You are on page 1of 6

Sensors and Actuators B 192 (2014) 601–606

Contents lists available at ScienceDirect

Sensors and Actuators B: Chemical


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

Influence of cobalt content on nanostructured alpha-phase-nickel


hydroxide modified electrodes for electrocatalytic oxidation
of isoniazid
Paulo Roberto Martins, Luís Marcos Cerdeira Ferreira, Koiti Araki, Lúcio Angnes ∗
Universidade de São Paulo, Instituto de Química, Av. Professor 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: Fluorine doped tin oxide electrodes modified with nanostructured alpha nickel hydroxide, cobalt oxi-
Received 28 August 2013 hydroxide and mixed alpha nickel/cobalt hydroxide electrodes were evaluated for determining isoniazid
Received in revised form 4 November 2013 concentration by cyclic voltammetry (CV) and batch injection analyses (BIA) associated with ampero-
Accepted 8 November 2013
metric detection. The different cobalt contents showed distinct features regarding isoniazid oxidation
Available online 16 November 2013
potential and current magnitude. The modified electrodes containing 25% cobalt presented about 3- and
7.5-fold greater sensitivity as amperometric isoniazid sensors than pure Ni(OH)2 electrodes, respectively
Keywords:
for CV and BIA assays. This characteristic associated with high stability and reproducibility ensured BIA
Nickel-cobalt mixed hydroxides
Isoniazid
detection limits as low as 1.37 × 10−7 , 5.19 × 10−8 and 4.94 × 10−8 mol L−1 for Ni-100, Ni-75 and Ni-50
Amperometric BIA electrodes respectively, as well as a high analyses throughput (about 400 h−1 ). The resulting electrodes
Nanostructured electrode are promising for other applications, once presents elevated sensitivity and the additional advantage of
allow the researcher choose the best working potential, just varying the constitution of the modifier film.

© 2013 Published by Elsevier B.V.

1. Introduction frequencies makes BIA-amperometry a powerful technique for high


demand analysis and has been widely demonstrated in pharmaceu-
Pyridine-4-carboxylic acid hydrazide, or simply isoniazid, is a tical products analysis [12–15].
bacteriostatic drug widely used for treatment of tuberculosis [1]. In recent years, the search for new electroanalytical sensors
Several isoniazid products are available for tuberculosis treatment based on nanostructured materials has intensified. These materials
including tablets, syrup and intramuscular injections [2]. However, act as modifying agents of electrode surfaces, making them very
the dosage should be strictly controlled for an effective treatment useful for the determination of chemical species of environmental,
while avoiding overdose, because of isoniazid’s high hepatotox- clinical and industrial interest [16–19]. This search is due to differ-
icity. Therefore, spectrophotometric [3–5], electrochemical [6–9] entiated features that materials can present, producing modified
and chromatographic [10,11] methods have been used for isoniazid electrodes with greater sensibility, lower detection limits, as well
quantification. However, most of these techniques are laborious as, fast electron transfer kinetics [20].
and inappropriate for rapid and simple quality control analysis, Here, we report the preparation of nanostructured materials
thus justifying the efforts for the development of high speed and based on nickel hydroxide (Ni-100), cobalt hydroxide (Co-100) and
precision methods for isoniazid analysis. nickel hydroxide containing 25% (Ni-75) and 50% (Ni-50) of cobalt
Batch injection analysis (BIA) associated with amperomet- ions, stabilized into the alpha phase, and the use of these nano-
ric detection involves the insertion of a small volume (a few materials as surface modifiers of fluorine-doped tin oxide (FTO)
microlitres) of analyte onto the surface of an electrode positioned electrodes to perform electrocatalytic oxidation of isoniazid using
inside a cell filled with supporting electrolyte. The high con- cyclic voltammetric and BIA-amperometry.
vectional transport onto the electrode surface followed by quick
dispersion generates a characteristic sharp transient signal of cur-
rent. The good accuracy, low reagent consumption and high sample 2. Experimental

Analytical grade reagents were used throughout as received


without further purification. Potassium hydroxide, isopropyl alco-
∗ Corresponding author. Tel.: +55 11 3091 3828; fax: +55 11 3091 3781. hol, methanol and n-butyl alcohol were purchased from Synth
E-mail address: luangnes@iq.usp.br (L. Angnes). Chemicals. Anhydrous glycerin (Sigma–Aldrich), nickel acetate

0925-4005/$ – see front matter © 2013 Published by Elsevier B.V.


http://dx.doi.org/10.1016/j.snb.2013.11.029
602 P.R. Martins et al. / Sensors and Actuators B 192 (2014) 601–606

tetrahydrate and cobalt acetate tetrahydrate were acquired from


Vetec Chemicals. Milli-Q DI-water (resistivity ≥18.2 M cm) was
used in all experiments.

2.1. Synthesis of nanostructured materials

The Ni(OH)2 , mixed Ni1−x Cox (OH)2 , and Co(OH)2 were prepared
using a modified Tower’ method [21–23]. Each preparation started
with a mixture of nickel acetate and cobalt acetate in appropri-
ate stoichiometric proportions and always totalling 4.82 mmol.
The salts were solubilized in 25 mL glycerin, and then 9.46 mmol
of KOH in n-butanol solution was slowly added under stirring
at room temperature. In this way, sol dispersions containing
Ni(OH)2 , NiCo(OH)2 – 75:25, NiCo(OH)2 – 50:50 and Co(OH)2
nanoparticles were prepared as described previously [21,23]
and here are respectively denoted as Ni-100, Ni-75, Ni-50 and
Co-100.

2.2. Apparatus
Fig. 1. Schematic representation of the batch injection cell adapted to FTO elec-
trodes comprising (A) motorized micropipette; (B) output for electrical contact
A Higaku Miniflex powder diffractometer, equipped with a Cu wire; (C) reference electrode; (D) circular magnetic plate over cylindrical platform;
K␣ radiation source (30 kV, 15 mA, 1.541 Å and step 0.03◦ ) was used (E) auxiliary electrode; and (F) cell drain. (G) Front and (H) back views of the FTO
to register the X-ray diffractograms (2 range = 1.5–70◦ ). electrode.
Voltammetric and amperometric measurements were carried
out using a ␮Autolab Type III. The arrangement of the electrochem- 3. Results and discussion
ical cell was constituted by an arrangement of three electrodes; a
FTO working electrode modified with Ni-100, Ni-75, Ni-50 or Co- 3.1. X-ray diffractrometry and scanning electron microscopy
100, a Pt wire coiled counter electrode and a reference electrode
(Ag/AgClKCl 1 mol L−1 ). The cyclic voltammograms were recorded in The polymorphic form of nickel hydroxide and mixed
alkaline medium (1 mol L−1 KOH solution) in the −0.15 to +0.5 V nickel/cobalt hydroxide electrodes were evaluated by X-ray diffrac-
range and scan rate of 20 mV s−1 . tion, as described previously [21]. The alpha phase presents a
disordered structure along the c axis and a basal plane distance
larger than the beta one. Thus, the alpha phase presents diffrac-
2.3. Preparation of nanostructured electrodes tion patterns significantly different to the beta phase. The X-ray
diffractograms of the Ni-100, Ni-75 and Ni-50 materials were char-
A Gravograph model LS100 laser printer was used to remove acterized by the presence of three broad peaks and low intensity in
the FTO layer from the plates defining a circular active area and the 1.5 to 70◦ (2 range), whereas Co-100 presents only one peak in
a narrow wire for electrical contact. Then, the electrodes were the region situated between 1.5◦ and 15◦ (2 range). The peak (003)
cleaned with soap, rinsed copiously with DI-water and cleaned with was observed with a maximum 2 = 7.91◦ , 9.37◦ , 8.50◦ and 8.81◦ ,
isopropyl alcohol and methanol in an ultrasonic bath for 15 min. respectively, for Ni-100, Ni-75, Ni-50 and Co-100 (Fig. 2A). Using
Then, 25 ␮L of colloidal Ni-100, Ni-75, Ni-50 or Co-100 were trans- the Bragg’s equation, the following basal plane distances were
ferred to the electrode surface and spin-coated at 2500 rpm for found using the peak (0 0 3) as reference, Ni-100 (11.16 Å), Ni-75
2 min. Finally, the FTO electrodes modified with thin nanostruc- (9.45 Å), Ni-50 (10.41 Å) and Co-100 (10.01 Å). Thus, the diffraction
tured films of the mixed hydroxides were fired at 240 ◦ C for 30 min patterns observed for all nanostructured materials were consistent
to remove most organics and to improve mechanical properties and with alpha phase materials [25]. Also, the film morphology and
conductivity. size distribution were confirmed by scanning electron microscopy
(Fig. 2B). In this figure, we can observe that the surface of the
2.4. Batch injection analysis substrate is homogeneously covered by nanoparticles of mixed
hydroxides with a size distribution of 5–10 nm.
Standard solutions of isoniazid were injected directly onto the
electrode surface using a Rainin Instruments EDP Plus Ep-100 3.2. Electrochemical behaviour of nanostructured electrodes
motorized electronic micropipette [24]. An electrochemical batch
injection cell was adapted to use modified FTO plate electrodes, as The electrochemical behaviour of the nanostructured elec-
shown in Fig. 1. A circular magnet was fixed on the top-centre of a trodes based on Ni-100, Ni-75 and Ni-50 was verified by cyclic
2.0 cm high cylindrical platform, placed inside a glass electrochem- voltammetry and electrochemical impedance spectroscopy. A
ical cell and positioned underneath the motorized micropipette. wave couple attributed to the NiIII/II process was observed in the
Also, a magnetizable metallic plate was fixed on the opposite side cyclic voltammograms of the modified electrodes, but it shifted
of the FTO glass pieces exactly underneath the electrode area. In as a function of the nanostructured material composition (Fig. 3).
this way, the modified FTO electrodes were magnetically attached The anodic and cathodic waves of Ni-100 modified electrode were
to the platform and positioned in such a way that the pipette tip found respectively at Epa = +0.38 V and Epc = +0.30 V, whereas the
was at the centre of the active electrode area in a reproducible Ni-75 electrodes exhibited a similar profile but the waves were
way. A sliding contact, protected by an insulating rubber cover, found at Epa = +0.33 V and Epc = +0.27 V. Clearly the anodic wave
and connected to an insulated wire passing through the lateral was shifted to a greater extent than the cathodic wave, such that
orifice (B), was used to establish the electrical contact with the Ep decreased from 80 to 60 mV with 25% cobalt addition. This
potentiostat/galvanostat. trend was observed for the Ni-50 modified electrode where Epa and
P.R. Martins et al. / Sensors and Actuators B 192 (2014) 601–606 603

CoIII/II was not registered in the potential range, probably due to


heating at 240 ◦ C for 30 min of the modified electrode. In this way,
the Co(OH)2 is oxidized to CoOOH, as previously described [23].
The mixed hydroxide materials were stable and no changes
were observed in the voltammetric profile, indicating no prefer-
ential lixiviation of NiII or CoIII species as shown previously [21].
Furthermore, as the amount of cobalt increased, the waves were
broadened and the charge associated with the Ni(II/III) process
increased. In fact, the integrated charge under the anodic wave was
estimated as 0.07, 0.13, 0.17 and 0.02 mC, respectively, for Ni-100,
Ni-75, Ni-50 and Co-100 modified electrodes, considering that the
same amount of material was deposited on the electrodes. This can
be assigned to a decrease of the mixed hydroxides’ grain size that
is also reflected in the higher capacitive currents in the 0.35–0.45 V
range.

3.3. Determination of the isoniazid

The cyclic voltammograms of FTO electrodes modified with Ni-


100, Ni-75, Ni-50 and Co-100 in the absence (dashed line) and
presence (black lines) of increasing concentrations of isoniazid, in
alkaline medium (1 mol L−1 KOH) and scan rate of 20 mV s−1 , are
shown in Fig. 4. It is interesting to note that Ni-100 electrodes pre-
sented an unusual behaviour along the −0.15 to +0.5 V range. In this
case, an anodic wave starting at −0.1 V and reaching the maximum
current at 0.08 V was observed in the presence of isoniazid, such
that its oxidation was anticipated by ∼0.3 V relative to the NiII/III
redox process. Notice that the current intensity is a direct func-
tion of the isoniazid concentration (Fig. 4E) and there is virtually
no response at the Ni(II/III) wave (∼0.3 V). In this region, the peak
shift proportionally to the base line signal.
In contrast, the oxidation of isoniazid at the Ni-75 and Ni-
50 modified electrodes occurred at the NiII/III wave leading to an
increase of the anodic current but keeping the wave profile. Thus,
the maximum signal was observed at +0.33 V and +0.27 V, respec-
Fig. 2. (A) XRD patterns of Ni-100, Ni-75 and Ni-50 in the 1.5–70◦ (2) range and
(B) SEM image of a Ni-50 sample on a Pt-coated Si-wafer kept in a desiccator under tively (Fig. 4B and C). This behaviour is quite unexpected since a
vacuum overnight and heated at 240 ◦ C. high electrocatalytic activity associated with high conductivity usu-
ally results in rising of a sharp wave at lower potentials than Epa ,
because even a low analyte concentration is enough to promote
Epc were cathodically shifted to +0.28 and +0.24 V, respectively.
the redox process. In contrast, low electrocatalytic activity elec-
A similar behaviour was described by Zimmerman et al. [26],
trodes generally lead to current intensification at higher potentials
who attributed this behaviour to the decrease of proton diffusion
than Epa , where the concentration of the electroactive species and
resistance as a function of cobalt addition. In this way, the charge
the conductivity are maximized. For the measurements using the
transfer resistance was reported to become almost two orders of
electrode modified with Co-100 (Fig. 4 D), the oxidation wave of
magnitude lower than for pure Ni(OH)2 .
isoniazid was recorded at Epa = +0.40 V.
For purposes of comparison, the electrochemical behaviour of
Fig. 4E presents the peak current as a function of isoniazid con-
the FTO electrode modified with Co-100 was verified by cyclic
centration using the four modified electrodes where it is clearly
voltammetry (Fig. 3). The cyclic voltammogram of the Co-100
demonstrated that Ni-75 electrodes present the best sensitivity,
electrode presented only one pair of waves at Epa = +0.42 and
whereas Ni-100 electrodes exhibited the lowest ones, but also the
Epc = +0.39 V related to the CoIV/III redox process. The redox process
lowest oxidation potential. Indeed, the Ni-75 electrodes presented
sensitivity almost three times larger than Ni-100 electrodes (140
compared with 50 ␮A mmol L−1 ). The Ni-100 and Co-100 electrode
presented the lower sensitivity (50 and 89.4 ␮A mmol L−1 respec-
tively). The inclusion of cobalt atoms into ␣-Ni(OH)2 lattice results
in an increase of structural defects of the nickel hydroxide, there-
fore more active sites are exposed to isoniazid oxidation leading to
increase of the sensitivity. Nevertheless, when the cobalt percent-
age increase to 50% into ␣-Ni(OH)2 lattice the sensitivity instead of
increasing it decrease, due decrease of amount of the nickel [27].
Moreover, the electrodes modified with nanostructured mate-
rials showed a good linear relationship between anodic current
and the isoniazid concentration, in the concentration 9.9 × 10−5 to
9.5 × 10−4 mol L−1 range, giving a correlation of R2 = 0.998, 0.999,
Fig. 3. Cyclic voltammograms of Ni-100, Ni-75, Ni-50 and Co-100 in 1 mol L−1 of 0.999 and 0.999, respectively, for Ni-100, Ni-75, Ni-50 and Co-100
KOH solution and scan rate 20 mV/s. modified electrodes.
604 P.R. Martins et al. / Sensors and Actuators B 192 (2014) 601–606

Fig. 4. Cyclic voltammograms of the modified FTO electrodes with (A) Ni-100, (B) Ni-75, (C) Ni-50 and (D) Co-100 in the absence (dashed line) and presence (black line) of
increasing isoniazid concentrations (9.9 × 10−5 to 9.47 × 10−4 mol L−1 ) and (E) plot of the current peak oxidation (␮A) as a function of isoniazid concentration (mmol L−1 ).

3.4. Batch injection analysis assay concentration was found from 1.0 × 10−6 to 1.0 × 10−3 mol L−1
(inset of Fig. 5), a very wide concentration range, which can be
Batch injection analysis is based on the injection of microlitre fitted by the following equations:
volumes of analyte directly onto a detector immersed in an elec-
trolyte with a micropipette. Transient response peaks, similar to I(Ni-100) = (0.026 ± 0.001) CIsoniazid
those obtained by flow injection analysis, are observed during the + (1.3 ± 0.4) × 10−6 (R2 = 0.975);
passage of the analyte sample on the detector. To optimize the
experiments, parameters such as distance of the pipette tip to the
electrode surface, volume injected and speed of injection were eval-
uated. The optimal distance between the electrode and the pipette I(Ni-75) = (0.195 ± 0.003) CIsoniazid
tip was found to be 2 mm. The most favourable injected volume was + (3.3 ± 1.0) × 10−6 (R2 = 0.997);
80 ␮L; the best injection speed of the programmable pipette was
the fastest one (75.2 ␮L s−1 ). To choose the optimal potential for BIA
determination, hydrodynamic experiments were carried out at the
optimized conditions described above for each modified electrode, I(Ni-50) = (0.144 ± 0.003) CIsoniazid
i.e. 0.15 V for Ni-100, 0.40 V for Ni-75, 0.35 V for Ni-50 and 0.50 V + (1.7 ± 1.1) × 10−6 (R2 = 0.994);
for Co-100 modified electrodes.
A set of experiments in triplicate under the optimized condi-
tions was carried out with all four modified electrodes in the BIA
system using isoniazid standard solutions in 1.0 mol L−1 of KOH,
I(Co-100) = (0.091 ± 0.003) CIsoniazid
and the amperogram obtained using Ni-75 is presented in Fig. 5.
A good linear relationship between current signal and analyte + (2.6 ± 1.1) × 10−6 (R2 = 0.984).
P.R. Martins et al. / Sensors and Actuators B 192 (2014) 601–606 605

7.5-fold better sensibility than Ni-100 among the electrodes, using


cyclic voltammetry and BIA, respectively. Also, the limit of detec-
tion of Ni-75 electrodes was estimated as 5.19 × 10−8 mol L−1 using
the BIA with no memory effect and a high sampling frequency.

Acknowledgments

The authors are grateful to Fundação de Amparo à Pesquisa


do Estado de São Paulo, FAPESP (process 2011/21227-8), Con-
selho Nacional de Desenvolvimento Científico e Tecnológico, CNPq
(process 306504/2011-1) and Coordenação de Aperfeiçoamento de
Pessoal de Nível Superior, CAPES, for the financial support.

References

[1] P. Nagaraja, K.C.S. Murthy, H.S. Yathirajan, Talanta 43 (1996) 1075–1080.


[2] D.A. Notterman, M. Nardi, J.G. Saslow, Pediatrics 77 (1986) 850–852.
[3] A. Safavi, M. Bagheri, Spectrochim. Acta A 70 (2008) 735–739.
[4] B. Haghighi, S. Bozorgzadeh, Microchem. J. 95 (2010) 192–197.
[5] Behrooz Zargar, A. Hatamie, Spectrochim. Acta A 106 (2013) 185–189.
Fig. 5. BIA-amperometric response of Ni-75 modified FTO electrode after injec- [6] M.F. Bergamini, D.P. Santos, M.V.B. Zanoni, Bioelectrochemistry 77 (2010)
tions of increasing concentrations of isoniazid solutions (from 1.0 × 10−6 to 133–138.
1.0 × 10−3 mol L−1 ). Inset: calibration plots for the three modified electrodes: (䊉) [7] S. Shahrokhian, E. Asadian, Electrochim. Acta 55 (2010) 666–672.
Ni-100; () Ni-75, () Ni-50 and () Co-100. The applied potential for measure- [8] U.P. Azad, V. Ganesan, J. Solid State Electrochem. 16 (2012)
ments varied according to the optimal conditions for each electrode presented in 2907–2911.
the text. [9] P.R. de Oliveira, M.M. Oliveira, A.J.G. Zarbin, L.H. Marcolino, M.F. Bergamini,
Sens. Actuator B: Chem. 171 (2012) 795–802.
[10] M.Y. Khuhawar, F.M.A. Rind, J. Chromatogr. B 766 (2002)
357–363.
[11] P.F. Fang, H.L. Cai, H.D. Li, R.H. Zhu, Q.Y. Tan, W. Gao, P. Xu, Y.P. Liu, W.Y. Zhang,
The detection limits (LDs) of the isoniazid BIA assays using Y.C. Chen, F. Zhang, J. Chromatogr. B 878 (2010) 2286–2291.
[12] M.S.M. Quintino, K. Araki, H.E. Toma, L. Angnes, Electroanalysis 14 (2002)
the modified FTO electrodes were estimated from the slopes of 1629–1634.
the respective linear correlations and the standard deviation in [13] M.S.M. Quintino, L. Angnes, Talanta 62 (2004) 231–236.
the absence of analyte (LD = 3 × blank standard deviation/slope). [14] F.S. Felix, L.M.C. Ferreira, P.D. Rossini, C.L. do Lago, L. Angnes, Talanta 101 (2012)
The calculated detection limits were 1.37 × 10−7 , 5.19 × 10−8 , 220–225.
[15] L.M.C. Ferreira, F.S. Felix, L. Angnes, Electroanalysis 24 (2012) 961–966.
4.94 × 10−8 and 4.45 × 10−7 mol L−1 , respectively, for Ni-100, Ni- [16] C.M. Welch, M.E. Hyde, C.E. Banks, R.G. Compton, Anal. Sci. 21 (2005)
75, Ni-50 and Co-100 modified electrodes. These results are 1421–1430.
consistent with the higher sensitivity achieved with Ni-75 [17] B. Rezaei, S. Damiri, Electrochim. Acta 55 (2010) 1801–1808.
[18] M. Tyagi, M. Tomar, V. Gupta, Biosens. Bioelectron. 41 (2013) 110–115.
electrodes as compared with the other nanostructured elec- [19] Y.Y. Yu, Y. Yang, H. Gu, T.S. Zhou, G.Y. Shi, Biosens. Bioelectron. 41 (2013)
trodes. In fact, its sensibility was 7.5-fold higher than that 511–518.
of Ni-100, which presented the lowest detection limit and [20] H. Dong, X.D. Cao, J. Phys. Chem. C 113 (2009) 603–609.
[21] P.R. Martins, A.L.A. Parussulo, S.H. Toma, M.A. Rocha, H.E. Toma, K. Araki, J.
sensitivity. Power Sources 218 (2012) 1–4.
Repetition of alternating injections of 1.0 × 10−5 and [22] M.A. Rocha, H. Winnischofer, K. Araki, F.J. Anaissi, H.E. Toma, J. Nanosci. Nan-
1.0 × 10−6 mol L−1 isoniazid solutions (figure not shown) resulted otechnol. 11 (2011) 3985–3996.
[23] P.R. Martins, S.H. Toma, M. Nakamura, H.E. Toma, K. Araki, RSC Adv. 3 (2013)
in a very good accuracy of current signals and low relative standard 20261–20266.
deviations (RSD). The RSD values for the highest and the lowest [24] J. Wang, L. Chen, L. Angnes, B.M. Tian, Anal. Chim. Acta 267 (1992)
concentrations were, respectively, 2.54% and 2.01% for Ni-100; 171–177.
[25] A. DelahayeVidal, B. Beaudoin, N. SacEpee, K. TekaiaElhsissen, A. Audemer, M.
2.78% and 1.69% for Ni-75; 3.83% and 2.13% for Ni-50; and 3.03% Figlarz, Solid State Ionics 84 (1996) 239–248.
and 2.55% for Co-100 electrodes. These experiments also demon- [26] A.H. Zimmerman, P.K. Effa, J. Electrochem. Soc. 131 (1984) 709–713.
strated that there is no memory effect as demonstrated by the [27] W. Yan, D. Wang, G.G. Botte, Electrochim. Acta 61 (2012) 25–30.
complete recovery of the low and high concentration signals after
successive alternating injections. Furthermore, the fast increase Biographies
and decrease of isoniazid oxidation signal allows a high sample
throughput. In fact, a maximum (theoretical) sampling frequency
Paulo Roberto Martins graduated in chemistry (Universidade Federal de Santa
of about 400 h−1 can be estimated from the cleaning time of 8.9 s Maria) in 2007, concluded his PhD in inorganic chemistry at São Paulo State
(time required to attain the maximum signal and decrease to 5% of University in 2012, under guidance of Koiti Araki, working on electrochemistry, elec-
that value). troanalysis and development of amperometric sensors utilizing new materials. He
is a post-doc in Prof. Angnes Lab and works on the construction of electrodes based
on nickel and cobalt hydroxides and their application in association with flow and
4. Conclusion batch analysis.

Luís Marcos Cerdeira Ferreira graduated in chemistry (Universidade Federal de


Nanostructured electrodes based on alpha-Ni(OH)2 (Ni-100), Juiz de Fora) in 2008, moving to São Paulo University where he finished his masters
in Analytical Chemistry (2011). He is now doing his PhD under the supervision of
Co(OH)2 (Co-100) and mixed nickel and cobalt hydroxides (Ni-75 Prof. Lucio Angnes, working with electroanalytical chemistry, with emphasis on the
and Ni-50) were prepared, characterized and utilized as amper- construction of modified electrodes using metalloporphyrins, and the use of these
ometric sensors for isoniazid determination. The electrochemical electrodes in association with automated analysis.
behaviour of the FTO modified electrodes in alkaline media Koiti Araki graduated at the University of São Paulo at Ribeirão Preto (1986), moved
(1 mol L−1 KOH) was studied by CV, in which the Epa was shown to to USP–São Paulo where he did his masters (1989) and PhD (1994) in Inorganic
shift to less positive potentials as a function of the cobalt concentra- Chemistry, both under the guidance of Prof. Henrique Toma. In 1989, he started
to teach in USP São Paulo and in 2006 ascended to full professor. Koiti has expe-
tion. The four modified electrodes were evaluated for quantification rience in supramolecular chemistry, nanotechnology, focusing nanoparticles and
of isoniazid where the Ni-75 electrodes presented about 3- and their functionalization, and the development of chemical interfaces and devices.
606 P.R. Martins et al. / Sensors and Actuators B 192 (2014) 601–606

Lúcio Angnes received his B.Sc. (1978) from University of Santa Maria (RS–Brazil) of modified electrodes, searching for new ways to create arrays of microelectrodes,
then moved to University of São Paulo, at São Paulo city, Brazil, where he com- exploring enzymes naturally immobilized in vegetal tissues for analytical purposes,
pleted his M.Sc. (1981) and PhD (1987). In 1990–1992, he developed his postdoctoral the construction of microelectrodes and its application as a detector in capillary
work in Las Cruces (NM, USA) in Joseph Wang’s lab. In 1996, he rose to associated electrophoresis, and the association of new sensors with flow and batch injection
professor and in 2002 ascended to a full professor position. His research inter- analysis.
ests include the construction of electrodes with new materials, the development

You might also like