You are on page 1of 5

Journal of Electroanalytical Chemistry 775 (2016) 72–76

Contents lists available at ScienceDirect

Journal of Electroanalytical Chemistry

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

Development of a tetraphenylporphyrin cobalt (II) modified glassy


carbon electrode to monitor oxygen consumption in biological samples
L.P.H. Saravia a, S. Anandhakumar a,⁎, A.L.A. Parussulo a, T.A. Matias a, C.C. Caldeira da Silva b, A.J. Kowaltowski b,
K. Araki a, M. Bertotti a,⁎
a
Departamento de Química Fundamental, Instituto de Química, Universidade de São Paulo
b
Departamento de Bioquímica, Instituto de Química, Universidade de São Paulo, Av Prof. Lineu Prestes, 748 São Paulo, SP, Brazil

a r t i c l e i n f o a b s t r a c t

Article history: A glassy carbon (GC) electrode modified with a CoTPP (5, 10, 15, 20-meso tetraphenylporphyrin cobalt (II)) film
Received 25 February 2016 was prepared and used as a dissolved oxygen electrochemical sensor. The as-prepared electrode was character-
Received in revised form 29 April 2016 ized using atomic force microscopy (AFM) and cyclic voltammetry (CV). The electrochemical behaviour of the
Accepted 18 May 2016
CoTPP/GC modified electrode for oxygen reduction in 0.1 mol L−1 KNO3 aqueous solution was compared to
Available online 19 May 2016
that of a bare GC electrode and a decrease in overpotential (~0.30 V) and onset potential (~0.15 V) was noticed.
Keywords:
The proposed platform was used as an electrochemical sensor for dissolved oxygen detection, and analytical pa-
Dissolved oxygen sensor rameters such as reproducibility, repeatability and long-term stability were determined. The usefulness of the de-
CoTPP veloped sensor was demonstrated by continuous amperometric monitoring of mitochondrial oxygen
Amperometry consumption and the results agreed well with those obtained in parallel measurements performed using a com-
Mitochondrial respiratory cycle mercial high-resolution “Clark-type” electrode.
Electrochemical sensor © 2016 Elsevier B.V. All rights reserved.

1. Introduction extensive research has been performed to develop alternative materials


and/or catalysts with high electrocatalytic activity for the electrochem-
The development of economically viable sensors and/or devices to ical reduction of oxygen. Cobalt phthalocyanine was first reported as an
get real-time information on dissolved oxygen concentration in envi- excellent catalyst for oxygen reduction [25] and different mediators
ronmental, clinical and biological samples is one of the most active re- were then reported, such as DNA [26], antraquinone [27,28],
search areas in electroanalytical chemistry [1–3]. The well-known polymethylene blue [29], peptide nanostructure [30], cobalt complexes
commercially available Clark electrode is widely used for this purpose (cobalt (II) porphyrins, cobalt phthalocyanines etc.) [31–37], iron (II)
[4–7]. Despite its popularity, several drawbacks limit its application tetrasulfonated phthalocyanine and iron (II) tetra-(N-methyl pridyl)-
e.g. (i) the resultant current is particularly unstable and therefore regu- porphyrin multilayers [21], magnesium (II) phthalocyanine [38],
lar pre-calibrations are required, which is not convenient for continuous nickel-salen polymeric film [39], vitamin B-12 [40], ABTS-laccase [41],
monitoring, especially in biological systems; (ii) even though platinum sub-micron ZnO doped RuO2 [42], diamino-o-benzoquinone [43],
and gold are inert for surface sensitive reactions, from the commercial indigotetrasulfonate [44] and eugenol [45].
point of view the cost of these materials makes them undesirable [8,9]. To extend our research interest to dissolved oxygen sensing, we
Conventional titrations and instrumental methods based on colori- have chosen cobalt (II) tetraphenylporphyrin as a mediator attached
metric, electrochemical (including potentiometry, amperometry, and to a glassy carbon support. The kinetics of the electrode process and
voltammetry), fluorescence and chemiluminescence information have the possibility of using the proposed platform for real-time monitoring
also been used [10–18]. However, electrochemical methods are most of dissolved oxygen have been investigated in this work.
widely used due to their simplicity, sensitivity, selectivity, ease of oper-
ation and low cost [19]. The oxygen reduction reaction in solid elec- 2. Experimental methods
trodes usually depends on high overpotential because of its slow
kinetics [20,21]. Platinum-based catalysts are used mostly due to their 2.1. Reagents and apparatus
high catalytic activity, but the cost of this metal has bolstered attention
towards finding alternative, platinum-free, catalysts [22–24]. Therefore, Analytically pure chemicals were used in this work and the aqueous
solutions were prepared using 18.2 MΩ cm resistivity water (Millipore).
⁎ Corresponding authors. Nafion, potassium nitrate (KNO3), potassium chloride (KCl), HEPES (4-
E-mail addresses: anandchemist@gmail.com, mbertott@iq.usp.br (M. Bertotti). (2-hydroxyethyl)-1-piperazineethanesulfonic acid), phosphate, MgCl2,

http://dx.doi.org/10.1016/j.jelechem.2016.05.026
1572-6657/© 2016 Elsevier B.V. All rights reserved.
L.P.H. Saravia et al. / Journal of Electroanalytical Chemistry 775 (2016) 72–76 73

succinate, ADP and oligomycin were purchased from Sigma Aldrich and 2.5. Mitochondrial isolation and incubation
used as received. Cyclic voltammetry and amperometry experiments
were carried out using a PalmSens portable Potentiostat (Netherland). All experiments were approved by the local Animal Care and Use
A three electrode electrochemical cell containing a CoTPP/GC modified Committee and follow international NIH-like vertebrate research proto-
electrode (3 mm diameter), a Ag/AgCl (saturated KCl) and a platinum cols. Mitochondria were isolated from adult mouse livers using standard
foil served as working, reference and counter electrodes, respectively. differential centrifugation protocols described in [47] and kept over ice
All measurements were performed at room temperature (25 ± 1 °C). until the experiments were conducted. Oxygen consumption traces
Atomic force microscopy (AFM) images were obtained using a were obtained by incubating the isolated mitochondria in a 37 °C
PicoSPM-LE Molecular Imaging System, slightly below their resonance media composed of 125 mmol L− 1 sucrose, 65 mmol L− 1 KCl,
frequency of approximately 294 kHz in air with cantilevers operating 10 mmol L−1 HEPES (4-(2-hydroxyethyl)-1-piperazineethanesulfonic
in the intermittent-contact mode (AAC mode) to measure the topogra- acid), 2 mmol L−1 Na₂HPO₄ and 4 mmol L−1 MgCl2. The pH was adjusted
phy of the surface. to 7.2 using KOH.

2.2. Synthesis of CoTPP 3. Results and discussion

TPP (meso-tetraphenylporphyrin) (0.0594 g) was dissolved in 9 mL 3.1. Microscopic characterization of CoTPP


of DMF and heated at 145 °C in a silicone bath for 20 min. Subsequently,
CoCl2·6H2O (0.0596 g) was added into the solution and the reaction The prepared CoTPP dispersed solution was coated on a mica surface
was continued for another 20 min (145–150 °C). After the reaction, for atomic force microscopy analysis (AFM) and the obtained results are
the solution was cooled at room temperature and kept in a refrigerator given in Fig S1. From the AFM images, it is clearly seen that CoTPP par-
for 12 h. Finally, the reaction product (CoTTP) was filtered, dried and ticles have a rod-like shape with aspect ratio (~0.4:0.1 μm) of ~4 and
used for the modification of the GC surface. were uniformly coated on the surface. The edges of the rod-like struc-
tures may have more active sites which can accelerate the electron
2.3. Preparation of the CoTPP/GC electrode transfer process involving oxygen reduction on the CoTPP electrode
surface.
The GC electrode was well polished with a polishing cloth using
0.05 μm alumina powder and sonicated for 5 min. The synthesized
CoTPP powder (1 mg mL−1) was dispersed in chloroform and Nafion 3.2. Electrochemical characterization of the CoTPP/GC modified electrode
(0.1%) solution and sonicated for 10–15 min to form a homogeneous
suspension. Afterwards, 10 μL of this solution were drop casted on the Cyclic voltammograms recorded with the CoTPP/GC modified elec-
GC surface and then allowed to dry in air. trode in a 0.1 mol L− 1 KNO3 solution at different scan rates (0.01–
0.15 V s− 1) are shown in Fig. 1A. The well-defined redox couple ob-
served at around 0.2 V corresponds to the CoII/CoIII electron transfer
2.4. Preparation of standard oxygen solutions process. The oxidation and reduction peak current values increase line-
arly with the scan rate (Fig. 1B), demonstrating that this is a typical sur-
The saturated dissolved oxygen solution (1.24 × 10−3 mol L−1) [46] face-controlled electrode process. In order to calculate the surface
was prepared by passing the oxygen gas into a 0.1 mol L−1 KNO3 solu- coverage (Г) of the CoTPP film, the slope obtained from the anodic cal-
tion. Solutions with known oxygen concentrations were prepared by ibration plot (Fig.1B) was used in the equation Г = ip4RT/n2F2Aυ [39],
transferring fixed amounts of the oxygen saturated solution to the elec- where ip is the anodic peak current, R is the universal gas constant
trochemical cell completely filled with supporting electrolyte solution, (8.314 J mol−1 K), T is the absolute temperature (298 K), n is the num-
which initially contained no oxygen. Scheme 1 shows the proposed ap- ber of electrons transferred in the redox process, F is the Faraday con-
paratus, which was adapted from the literature [3]. The concentration of stant (96,485 C mol−1), A is the electrode area (0.07065 cm2) and υ is
oxygen in the electrochemical cell after each addition of the saturated the scan rate (V s−1). The calculated surface coverage for the CoTPP/
oxygen solution can be calculated as described in [3]. GC modified electrode was found to be 4.75 × 10−11 mol cm−2.

Scheme 1. Oxygen calibration system design. A: oxygen cylinder, B: oxygen saturated solution, C: valve, D: piston burette, E: counter electrode, F: working electrode, G: reference
electrode, H: solution out, I: electrochemical cell, J: magnetic stirrer.
74 L.P.H. Saravia et al. / Journal of Electroanalytical Chemistry 775 (2016) 72–76

Fig. 1. Cyclic voltammograms recorded with a CoTPP/GC modified electrode in 0.1 mol L−1 KNO3 solution at different scan rates (A); corresponding ip vs υ plot (B).

3.3. Electrocatalytic reduction of dissolved oxygen at the CoTPP/GC modified 0.2 V s− 1 in 0.1 mol L−1 KNO3 aqueous neutral solution containing
electrode 100% dissolved oxygen (Fig. S2A). The oxygen reduction peak current
increases with scan rate and the linear dependence of peak current
Typical cyclic voltammograms of dissolved oxygen in 0.1 mol L−1 with the square root of scan rate (Fig. S2B) indicates the electrode pro-
KNO3 solution (pH 7) using GC and CoTPP/GC modified electrodes at cess is mass-transport controlled. Peak current and scan rate values
the scan rate of 0.1 V s−1 are shown in Fig. 2. The oxygen reduction were also plotted in logarithmic scale, and a slope value of ~ 0.44 was
peak is located at a potential of − 0.72 V for the bare GC electrode, obtained. This value is close to that expected (0.5) for a process con-
whereas the same process takes place at − 0.42 V for the CoTPP/GC trolled by diffusion.
modified electrode which is comparable to reported values for various Further, in order to get information on the mechanism of oxygen re-
modified electrodes (Table S1). By comparing both voltammograms, duction at the CoTPP/GC modified electrode, i.e. whether oxygen is re-
the overpotential for the electroreduction of oxygen can be clearly duced through a two electron pathway (formation of hydrogen
seen to be shifted towards a less negative potential value (about peroxide) or a direct four electron pathway (formation of water), the
0.30 V) when the experiment was carried out with the modified elec- transfer coefficient (α) and the number of electrons involved in the
trode. Furthermore, it is also interesting to compare onset potential rate determining step (nc) were calculated from the peak potential
values (the less negative potential at which a reaction product is (Epc) versus log scan rate plot (Fig. S2C). From the slope value
formed, i.e., the raising part of a cyclic voltammogram) of bare GC and (~ 0.054 V) of this plot, the Tafel slope (b) was calculated (from the
CoTPP/GC electrodes, which are −0.30 V and −0.15 V for the GC and equation Ep = b/2 log υ + constant) as 0.108 V. The values of α and nc
the CoTPP/GC electrodes, respectively. Therefore, in addition to the de- were calculated using the Tafel slope (from the equation y = 2.303RT/
crease in the overpotential, the onset potential was also significantly αncF, where y = Tafel slope and the other parameters have standard
shifted towards a less negative potential value. The results obtained meaning) [48,49] and were found to be 0.54 and 1.09, respectively.
confirm the electrocatalytic behaviour of the CoTPP/GC. This can be rea- Afterwards, the total number of electrons (n) involved in the oxygen
soned as (i) dioxygen binding to cobalt(II) axial position facilitates the reduction reaction was calculated from the slope value (120.6 μA)
bielectronic reduction reaction; (ii) coordination to Co(II)TPP changes obtained in Fig. S2B using the equation for a diffusion controlled
dioxygen energy levels shifting the redox potential to less negative po- irreversible process ip = 2.99 × 105ACo⁎D1/2 o υ
1/2
n (αnc)1/2 [49] where
tentials and (iii) the electrocatalytic activity of cobalt center is modulat- ip = peak current (A), A = electrode area (0.07065 cm2), Co⁎ = concentra-
ed by the porphyrin ring [34]. tion of dissolved oxygen (saturated solution = 1.24 × 10−6 mol cm−3),
To understand the kinetics of the overall electron transfer process in- D o = diffusion coefficient of oxygen in aqueous medium
volving the oxygen reduction at the CoTPP/GC modified electrode, cyclic (1.9 × 10−5 cm2 s− 1) [21], υ = scan rate (V s− 1), n = total number
voltammograms were recorded at various scan rates from 0.02 to of electrons, α = transfer coefficient (0.54), and nc = number of elec-
trons involved in the rate determining step). The n value was found to
be ~1.44, which suggests that the oxygen reduction follows a two elec-
tron pathway, where hydrogen peroxide is the main product and the
transfer of one electron is the rate determining step at the CoTPP/GC
modified electrode.

3.4. Analytical applications

The analytical performance of the CoTPP/GC modified electrode was


evaluated by varying the concentration of dissolved oxygen in the elec-
trochemical cell according to a procedure proposed in the literature and
shown in Scheme 1 [3]. Cyclic voltammetric responses in 0.1 mol L−1
KNO3 solution are presented in Fig. 3A and the resulting current mea-
sured at −0.42 V increases linearly as the dissolved oxygen concentra-
tion is increased in the range of 5 to 48.7% (62 to 604 × 10−6 mol L−1,
correlation coefficient of 0.9920, Fig. 3B). From the slope value of the
calibration plot, the limit of detection (LOD = 3Sb/S, where, Sb = stan-
Fig. 2. Cyclic voltammograms of dissolved oxygen (saturated) in a 0.1 mol L−1 KNO3
solution recorded with a GC (b) and a CoTPP/GC modified electrode (c). Curve (a)
dard deviation of the blank; S = slope of the analytical curve) [38] was
corresponds to the voltammogram of supporting electrolyte in absence of oxygen calculated and found to be 0.035% (0.4 × 10−6 mol L−1), which is rea-
recorded with a GC electrode. Scan rate: 0.1 V s−1. sonably comparable to some reported values (Table S1). The other
L.P.H. Saravia et al. / Journal of Electroanalytical Chemistry 775 (2016) 72–76 75

Fig. 3. Cyclic voltammograms recorded at various concentrations of dissolved oxygen (5–48.7%) in 0.1 mol L−1 KNO3 solution using the CoTPP/GC modified electrode (A). Corresponding
calibration plot of current (measured at −0.42 V) versus oxygen concentration (B). Scan rate = 0.1 V s−1.

important analytical parameters such as reproducibility, repeatability rates detected by both systems. ADP (c, c′) further increased this con-
and long-term stability were evaluated by cyclic voltammetry. The re- sumption by activating mitochondrial oxidative phosphorylation [47].
producibility studies were carried out using five different CoTPP/GC Finally, the addition of oligomycin blunted the ADP-induced increase
modified electrodes and the repeatability measurements were per- in oxygen consumption due to its inhibitory effect on mitochondrial ox-
formed with one modified electrode at different interval times (n = idative phosphorylation (d, d′). All measurements were very similar
10). Relative standard deviation (RSD) values were found to be 2.6 using both electrodes.
and 2.7%, respectively. Long-term stability was assessed by measuring The similarity in detection speed and range to a high-resolution
the oxygen response for five consecutive days using the same modified commercial electrode confirms the usefulness of the proposed modified
electrode and the calculated RSD was found to be 2.6% (between mea- electrode as an amperometric sensor for oxygen in biological samples.
surements, the electrode was stored at ambient temperature).
In order to evaluate the practical application of the proposed sensing 4. Conclusions
platform for continuous monitoring of dissolved oxygen in aqueous
neutral solutions, amperometric measurements at an applied potential We have successfully prepared a simple and effective dissolved oxy-
of −0.5 V were performed. The experiment was designed to vary oxy- gen sensor based on deposition of a CoTPP film onto a GC electrode. Ki-
gen concentration by bubbling oxygen and argon repeatedly. Fig. 4 netic studies were carried out systematically and the results suggested
shows the result of such experiment and the current increase (during the electroreduction of oxygen at the CoTPP/GC modified electrode is
O2 bubbling) and further decrease (during argon bubbling) clearly dem- a 2-electron process, i.e. hydrogen peroxide is the main product. The
onstrate that the sensor responds rapidly to changes in the dissolved ox- usefulness of the proposed sensor for real time oxygen monitoring
ygen concentration. was demonstrated by continuous measurements of the concentration
Finally, the possibility of using the proposed sensor to follow respira- of the target analyte in aqueous solution as well as in a biological
tory activity in a biological system was evaluated by using this electrode samples.
to monitor oxygen consumption in isolated liver mitochondria. The re- Further studies are under progress to use such a sensor as a tip in
sponse towards oxygen was continuously measured using the proposed SECM (Scanning Electrochemical Microscopy), a technique with several
electrochemical sensor (Fig. 5, blue trace) and parallel measurements advantages for studying live cells, cellular processes and cell response
were performed with a high-resolution Clark electrode Oroboros sys- towards drug administration. Localized microelectrochemical informa-
tem (Fig. 5, red trace), an equipment considered the gold standard sys- tion on oxygen uptake by a single cell can be obtained by scanning the
tem for oxygen consumption detection in liquid biological systems [50].
A slow decrease in oxygen concentrations was detected by both systems
in the presence of mitochondria (additions a, a′). The subsequent addi-
tion of succinate (b, b′), which acts as a substrate for mitochondrial elec-
tron transport to oxygen, substantially increased oxygen consumption

Fig. 5. Amperometric response of the CoTPP/GC modified electrode (E = −0.5 V) and


oxygen concentration measured with a Clark electrode in cellular medium upon
Fig. 4. Amperometric response measured with the CoTPP/GC modified electrode in addition of mitochondria (a, a′, ~0.5 mg mL−1), succinate (b, b′ 1 mmol. L−1), ADP (c, c′,
0.1 mol L−1 KNO3 solution upon bubbling of oxygen and argon. E = −0.5 V. 1 mmol L−1) and oligomycin (d, d′, 1 μg mL−1).
76 L.P.H. Saravia et al. / Journal of Electroanalytical Chemistry 775 (2016) 72–76

tip in close proximity across the cell and this information can be used as [21] J.C. Duarte, R.C.S. Luz, F.S. Damos, A.A. Tanaka, L.T. Kubota, Anal. Chim. Acta 612
(2008) 29–36.
an indicator of respiratory activity. [22] A.S.H. Smith, A. Hamill, M. Campbell, M. Uttamlal, Analyst 124 (1999) 1463–1466.
[23] L. Hutton, M.E. Newton, P.R. Unwin, J.V. Macpherson, Anal. Chem. 81 (2009)
Acknowledgements 1023–1032.
[24] S.C. Perry, G. Denuault, Phys. Chem. Chem. Phys. 17 (2015) 30005–30012.
[25] R. Jasinski, Nature 201 (1964) 1212–1213.
The authors thank FAPESP grants 13/7937-8, 10/51906-1, 12/20676-6, [26] F. Wang, J. Zhao, Y. Xu, S. Hu, Bioelectrochemistry 69 (2006) 148–157.
CNPq grant 471236/2013-6 and CAPES for financial support and SAK [27] I. Tiwari, M. Gupta, R. Prakash, C.E. Banks, Anal. Methods 6 (2014) 8793–8801.
[28] I. Tiwari, M. Singh, M. Gupta, J.P. Metters, C.E. Banks, Anal. Methods 7 (2015)
gratefully acknowledges the Sao Paulo Research Foundation (FAPESP) 2020–2027.
for the Postdoctoral Fellowship grant [14/15215-5]. [29] X. Xiao, B. Zhou, L. Tan, H. Tang, Y. Zhang, Q. Xie, S. Yao, Electrochim. Acta 56 (2011)
10055–10063.
[30] C.P. Sousa, M.D. Coutinho-Neto, M.S. Liberato, L.T. Kubota, W.A. Alves, J. Phys. Chem.
Appendix A. Supplementary data
C 119 (2015) 1038–1046.
[31] H.H. De Paz, C. Medard, M. Morin, J. Electroanal. Chem. 648 (2010) 163–168.
Supplementary data to this article can be found online at http://dx. [32] T.D. Chung, F.C. Anson, J. Electroanal. Chem. 508 (2001) 115–122.
doi.org/10.1016/j.jelechem.2016.05.026. [33] Y. Zhang, S. Liu, L. Wang, W. Liu, R. Sun, J. Ye, Mater. Lett. 144 (2015) 5–8.
[34] W. Yin, C. Chen, H. Fa, L. Zhang, J. Solid State Electrochem. 17 (2013) 3095–3099.
[35] E.S. Ribeiro, S.L.P. Dias, Y. Gushikem, L.T. Kubota, Electrochim. Acta 49 (2004)
References 829–834.
[36] A. Rahim, L.S.S. Santos, S.B.A. Barros, L.T. Kubota, Y. Gushikem, Sensors Actuators B
[1] M.L. Hichman, The Measurement of Dissolved Oxygen, John Wiley, New York, 1978. Chem. 177 (2013) 231–238.
[2] R. Ramamoorthy, P.K. Dutta, S.A. Akbar, J. Mater. Sci. 38 (2003) 4271–4282. [37] R.C.S. Luz, F.S. Damos, A.A. Tanaka, L.T. Kubota, Sensors Actuators B Chem. 114
[3] M. Sosna, G. Denuault, R.W. Pascal, R.D. Prien, M. Mowlem, Sensors Actuators B (2006) 1019–1027.
Chem. 123 (2007) 344–351. [38] L.S.S. Santos, R. Landers, Y. Gushikem, Talanta 85 (2011) 1213–1216.
[4] L.C. Clark Jr., R. Wolk, D. Granger, Z. Talor, J. Appl. Physiol. 6 (1953) 189–193. [39] C.S. Martin, T.R.L. Dadamos, M.F.S. Teixeira, Sensors Actuators B Chem. 175 (2012)
[5] L.C. Clark Jr., F. Gollan, V.B. Gupta, Science 111 (1950) 85–87. 111–117.
[6] H. Suzuki, A. Sugama, N. Kojima, Sensors Actuators B Chem. 10 (1993) 91–98. [40] M.S. Lin, H.J. Leu, C.H. Lai, Anal. Chim. Acta 561 (2006) 164–170.
[7] M. Koudelka, Sensors Actuators 9 (1986) 249–258. [41] I. Zawisza, J. Rogalski, M. Opallo, J. Electroanal. Chem. 588 (2006) 244–252.
[8] L. Nei, R.G. Compton, Sensors Actuators B Chem. 30 (1996) 83–87. [42] S. Zhuiykov, E. Kats, Sensors Actuators B Chem. 187 (2013) 12–19.
[9] L. Xiong, R.G. Compton, Int. J. Electrochem. Sci. 9 (2014) 7152–7181. [43] F. Jalali, A.M. Ashrafi, D. Nematollahi, Electroanalysis 21 (2009) 201–205.
[10] I. Helm, L. Jalukse, M. Vilbaste, I. Leito, Anal. Chim. Acta 648 (2009) 167–173. [44] T.H. Tsai, S.H. Wang, S.M. Chen, J. Electroanal. Chem. 659 (2011) 69–75.
[11] W.W. Broenkow, J.D. Cline, Limnol. Oceanogr. 14 (1969) 450–454. [45] D.W. Paul, I. Prajapati, M.L. Reed, Sensors Actuators B 183 (2013) 129–135.
[12] R.M. Manez, J. Soto, J.L. Sabater, E.G. Breijo, L. Gil, J. Ibanez, I. Alcaina, S. Alvarez, Sen- [46] A. Sukeri, L.P.H. Saravia, M. Bertotti, Phys. Chem. Chem. Phys. 17 (2015)
sors Actuators B Chem. 101 (2004) 295–301. 28510–28514.
[13] P. Fisicaro, A. Adriaens, E. Ferrara, E. Prenesti, Anal. Chim. Acta 597 (2007) 75–81. [47] E.B. Tahara, F.D.T. Navarete, A.J. Kowaltowski, Free Radic. Biol. Med. 46 (2009)
[14] W. Glasspool, J. Atkinson, Sensors Actuators B Chem. 48 (1998) 308–317. 1283–1297.
[15] Y.Y. Lau, T. Abe, A.G. Ewing, Anal. Chem. 64 (1992) 1702–1705. [48] R. Nissim, R.G. Compton, Phys. Chem. Chem. Phys. 15 (2013) 11918–11925.
[16] Y. Koo, Y. Cao, R. Kopelman, S.M. Koo, M. Brasuel, M.A. Philbert, Anal. Chem. 76 [49] R. Rajaram, S. Anandhakumar, J. Mathiyarasu, J. Electroanal. Chem. 746 (2015)
(2004) 2498–2505. 75–81.
[17] A.K. McEvoy, C.M. McDonagh, B.D. MacCraith, Analyst 121 (1996) 785–788. [50] E. Gnaiger, R. Steinlechner-Maran, G. Méndez, T. Eberl, R. Margreiter, J. Bioenerg.
[18] K. Fujimori, N. Takenaka, H. Bandow, Y. Maeda, Anal. Sci. 17 (2001) 161–166. Biomembr. 6 (1995) 583–596.
[19] A.J. Bard, L.R. Faulkner, Electrochemical methods: Fundamentals and Applications,
second ed. Wiley, New York, 1980.
[20] D.T. Sawyer, L.V. Interrante, J. Electroanal. Chem. 2 (1961) 310–327.

You might also like