Campi 183, 41100 Modena, Italy Dipartimento di Chimica, Universita e-mail: seeber@unimo.it

Received: March 29, 2001 Final version: July 25, 2001 Abstract Conventional-size electrodes and microelectrodes coated by electrogenerated poly[4,4'-bis(butylsulfanyl)-2,2'bithiophene] have been tested with respect to the electrocatalytic oxidation of a particularly interesting analyte, i.e., ascorbic acid, in different concentrations of phosphate buffer, aqueous solution. Linear sweep and cyclic voltammetry have been used and the quantities related to the analyte concentration were the peak current and the diffusion (t1/2) deconvoluted peak current in the case of conventional-size and of microelectrode, respectively. Fairly good linear correlation could be found; a particularly wide linearity range was obtained by working with the microelectrode. It showed to give good results also at a very low (104 M) phosphate buffer-supporting electrolyte concentration. The actual interference on the analysis of compounds often coupled with ascorbic acid in natural or pharmaceutical products has been studied. Keywords: Modified electrode, Microelectrode, Amperometric sensors, Electrocatalysis, Ascorbic acid

1. Introduction
Ascorbic acid (vitamin C, AA) is a water-soluble substance present in a wide number of foods, such as fruits and vegetables. It is also added to foodstuffs as an antioxidant, for stabilization of color and aroma, as well as for prolonging the life of commercial products. Pharmaceuticals also often contain AA, as a supplementary source to human diets, for its important role as free-radical scavenger, which may help prevent free radical-induced diseases, such as cancer and Parkinsons disease. Since interest in oxidative changes in food products is increasing more and more, the possibility of determining the AA content during production, transformation, and storage of all these products has become particularly important. For this reason many analytical procedures have been proposed for AA detection, in many different matrices and at many different concentration levels. Traditional procedures for AA determination are generally based on enzymatic methods [1], on titration with oxidizing agent, like iodine or 2,6-dichloro-phenolindophenol [2] and on HPLC analysis with fluorimetric [2] or UVvisible detection [3, 4]. Electroanalytical techniques have also been frequently used for this purpose. Electrochemical oxidation of AA at different bare electrodes has been examined [5 7], the oxidation mechanism at there conventional electrodes being well known [8]. It is quite a complicated one: first a one-electron transfer leads to the dehydroascorbic acid anion radical that is adsorbed on the electrode surface, where it is further oxidized to adsorbed dehydroascorbic acid; this species undergoes slow deElectroanalysis 2002, 14, No. 78

adsorption to hydrolyze to the final dihydrate species. Electrode fouling by the intermediate adsorbed species takes place and the relevant voltammetric response was poorly reproducible. For this reason, a wide series of redox catalysts, ranging from metal ions and metal oxides to redox organic polymers, to macrocyclic complexes and to Prussian Blue, have been proposed as electrode coatings in order to mediate the oxidation of AA [9 11], strongly reducing the oxidation overpotential of the electrochemical process. For example, metal porphyrine-modified electrodes can lower the potential required for AA oxidation by several hundreds mV; the decrease of the charge transfer overvoltage renders the rate of the electron transfer so much faster that the voltammetric response becomes much better defined [12]. Osmium-containing redox polymer-modified electrodes are also reported to induce easier AA oxidation [13]; other electrodes used for the electrocatalytic determination of AAwere modified with 3,4-dihydroxybenzaldehyde [14], Ni (II) [15, 16], or Zn(II) [17] macrocyclic complexes. Recently, the use of electrodes modified with conducting polymers such as polyaniline, polypyrrole [18], and poly(3methylthiophene) [18 21] has received considerable attention with respect to amperometric quantitative analyses for AA determination. This can be ascribed to the advantages of such modified electrodes over conventional ones with respect to efficient electrocatalysis, sensitivity, selectivity, and low detection limit. Lyons and co-workers studied polypyrrole coated electrodes doped with chloride and dodecylbenzene sulfonate for electrocatalytic determination of AA [22]. Casella and Guascito [23] demonstrated

WILEY-VCH Verlag GmbH, 69469 Weinheim, Germany, 2002 1040-0397/02/0704-0519 $ 17.50+.50/0

520 that polyaniline film acts as an efficient electrocatalyst for AA oxidation. Chen and co-workers [24, 25] reported the decrease of ascorbate oxidation overpotential on a microdisk electrode covered by a copolymer of aniline with maminobenzoic acid and 3,4-dihydroxybenzoic acid. Polyaniline modified electrodes are also reported by Guilbault and co-workers [26] to selectively catalyze the oxidation of AA, strongly minimizing the effect of interferents present in the real sample, such as paracetamol, uric and citric acids. The use of (ultra)microelectrodes is very appealing when aiming to obtain an electrochemical sensor, thanks to the capability of working in low conductive media [27]. Moreover, the higher mass transport rate at a microelectrode reduces the detection limit of the analyte and permits a short response time. These peculiarities, together with the ability of a conducting polymer to incorporate ions during p-and ndoping, were exploited for the detection of electroinactive anions [28, 29]. The present article reports the electrochemical behavior of AA at both conventional-size and micro Pt electrodes modified with poly[4,4'-bis(butylsulfanyl)-2,2'-bithiophene] (PBSBT), as well as the performance of similar systems in the quantitative estimation of AA, under different experimental conditions, also in the presence of possible interferents of practical interest. Our previous studies have shown that this polymer possesses particularly interesting properties and extraordinary good stability [30 34], so much so that we choose using it when carrying out timeconsuming measurements in which repeatability has to be assured. Under these conditions, in fact, usual polymers easily undergo modifications that render the electrode somehow different when performing the very first and the very last run, respectively.

S. Lupu et al.

saturated calomel electrode (SCE, Amel) as the reference electrode and a glassy carbon rod as the auxiliary electrode. Alternatively, the working electrode was a homemade 10 mm diameter Pt disk microelectrode, assembled by sinking a thin Pt wire (Goodfellow, 99.99 % pure) inside a double component insulating epoxy glue. The success of the operation, as well as the evaluation of the actual electroactive area of the resulting disk microelectrode, were performed by a test in 103 M ferrocene, acetonitrile solution, using 0.2 M LiClO4 as the supporting electrolyte [36]. The working electrodes were polished subsequently with 1, 0.3, and 0.05 mm alumina powder to a mirror finish, dipped into an ultrasonic bath for 5 min, and then rinsed with doubly distilled water. The pH of the solutions was measured with a Denver Instrument Model 215 pH-meter (Denver Instrument Company, Denver, CO, USA).

2.3. Preparation of Polymer Modified Electrodes The polymer electrode coating was prepared by direct electrochemical polymerization-deposition, carried out in a solution containing 5 103 M BSBT and 0.1 M tetrabutylammonium hexafluorophosphate supporting electrolyte in deaerated acetonitrile solvent. The PBSBT polymer film was grown on both conventional-size and micro Pt electrodes by bulk electrolysis at a controlled potential of 1.05 V (vs. SCE) [32]. The electrolysis time was 16 s for the conventional-size electrode and 8 s for the microelectrode. After removal from the solution, rinsing with acetonitrile and then with doubly distilled water, the electrode was dipped into a 0.1 M KH2PO4 K2HPO4 buffer aqueous solution (pH 7.1) and the potential scanned from 0.20 to 1.00 V at a rate of 0.05 V s1 in order to obtain characterization of the polymer coating in the aqueous medium. The electrode was then rinsed once more with doubly distilled water and transferred into the phosphate buffer solution to perform electroanalysis.

2. Experimental
2.1. Chemicals Ascorbic acid (Aldrich, 99 % pure), tetrabutylammonium hexafluorophosphate (Fluka, puriss.), acetonitrile (Aldrich, 99.8 % pure, anhydrous, packaged under nitrogen), monobasic and dibasic potassium phosphate (Carlo Erba, RPE) were used as received. 4,4'-bis(butylsulfanyl)-2,2'-bithiophene (BSBT) was synthesized as described previously [35]. Solutions of the different analytes were prepared immediately before use with doubly distilled water. The solutions for electrochemical tests were deaerated by bubbling nitrogen for 20 min before the experiments; a continuous nitrogen flow was maintained over the sample solution during the measurements. 2.2. Apparatus and Electrodes The electrochemical experiments were carried out with an Autolab PGSTAT12 potentiostat/galvanostat (Ecochemie, Utrecht, The Netherlands) in a single-compartment, threeelectrode cell configuration comprising a 3 mm diameter Pt disk electrode (Metrohm) as the working electrode, a
Electroanalysis 2002, 14, No. 78

3. Results and Discussion


3.1. Conventional-Size Electrodes As already cited, a bare Pt electrode is unsuitable to monitor AA oxidation efficiently. In Figure 1 oxidation of AA at a bare Pt electrode is shown. The electrochemical response consists of an ill-defined anodic peak, decreasing in height by repetitive cycling of the electrode potential. This behavior supports the necessity of finding an alternative electrode system for monitoring AA. The effective electrocatalytic activity of PBSBT towards AA oxidation is shown by the i E curves reported in Figure 2, where the cyclic voltammograms of the modified electrode in phosphate buffer solution, in the absence and in the presence of AA, respectively, are shown. Two important features of the electrochemical behavior of the modified electrode in aqueous medium, in compar-

Determination of Ascorbic Acid

521 according to a classical electrocatalytic regeneration of the reactant at the electrode, results in an enhancement of the voltammetric peak, the current being limited by the rate either of the chemical reaction or of the substrate diffusion. The lack of any associated negative currents in the backward scan is a consequence of the high enough rate of the polymer-substrate reaction. However, a pure redox-mediated process is only approximately in agreement with the value of the potential at which AA oxidation occurs, which corresponds exactly with that of polaron formation. Actually, at variance with what happens with solution-soluble redox mediators, no negative shift of the potentials of the response relative to the oxidation of the catalyst in the presence of substrate is observed, which suggests clear differences between the two situations. This finding, as well as other results reported thereafter, are better accounted for by considering the polymer as something that only in part resembles an actual redox mediator, behaving rather like a particular electrode surface. As an additional peculiarity, the insulating character of the film does not allow oxidation of AA to occur at potentials lower than those of polaron formation. It is reasonable that inner-sphere electrode charge transfers exhibit specific overvoltages on a surface with chemical and electric peculiarities very different from those of conventional electrode materials. Being the oxidizing centers, i.e., the polarons on the polymer, confined onto the electrode surface, a diffusive character has to be expected under the conditions that: i) the polymer is either absolutely compact or the holes present in the porous structure are wide enough not to induce a lowering in the analyte diffusion rate inside them; ii) the reaction rate between the polaron centers and the substrate is high with respect to the time scale of the experiment. From the constancy of the values obtained for the ip, a/(qE/qt)1/2 ratio (ip, a anodic peak current; qE/qt potential sweep rate) at varying v in the range 0.020 to 0.500 V s1, both hypotheses are verified. The repeatability of the response relative to AA oxidation process on the modified electrode was tested by repeatedly scanning the potential from 0.20 to 0.80 V, in phosphate buffer solutions containing 6 103 M AA. As usual, the voltammogram recorded on the first scan is different from the subsequent ones; on going cycling, the current slowly decreases, reaching a stable value after a few cycles. It should be noted that the upper positive value reached in the potential sweep is critical, since if it is as high as 1.00 V the current peak exhibits a continuous decrease, suggesting a progressive degradation of the coating. In particular, the stability was tested by series of 10 consecutive scans; up to 20 consecutive series could be performed without observing any significant changes of the responses. These results support the possibility of performing efficient analyses for long periods of time, and of using the modified electrode in flow systems. Calibration analysis at different AA concentrations was also performed in order to test the actual capabilities of this system as an amperometric sensor. A linear dependence of the oxidation peak current on AA concentration, in the
Electroanalysis 2002, 14, No. 78

Fig. 1. Cyclic voltammograms of a 3 103 M AA (pH 7.1), 0.1M phosphate buffer, aqueous solution. 0.3 mm diameter Pt electrode; 0.050 V s1 potential scan rate. First scan and steady state curves (solid and dotted line, respectively) are reported.

Fig. 2. Cyclic voltammograms of a PBSBT/Pt modified electrode in 0.1 M phosphate buffer, aqueous solution (pH 7.1) in the absence (solid line) and in the presence (broken line) of 3 103 M AA; 0.050 V s1 potential scan rate. 2nd scan in both cases is shown.

ison with that observed in MeCN solution [32], should be emphasized: i) the peak associated to the polaron formation is broader and worse defined; ii) the bipolaron peak shifts to more anodic potentials, so that it is observable as a poorly defined shoulder of the solvent discharge current increase. These differences could be ascribed to higher difficulty for the polymer to swell in aqueous media, which is a necessary condition to permit the entrance of counterions into the polymer coating to compensate the positive charge arising during the oxidation process. In the presence of AA analyte, clear enhancement of the anodic peak current and concomitant disappearance of the associated cathodic peak are observed. This behavior is consistent with a fast process in which the polymer coating acts as a redox mediator with respect to the analyte substrate. When the potential necessary for p-doping to occur is reached, the oxidized polymer formed oxidizes AA, being reduced to the neutral state. The cyclic process,

522 range from 2 103 to 1.5 102 M (R2 0.98), was obtained, using a 0.050 V s1 potential scan rate. Tests were also performed with different analytes which are present with AA in different matrices: glucose and uric acid, which are also present in fruits and vegetables, 4acetaminophenol (paracetamol) and acetylsalicylic acid, often accompanying AA in pharmaceutical preparations. In the present article, we did not perform tests for calibrating the system for quantitative determination of these analytes. The tests reported in the following were rather addressed to the study of possible interferences and of the nature of the operative electrode process. Figure 3a reports cyclic voltammograms recorded on the modified electrode in phosphate buffer solutions, in the absence and in the presence of increasing concentrations of glucose, respectively. No significant increase of the current due to polaron formation was observed, only a poorly detectable increase of the current in the correspondence of the bipolaron formation being noted. However, as already observed, repetitive scans in this potential region cause a progressive degradation of the deposit, leading to very poor reproducibility of the analytical measures. Similar analyses

S. Lupu et al.

carried out at the same glucose concentration, though in the presence of 3 103 M AA, show a completely different electrochemical behavior: the current of the first oxidation peak progressively increases by addition of increasing amounts of glucose to the solution (see Figure 3b). This result supports the hypothesis advanced above, i.e., that the polymeric film does not actually act as a true redox mediator, capable of eliminating the charge transfer overvoltage affecting irreversible electrochemical process. Only the presence of AA that, in a classical homogeneous catalytic process, mediates the oxidation of glucose, permits the oxidation of the sugar molecule to occur according to pure thermodynamic arguments: the oxidation of the glucose substrate is anticipated and detected in the correspondence of the polaron formation peak. The current increase is not significant when the AA concentration is below a given value, suggesting that the rate of the redox reaction between oxidized AA and glucose is not particularly high. In view of the oxidation mechanism of AA, it should be considered that oxidation of glucose by dehydroascorbic acid competes with deactivation of the oxidation product of AA by hydrolysis to the corresponding dihydrate. Electrochemical tests were also carried out on the oxidation of uric acid. As shown in Figure 4, the addition of this analyte to the solution causes an increase of the signal relative to the polymer oxidation, in a way similar to that induced by the presence of AA, the redox mediation of PBSBT being effective in this case. The further addition of an equal amount of AA further increases the peak current, no resolution between the two signals, due to uric acid and to AA, respectively, being possible under this experimental condition, i.e., at 0.050 Vs1 potential scan rate. This value of the scan rate was, however, chosen in order to meet with the best conditions to observe the polaron formation response, i.e., conditions comparable with those of the other tests performed. Experiments specifically aimed to deconvolve

Fig. 3. Cyclic voltammograms of a PBSBT/Pt modified electrode in 0.1 M phosphate buffer, aqueous solution, in the absence (solid line) and in the presence of two different glucose concentrations (dotted line: 1 103 M; broken line: 7.8 103 M). a) in the absence and b) in the presence of 3 103 M AA. 0.050 V s1 potential scan rate. In all cases the 2nd scan is reported.
Electroanalysis 2002, 14, No. 78

Fig. 4. Cyclic voltammograms of 3 103 M uric acid, 0.1 M phosphate buffer aqueous solution, on a PBSBT/Pt modified electrode, in the absence (solid line) and in the presence (broken line) of 1 103 M AA. 0.050 V s1 potential scan rate. In both cases the 2nd scan is reported.

Determination of Ascorbic Acid

523 conventional-size modified electrodes. According to the results obtained for the conventional-size electrodes, the voltammetric responses for such a surface electrocatalytic mechanism is expected to posses a diffusive character. As a consequence, in agreement with the constancy of the ratio ip, a/(qE/qt)1/2 observed in the case of conventional-size electrodes at varying qE/qt, a limiting current value independent of qE/qt has to be expected for the S-shaped responses recorded on the microelectrode. Actually, the values along the plateau of the voltammetric i E curves obtained in our study with similar systems often result not so nicely constant, as clearly shown by the curves in Figure 5, to allow for a precise evaluation of the relevant current level. However, the sigmoidal-shaped i E curve at a microelectrode can be considered as the result of the convolution of the peak-shaped i E curve recorded at a conventional-size electrode with a t1/2 function that accounts for the additional contribution to diffusion due to the small dimension of the electrode. Hence, deconvolution of the forward trace of the response recorded at a microelectrode by t1/2 leads to a peak-shaped curve (see Figure 5, curve c) that retains the properties of the linear sweep response at a conventionalsize electrode [38]. In other words, it assumes the same physical meaning as the corresponding signal on a conventional-size electrode. This implies that the peak currents of responses obtained by time semidifferentiation, which is an operation coincident with deconvolution by t1/2, recorded at the same potential scan rate result proportional to the concentration of the electroactive species in the bulk of the solution. In other words, it is not just the outcome of a mathematical manipulation of the original signal. The result of application of such a procedure to the AA oxidation signal is quite satisfactory, since the peak current on the resulting signal can be much more precisely evaluated than the plateau current on the original curve. According to the above arguments, the occurrence of an electrocatalytic regeneration of the electrode reactant could be checked on the deconvoluted signals by testing the constancy of the Ip, a/(qE/qt)1/2 ratio at varying the scan rate (Ip, a (time)

the responses relative to the two different processes are beyond the scope of the present work. The conclusion is, however, that detection of both analytes is possible once resolution of the analytes is performed in advance to the analysis of each one at a time. Analogous tests were carried out for components of pharmaceutical interests, i.e., acetylsalicylic acid and paracetamol. No signal relative to the former species could be detected, even after addition of AA to the solution: analysis with increasing amount of the two analytes show that only AA causes an enhancement of the voltammetric response. On the contrary, the addition of paracetamol to the buffer solution causes a high increase of the anodic current that, however, progressively decreases by repetitive cycling, without reaching steady-state values. Poisoning of the electrode surface by the oxidation products should be invoked to explain this behavior, making the analysis of this analyte unfeasible by such an electrode system. Tests with the contemporary presence of AA led to a similar progressive lowering of the response. On the other hand, similar behavior is also obtained on a bare Pt electrode.

3.2. Microelectrodes The electrochemical oxidation of AA was also studied at PBSBT/Pt modified microelectrodes, both in the presence of usually high buffer concentration and, based on preliminary results obtained by us in low-conducting media [37], at very low buffer concentration. These latter tests were performed with the aim of testing the system as a possible amperometric sensor suitable to work in unmodified real matrices. Figure 5 compares the voltammetric responses recorded in the same aqueous solution when only containing 0.1 M phosphate buffer, and when added with 3 103 M AA, respectively. The presence of AA causes a marked enhancement of the anodic current, which suggests that the PBSBT film acts as a redox mediator, similar to the case of

Fig. 5. Cyclic voltammograms of a PBSBT/Pt modified microelectrode in 0.1 M phosphate buffer aqueous solution (pH 7.1) in the absence a) and in the presence b) of 3 103 M AA. c): deconvoluted (b) curve (forward trace). 0.05 V s1 potential scan rate. Left and right axes ordinate values refer to (a, b) and to (c) curves, respectively. In both cases the 2nd scan is reported.
Electroanalysis 2002, 14, No. 78

524 semidifferential [diffusion (t1/2) deconvoluted] peak current). Tests were carried out at different AA concentrations in the range 2 105 to 1 102 M, leading to a satisfactory linear dependence of the deconvoluted oxidation peak current on the AA concentration (R2 0.99). A detection limit coincident with the lower extreme of the linearity range was calculated, as the concentration corresponding to a signal equal to the zero intercept of the calibration line, incremented by three times the standard deviation of regression (sy/x). The AA oxidation has also been studied in solutions containing a much lower phosphate buffer concentration. The PBSBT coated microelectrode still presents an electrocatalytic effect towards AA oxidation, the value of the ratio between the peak current of the deconvoluted signal and the square root of the potential scan rate assuming constant values within the range 0.020 to 0.500 V s1. Calibration plots were drawn out also under similar experimental conditions, once more obtaining quite a good linear dependence of the deconvoluted oxidation peak current on the AA concentration, in the range 2 105 to 1 102 M (R2 0.98). Once more a detection limit coincident with the lower extreme of the linearity range was calculated.

S. Lupu et al.

6. References
[1] A. Marchesini, F. Montuori, D. Muffato, D. Mestri, J. Food Sci. 1974, 39, 568. [2] Pearsons Composition and Analysis of Food, 9th ed. (Eds.: R. Kirk and R. Sawyer), Longman, Harlow UK, 1991. [3] E. S. Wagner, B. Lindley, R. D. Coffin, J. Chromatogr. 1979, 163, 225. [4] J. W. Finley, E. Duang, J. Chromatogr. 1981, 207, 449. [5] M. Dominguez, A. Aldaz, F. Sanchez-Burgos, J. Electroanal. Chem. 1976, 68, 345. [6] M. Rueda, A. Aldaz, F. Sanchez-Burgos, Electrochim. Acta 1978, 23, 419. [7] P. Karabinas, D. Jannakoudakis, J. Electroanal. Chem. 1984, 160, 159. [8] O. Hammerich, B. Svensmark in Organic Electrochemistry. An Introduction and a Guide, 3rd ed., (Eds.: H. Lund, M. M. Baizer) Marcel Dekker, New York 1991, ch. 16. [9] C.-X. Cai, K.-H. Xue, S.-M. Xu, J. Electroanal. Chem. 2000, 486, 111. [10] N. Totir, S. Lupu, E. M. Ungureanu, M. Giubelean, A. Stefanescu, Rev. Roum. Chim., in press. [11] M. Chen, H. Li, Electroanalysis 1998, 10, 477. [12] J. Wang, T. Golden, Anal. Chim. Acta 1989, 217, 343. [13] A. P. Doherty, M. A. Stanley, J. G. Vos, Analyst 1995, 120, 2371. [14] Z. Q. Gao, K. S. Siow, A. Ng, Y. M. Zhang, Anal. Chim. Acta 1997, 343, 49. [15] Z. U. Bae, J. H. Park, S. H. Lee, H. Y. Chang, J. Electroanal. Chem. 1999, 468, 85. [16] C. R. Raj, T. Ohsaka, J. Electroanal. Chem. 2001, 496, 44. [17] V. S. Ijeri, P. V. Jaiswal, A. K. Srivastava, Anal. Chim. Acta 2001, 439, 291. [18] G. Erdogdu, A. E. Karagozler, Talanta 1997, 44, 2011. [19] H. B. Mark Jr., N. Atta, Y. L. Ma, K. L. Petticrew, H. Zimmer, Y. Shi, S. K. Lunsford, J. F. Rubinson, A. Galal, Bioelectrochem. Bioenerg. 1995, 38, 229. [20] A. Galal, Electroanalysis 1998, 10, 121. [21] A. Malinauskas, Synth.Met. 1999, 107, 75. [22] M. E. Lyons, W. Breen, J. Cassidy, J. Chem. Soc. Faraday Trans. 1991, 87, 115. [23] I. G. Casella, M. R. Guascito, Electroanalysis 1997, 9, 1381. [24] D. M. Zhou, J. J. Xu, H. Y. Chen, H. Q. Fang, Electroanalysis 1997, 9, 1185. [25] J. J. Sun, D. M. Zhou, H. Q. Fang, H. Y. Chem, Talanta 1998, 45, 851. [26] P. J. OConnel, C. Gormally, M. Pravda, G. G. Guilbault, Anal.Chim.Acta 2001, 431, 239. [27] Microelectrodes: Theory and Applications (Eds.: M. I. Mon s, and J. L. Daschbach), Nato ASI tenegro, M. A. Queiro Series, Kluwer, Dordrecht, The Netherlands 1991. [28] O. A. Sadik, G. G. Wallace, Electroanalysis 1993, 5, 555 and reference therein. [29] O. A. Sadik, G. G. Wallace, Electroanalysis 1994, 6, 860. [30] D. Iarossi, A. Mucci, L. Schenetti, R. Seeber, F. Goldoni, M. Affronte, F. Nava, Macromolecules 1999, 32, 1390. [31] H. Ding, L. Pigani, R. Seeber, C. Zanardi, J.New Mat. Electrochem. Systems 2000, 3, 337. [32] B. Ballarin, F. Costanzo, F. Mori, A. Mucci, L. Pigani, L. Schenetti, R. Seeber, D. Tonelli, C. Zanardi, Electrochim. Acta 2001, 46, 881. [33] H. Ding, Z. Pan, L. Pigani, R. Seeber, C. Zanardi, Electrochim. Acta 2001, 46, 2721. [34] H. Ding, Z. Pan, L. Pigani, R. Seeber, J.New Mat. Electrochem. Systems 2001, 4, 63.

4. Conclusions
The results obtained using PBSBT modified electrodes on a first analyte considered by us under a wide range of experimental conditions are quite promising: the electrocatalytic response relative to oxidation of AA allows quantitative determination over a wide range of concentrations, with fast time of response. In particular, the use of a microelectrode system leads to quite satisfactory quantitation even of very low AA concentration and allows the analysis to be also performed in very low buffer concentration, i.e., in very poorly conductive media, which is a condition that approximates those met with when working on real matrices. Analyses performed on other analytes present in food together with AA show a similar, fast-response, electrocatalytic behavior, which is an interesting result in view of the possible use of similar modified electrodes for detecting the different components in a flow coming out from a separation system.

5. Acknowledgements
Financial support of MURST (Rome) Ricerche di Interesse Nazionale is acknowledged. S.L. is grateful to the Romanian Ministry of Education (MEN) for a governmental Grant financially supporting a six month stay at the University of Modena and Reggio Emilia, Department of Chemistry.

Electroanalysis 2002, 14, No. 78

Determination of Ascorbic Acid [35] U. Folli, F. Goldoni, D. Iarossi, A. Mucci, L. Schenetti, J. Chem. Res. 1996, 69. [36] N. Adams, Electrochemistry at Solid Electrodes, Marcel Dekker, New York 1969.

525
[37] M. Cocchi, G. Franchini, M. Manfredini, A. Marchetti, L. Pigani, R. Seeber, L. Tassi, A. Ulrici, M. Vignali, C. Zanardi, P. Zannini, Ann. Chim. (Rome), in press. [38] M. I. Pilo, G. Sanna, R. Seeber, J. Electroanal. Chem. 1992, 323, 103.

Electroanalysis 2002, 14, No. 78

Вам также может понравиться