Next Article in Journal
Colorimetric Chemosensor Array for Determination of Halides
Previous Article in Journal
Gold Nanoframe Array Electrode for Straightforward Detection of Hydrogen Peroxide
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Pb(II) Determination in a Single Drop Using a Modified Screen-Printed Electrode

1
Department of Chemistry and Applied Biosciences, ETH Zurich, CH-8093 Zurich, Switzerland
2
Faculty of Chemistry and Chemical Engineering, University of Maribor, Smetanova Ulica 17, 2000 Maribor, Slovenia
*
Author to whom correspondence should be addressed.
Chemosensors 2021, 9(2), 38; https://doi.org/10.3390/chemosensors9020038
Submission received: 15 January 2021 / Revised: 13 February 2021 / Accepted: 15 February 2021 / Published: 18 February 2021
(This article belongs to the Section Electrochemical Devices and Sensors)

Abstract

:
This work presents the development and validation of an electroanalytical method for Pb(II) determination in a single drop. The electrochemical sensors used were an unmodified screen-printed electrode (SPE) and a Bi-film SPE (BiFSPE). Anodic square wave stripping voltammetry (SWASV) was performed at an accumulation potential of −1.5 V vs. Ag/AgCl and an accumulation time of 60 s. Electroanalysis with an unmodified SPE did not yield satisfactory results, whereas the BiFSPE was a much better analysis method. The linear concentration using the BiFSPE was in the range of 138.8–162.5 µg/L. The accuracy and precision were evaluated for different spiked concentrations, but the method using the unmodified SPE was neither accurate nor precise. Using the BiFSPE, the method was found to be both accurate and precise for Pb(II) determination at a concentration of 140.0 μg/L, with recovery and relative standard deviation (RSD) of 106.6% and 12.1%, respectively. In addition, using the BiFSPE, LOD and LOQ values of 1.2 μg/L and 3.3 μg/L were obtained, respectively. The possible interference effect on Pb(II) stripping signal was checked in the presence of Cd(II), Zn(II), Cu(II), Sn(IV), Sb(III), Hg(II), Fe(III), As(V), K(I), I, Ca(II), and NO3. Electrochemical impedance spectroscopy measurements were also performed for the unmodified SPE and BiFSPE. The application of single drop Pb(II) analysis was tested by real water sample analysis.

1. Introduction

If only a small sample volume is available (for example, one drop, i.e., a volume of 60 µL), electrochemical analysis becomes essential. The latter became possible with the development of screen-printing technology. These screen-printed electrodes (SPEs) are promising for on-site analysis [1]. However, the analysis of a single drop has not yet been widely reported.
Few reports have been published on the determination of analytes in a sample drop on SPEs. For example, Knežević et al. [2] constructed a novel sensor based on gold nanoparticles decorated on manganese dioxide, which was used for histamine determination in a single sample drop. The method proved to be accurate, precise, selective, fast, and reproducible. In another study, norepinephrine was detected in a sample drop using a smartphone-based square-wave voltammetry (SWV) system with a graphene SPE [3]. Chiu et al. [4] reported the simultaneous determination of chloride, bromide, and iodide in aqueous solutions in a 50-µL solution. It was also shown that heavy metals can be determined in one drop of solution using a SPE [5,6].
Conventionally, for the determination of heavy metals atomic absorption spectroscopy (AAS), inductively coupled plasma mass spectrometry (ICP-MS) and inductively coupled plasma optical emission spectrometry (ICP-OES) are employed [7,8,9,10,11,12,13]. On the other hand, electrochemical determination of heavy metals using anodic or adsorptive stripping voltammetry have also been frequently employed [7,8,14,15]. Electroanalysis has certain advantages over conventional analytical methods, including low detection limits (LOD), high sensitivity and selectivity, relatively simple and inexpensive equipment, a short analysis time, and the possibility of on-site analysis [16,17]. Anodic stripping voltammetry is one of the most effective methods due to a preconcentration step that allows the trace analysis of heavy metals with high accuracy, precision, and sensitivity [1,18]. In 2000, Wang et al. [19] introduced bismuth electrodes. Bismuth has very low toxicity, a wide operational potential window, and favourable electroanalytical properties, and is insensitive to dissolved oxygen in the sample solution [20,21,22,23,24]. Bismuth-film electrodes can be prepared in situ and ex situ, and both give good electroanalytical performances in heavy metal determination [7,8,20,24,25,26,27,28]. The modification of the working electrode can be essential to ensure the desired analytical performance [8,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44].
The main routes of lead exposure are via the inhalation of contaminated aerosols or dust particles and via the ingestion of contaminated food and water. In the human body, most lead accumulates in the kidneys, followed by the liver and other soft tissues such as the heart and brain. Most of it is also accumulated in the skeleton. Lead also has a major effect on the nervous system; early symptoms thereof include headaches, a short attention span, irritability, and memory loss. Lead induces cell damage through the formation of reactive oxygen species. Lead compounds can inhibit DNA synthesis and repair, cause oxidative damage, and interact with various proteins. Studies indicate that lead is carcinogenic and induces renal tumours in rats and mice and is therefore considered a probable human carcinogen [45,46]. On that basis, lead determination was and will be a fundamental analytical problem [18,30,47,48,49,50,51]. To ensure drinking-water quality, besides the already-established procedures, new methods are being developed to remove lead ions and other metal ions from water, such as adsorption by functionalized membranes or polymeric nanocomposites [52].
The aim of this work is the development and validation of an electroanalytical method for Pb(II) determination in a sample drop using an SPE. The latter was performed with a graphite-based working electrode printed on an SPE holder and modified in situ with Bi-film. On this basis, the analytical performance of an unmodified SPE and a Bi-film-modified SPE (BiFSPE) was validated as a unique analysis of a very small amount (one drop) of sample solution using only a handheld potentiostat/galvanostat, a connecting cable, and a small SPE sensor.

2. Experimental

All electrochemical experiments were performed with a PalmSens 3 potentiostat/galvanostat controlled by PSTrace 5.6 software (PalmSens, Houten, Netherlands). An SPE (type AC1.W4.R2, BVT Technologies, Brno, Czech Republic) was used as a three-electrode system. The working electrode was made of graphite and had a diameter of 1 mm. The area of the counter electrode was 4 mm2. The reference electrode was made of silver and coated with AgCl. All potentials reported in this work referred to this Ag/AgCl reference electrode. After each measurement set, the SPE was thoroughly washed with ultrapure water (obtained by means of the Milli-Q system, Millipore Corporation, MA, USA; resistance 18.2 MΩ cm). All measurements were performed at 23 °C.
The solutions of standards were diluted with ultrapure water. For cyclic voltammetry (CV) analyses, a 1.0-mM solution of potassium hexacyanoferrate, K3[Fe(CN)6] (99%, Sigma-Aldrich, St. Louis, MO, USA) in 0.1 M KCl (Carlo Erba Reagents, Milan, Italy) was used.
For the square-wave anodic stripping voltammetry (SWASV) analyses, a 0.1 M acetate buffer was used as the supporting electrolyte with or without 0.5 mg/L Bi(III). The buffer was prepared using sodium acetate trihydrate (≥ 99%, Fisher Scientific, Loughborough, UK) and acetic acid (≥ 99.8%, Sigma-Aldrich, St. Louis, MO, USA). The Bi(III), Zn(II), Cd(II), As(V), Sn(IV), Sb(III), Cu(II), and Hg(II) standard solutions (1000.0 mg/L) were supplied by Merck KGaA (Darmstadt, Germany). The Pb(II) standard solution (1000.0 mg/L) was supplied by Merck KGaA and used for the analysis of metal traces. For the electroanalysis, this solution was diluted as required in the supporting electrolyte (0.1 M acetate buffer) with or without 0.5 mg/L Bi(III). The Fe(III) and KNO3 were supplied by Sigma-Aldrich (St. Louis, MO, USA). The KI was supplied by Carlo Erba Reagents (Val de Reuil, France). The CaCl2 was supplied by Acros Organics (Morris, NJ, USA).

2.1. Real Sample Analysis

The real sample analysis was performed using BiFSPE by preparing 0.1 M acetate buffer with tap water from our laboratory. The tap water was first allowed to run for 5 min before being used for solution preparation. All experimental conditions were the same as those used for the method validation.

2.2. CV of [Fe(CN)6]3–/[Fe(CN)6]4– Redox Couple

Before the SWASV measurement, each SPE was checked by CV measurement to determine if the sensor was functioning properly. A 60-μL drop of 1.0-mM potassium hexacyanoferrate solution in 0.1 M KCl was applied to the SPE, so that all three electrodes were covered by the drop. The CV scan started at 0.8 V and scanned in the cathodic direction to −0.3 V (switching potential) and then switched to 0.8 V, at a linear scan rate. Different scan rates were used, i.e., 10.00, 20.00, 50.00, 75.00, 125.00, 150.00, and 200.00 mV/s. The results obtained had to follow the characteristics of the reversible and diffusion-controlled process [53] in order to accept the SPE electrode as suitable. In cases of large deviations from the expected results (linearity of the current peak height vs. scan rate, the change in peak potential vs. scan rate, the ratio of the anodic and cathodic peak heights, and the expected potential difference for anodic and cathodic peaks, i.e., 59 mV [53]), the sensor was discarded and another one was used. After this test, the sensor was thoroughly rinsed with ultrapure water and dried under the stream of Ar gas.

2.3. SWASV Measurements

The analyses were performed with unmodified and modified SPEs. In this work, the term unmodified SPE refers to an SPE that has a graphite working electrode. A BiFSPE, on the other hand, refers to a Bi-film SPE where the working electrode was modified in situ during the preconcentration step with a thin bismuth film. The supporting electrolyte in this case was a 0.1-M acetate buffer containing 0.5 mg/L Bi(III).
First, a 60-μL drop of supporting electrolyte (without Pb(II)) was pipetted onto the SPE and the voltammogram of the solution without analytes (blank solution) was measured. The multiple standard addition method was used for quantification; each addition was 10 μL of a solution with Pb(II). After adding the 10-µL solution to the drop, the liquid on the sensor was drawn up into the pipette and dispensed again to form a drop on the sensor (several times) to ensure homogeneity of the solution. SWASV analyses were performed using a conditioning potential of 0.3 V, a conditioning time of 30 s, an accumulation potential (Eacc) of −1.5 V, an accumulation time (tacc) of 60 s, a start potential of −1.5 V, an end potential of 0.3 V, a potential step of 0.004 V, an amplitude of 0.05 V, a frequency of 25.0 Hz, a standby potential of 0.3 V, and a standby time of 60 s.
It should be noted that there are some limitations to the method using a single drop of solution. First, a maximum of nine standard additions of 10-μL drops could be added to the initial drop (the final volume of the drop was 150 μL), as larger volumes would result in the drop being spilled over the SPE holder. Secondly, there was a limit with the potential window: the oxidation of water takes place at more positive potentials (from about 0.8 V) and gaseous oxygen is produced. Furthermore, hydrogen starts to develop at more negative potentials (starting from approximately −1.4 V).

2.4. Electrochemical Impedance Spectroscopy Measurements

Electrochemical impedance spectroscopy (EIS) measurements were performed at −0.7 V using the sinusoidal perturbation signal with an amplitude of 10 mV (peak-to-peak). These measurements were performed in the frequency range from 100 kHz to 50 mHz with 5 measurement points per decade. Prior to the EIS measurements, the same preconcentration step as for SWASV was performed. These measurements were performed with a 60-µL drop.

3. Results and Discussion

3.1. LOD and LOQ Determination

The LOD and limit of quantification (LOQ) were determined by the determination of the signal-to-noise ratio (S/N). For the LOD, this ratio had to be equal to or greater than 3 (but less than 10), and for the LOQ, this ratio had to be equal to or greater than 10 (but close to 10). The signal S is the peak height, i.e., ΔiPb(II), and N is the background noise (determined as the difference between the highest and the lowest current measured for the background contribution). First, a volume consisting of a 60-μL drop of supporting electrolyte (blank) was dropped on the SPE and the SWASV measurement was performed, followed by 10-μL additions of the solution with a known Pb(II) mass concentration (γPb(II)) added to this drop, and voltammograms were measured until S/N ratios ≥ 3 (but < 10, for LOD) and S/N ≥ 10 (for LOQ) were obtained. This procedure was repeated at least three times.
Using the unmodified SPE, the baseline of an initial drop (blank solution) at the potentials for the Pb(II) stripping peak was intense and had a peak-like shape. We checked the ultrapure water by ICP-MS for possible contamination, but no Pb(II) was detected in the blank solution. Due to this intensive background contribution, the determination of the LOD and LOQ was less reliable, since the resulting S/N ratios were much more significant than 10 when clear peaks were developed with the addition of a solution containing Pb(II) to the initial drop (despite the subtraction of the background contribution). On the other hand, using the BiFSPE, the LOD and LOQ were determined to be 1.2 μg/L and 3.3 μg/L, respectively.

3.2. Linear Concentration Range Determination

The linear concentration range was determined by successively adding 10 µL of solution of the diluted Pb(II) standard (as needed) to the drop previously placed on the SPE sensor. Several repeated linearity determinations were performed with the same SPE and using additional new SPEs (several repeat measurements were also performed with these new SPEs). The same procedure was performed for the BiFSPE. Linearity was accepted when the square of the correlation coefficient, R2, was greater than 0.980.
The voltammograms measured using the unmodified SPE are shown in Figure 1a, the linear calibration curve is shown in Figure 1b, and the change in the peak potentials (EPb(II)) vs. γPb(II) is shown in Figure 1c. The same is also shown for the BiFSPE in Figure 2. Figure 1a and Figure 2a show a slight shift in the Pb(II) stripping peak potential. The reason is most likely that Pb can form an intermetallic compound (i.e., PbBi) or Pb is adsorbed on the Bi surface. On the one hand, the energy requirements to desorb/strip Pb from a carbon surface compared to the Bi surface/compound must be very different.
The linear concentration ranges obtained for the SPE and BiFSPE were 171.4–225.0 µg/L and 138.8–162.5 µg/L, respectively. It should be noted that repeat measurements with different SPEs had different linear concentration ranges (depending on each SPE used). However, the linear concentration ranges given here for the SPE and BiFSPE are the most representative ranges among the repeat measurements (all measurements covered the given linear concentration range). The lower limit of the linear concentration range of Pb(II) was at lower γPb(II) when the BiFSPE was used compared to the unmodified SPE, while the upper limit of the linear concentration range was at higher γPb(II) for the unmodified SPE. The sensitivity of the BiFSPE was higher compared to the unmodified SPE (the sensitivity was evaluated based on the slope of the calibration curve, Figure 1b vs. Figure 2b). In both cases, using the unmodified SPE and the BiFSPE, the Pb(II) stripping peak potential shifted to more positive potentials as γPb(II) increased (Figure 1c and Figure 2c).

3.3. Determination of Accuracy and Precision

For a single drop of the test solution, the accuracy of both methods using the unmodified SPE and BiFSPE was evaluated by determining the Pb(II) concentration of the spiked 0.1 M acetate buffer solution. Quantification was performed using the multiple standard addition method (the results for the BiFSPE are shown in Figure 3). Based on the γPb(II) obtained, the average recovery (Re) and relative standard deviation (RSD) were calculated. The method was considered accurate when the average Re was between 80.0% and 120.0%. Furthermore, the method was considered precise if the RSD was equal to or lower than 20.0% [54].
The accuracy and precision of γPb(II) determination were tested at a Pb(II) concentration of 140.0 μg/L. The method with the unmodified SPE did not yield satisfactory accuracy and precision, i.e., the Re was not within the 80.0–120.0% interval and the RSD was higher than 20.0%. On the other hand, the method with the BiFSPE reported an average Re of 106.6% and thus accurate results. In addition, the RSD value when using the BiFSPE was 12.1%. On this basis, the method with the BiFSPE for the determination of Pb(II) can be considered accurate and precise. Therefore, the addition of Bi(III) to form Bi-film in situ on the SPE is necessary to obtain both accurate and precise results. Figure 3 shows that each individual measurement set produces differences in slopes and y-intercepts between successive measurements. Therefore, the multiple standard addition methodology is required to determine Pb(II), while the calibration curve method can be problematic [46].

3.4. EIS Measurements

EIS measurements were performed at E = −0.7 V vs. Ag/AgCl, which is the potential at which the Pb(II) stripping peak begins to develop while sweeping towards more positive potentials (Figure 1 and Figure 2). The semicircles on the Nyquist plots (Figure 4) show that the real part of the impedance (Zreal) is at lower values for the BiFSPE compared to the unmodified SPE, indicating a lower charge transfer resistance (and also polarisation resistance) for the BiFSPE. Polarisation resistance is a measure of the resistance that the electrode presents to the transfer of electrons to or from the electrode to the electroactive species in solution. The lower it is, the more favourable the electron transfer and the better electroanalytical properties of the sensor are obtained. Therefore, the lower resistance behaviour of BiFSPE compared to unmodified SPE indicates a better analytical sensor [55].

3.5. Interference Study and Real Sample Analysis

The interference effect of the following possible interfering substances Cd(II), Zn(II), Cu(II), Sn(II), Sb(III), Hg(II), Fe(III), As(V), K(I), I, Ca(II), and NO3 on the Pb(II) stripping signal was checked by SWASV in 0.1 M acetate buffer by preparing solutions with a Pb(II):interferent mass concentration ratio of 1:1. The tested interferents could be present in real water samples. The influence of possible interferents on the Pb(II) stripping signal was calculated as the change in % = i p interferent     i p Pb II i p Pb II · 100 , where i p interferent   and i p Pb II are the Pb(II) stripping signals with and without the presence of a particular interferent, respectively. The calculated changes are shown in Table 1. The interference effect may occur due to the competition between Pb(II) and interferents in preconcentration step, thus the increase (positive change) or decrease (negative change) of the Pb(II) stripping signal is observed [56].
A significant effect on the Pb(II) stripping signal is observed with Sn(IV) (the stripping signal for Pb(II) decreased by 57.9%) and Sb(III) (the stripping signal for Pb(II) increased by 49.4%). This was expected since Sn(IV) and Sb(III) are used to form in-situ modified electrodes [18,57]. Ions such as Cd(II), Zn(II), Hg(II), Fe(III), As(V), K(I), Ca(II), I, and NO3 had a minor effect on the Pb(II) stripping signal as the change in the Pb(II) signal was between ± 23.0%. Cu(II) did not have a significant influence on Pb(II) stripping signal (Pb(II) signal increased by only 3.3%).
For real sample analysis, the BiFSPE was employed, as it showed accurate and precise results in the method validation procedure (as given above). The accuracy and precision were tested in the same manner, as described in Section 3.3. The real sample analysis showed the accurate and precise result as the average Re was between 80.0 and 120.0% with RSD value of less than 20.0%.

4. Conclusions

The aim of this work was to develop and validate an electroanalytical method for the determination of Pb(II) in a single drop (60 µL). An unmodified SPE (a bare SPE with an Ag/AgCl reference electrode) and a Bi-film SPE (BiFSPE) were used for the trace determination of Pb(II) by means of square wave anodic stripping voltammetry (SWASV). The electrode modification of the BiFSPE was performed in situ with 0.5 mg/L Bi(III) in 0.1 M acetate buffer.
For Pb(II), the LOD and LOQ with the unmodified SPE could not be determined due to the intensive background contribution. For the BiFSPE, the obtained LOD was 1.2 μg/L and the obtained LOQ was 3.3 μg/L. The linear concentration ranges of Pb(II) obtained for the SPE and BiFSPE were 171.4–225.0 µg/L and 138.8–162.5 µg/L, respectively. It was also found that the sensitivity of the BiFSPE was higher compared to the unmodified SPE, which was assessed by the slope of the calibration curve. Using the unmodified SPE, the accuracy and precision were not satisfactory (the average recovery was not within 80.0–120.0% and the RSD was higher than 20.0%). However, when the BiFSPE was used, the average recovery and the RSD were 106.6% and 12.1%, respectively; thus, the BiFSPE provides accurate and precise results for Pb(II) determination in a single drop solution. The interference effect was studied in the presence of Cd(II), Zn(II), Cu(II), Sn(IV), Sb(III), Hg(II), Fe(III), As(V), K(I), I, Ca(II), and NO3. The most significant change in the Pb(II) stripping signal was found in the presence of Sn(IV) and Sb(III). The method showed accurate and precise results in the analysis of real water sample using BiFSPE. In addition, electrochemical impedance spectroscopy measurements showed that the BiFSE has lower charge transfer resistance compared with bare SPE, which makes the BiFSE more useful for electroanalytical studies.
This work demonstrates a very promising research area due to the potential for rapid on-site analysis of heavy metals and other pollutants. This work shows some shortcomings of SPEs and may therefore contribute to further research on this topic, where various surface modifications can lead to even better analytical performance.

Author Contributions

Conceptualization, M.F.; Data curation, M.F., N.Ž. and D.M.; Formal analysis, M.F. and N.Ž.; Funding acquisition, M.F.; Investigation, M.F. and N.Ž.; Methodology, M.F., N.Ž. and D.M.; Project administration, M.F.; Resources, M.F.; Software, M.F., N.Ž. and D.M.; Supervision, M.F.; Validation, M.F., N.Ž. and D.M.; Visualization, M.F. and N.Ž.; Roles/Writing—original draft, M.F. and N.Ž.; Writing—review & editing, M.F., N.Ž. and D.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Slovenian Research Agency grant numbers J1-9169, J1-2470, P2-0032, and J3-1762.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Acknowledgments

The authors would like to acknowledge the financial support for this project received from the Slovenian Research Agency (Grant Nos. J1-9169, J1-2470, P2-0032, and J3-1762).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Renedo, O.D.; Alonso-Lomillo, M.A.; Martínez, M.J.A. Recent developments in the field of screen-printed electrodes and their related applications. Talanta 2007, 73, 202–219. [Google Scholar] [CrossRef] [PubMed]
  2. Knežević, S.; Ognjanović, M.; Nedić, N.; Mariano, J.F.M.L.; Milanović, Z.; Petković, B.; Antić, B.; Djurić, S.V.; Stanković, D. A single drop histamine sensor based on AuNPs/MnO2 modified screen-printed electrode. Microchem. J. 2020, 155, 104778. [Google Scholar] [CrossRef]
  3. Ji, D.; Shi, Z.; Liu, Z.; Low, S.S.; Zhu, J.; Zhang, T.; Chen, Z.; Yu, X.; Lu, Y.; Lu, D.; et al. Smartphone-based square wave voltammetry system with screen-printed graphene electrodes for norepinephrine detection. Smart Mater. Med. 2020, 1, 1–9. [Google Scholar] [CrossRef]
  4. Chiu, M.-H.; Cheng, W.-L.; Muthuraman, G.; Hsu, C.-T.; Chung, H.-H.; Zen, J.-M. A disposable screen-printed silver strip sensor for single drop analysis of halide in biological samples. Biosens. Bioelectron. 2009, 24, 3008–3013. [Google Scholar] [CrossRef]
  5. Niu, P.; Fernández-Sánchez, C.; Gich, M.; Navarro-Hernández, C.; Fanjul-Bolado, P.; Roig, A. Screen-printed electrodes made of a bismuth nanoparticle porous carbon nanocomposite applied to the determination of heavy metal ions. Microchim. Acta. 2016, 183, 617–623. [Google Scholar] [CrossRef]
  6. Yantasee, W.; Deibler, L.A.; Fryxell, G.E.; Timchalk, C.; Lin, Y. Screen-printed electrodes modified with functionalized mesoporous silica for voltammetric analysis of toxic metal ions. Electrochem. Commun. 2005, 7, 1170–1176. [Google Scholar] [CrossRef]
  7. Lu, Y.; Liang, X.; Niyungeko, C.; Zhou, J.; Xu, J.; Tian, G. A review of the identification and detection of heavy metal ions in the environment by voltammetry. Talanta 2018, 178, 324–338. [Google Scholar] [CrossRef]
  8. Ariño, C.; Serrano, N.; Díaz-Cruz, J.M.; Esteban, M. Voltammetric determination of metal ions beyond mercury electrodes. A review. Anal. Chim. Acta 2017, 990, 11–53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. He, M.; Huang, L.; Zhao, B.; Chen, B.; Hu, B. Advanced functional materials in solid phase extraction for ICP-MS determination of trace elements and their species—A review. Anal. Chim. Acta 2017, 973, 1–24. [Google Scholar] [CrossRef]
  10. Bettinelli, M.; Beone, G.M.; Spezia, S.; Baffi, C. Determination of heavy metals in soils and sediments by microwave-assisted digestion and inductively coupled plasma optical emission spectrometry analysis. Anal. Chim. Acta 2000, 424, 289–296. [Google Scholar] [CrossRef]
  11. Krishna, M.V.B.; Arunachalam, J. Ultrasound-assisted extraction procedure for the fast estimation of major, minor and trace elements in lichen and mussel samples by ICP-MS and ICP-AES. Anal. Chim. Acta 2004, 522, 179–187. [Google Scholar] [CrossRef]
  12. Mohammed, H.; Sadeek, S.; Mahmoud, A.R.; Zaky, D. Comparison of AAS, EDXRF, ICP-MS and INAA performance for determination of selected heavy metals in HFO ashes. Microchem. J. 2016, 128, 1–6. [Google Scholar] [CrossRef]
  13. Daşbaşı, T.; Saçmacı, Ş.; Çankaya, N.; Soykan, C. A new synthesis, characterization and application chelating resin for determination of some trace metals in honey samples by FAAS. Food Chem. 2016, 203, 283–291. [Google Scholar] [CrossRef] [PubMed]
  14. Alves, G.M.S.; Rocha, L.S.; Soares, H.M.V.M. Multi-element determination of metals and metalloids in waters and wastewaters, at trace concentration level, using electroanalytical stripping methods with environmentally friendly mercury free-electrodes: A review. Talanta 2017, 175, 53–68. [Google Scholar] [CrossRef] [PubMed]
  15. Injang, U.; Noyrod, P.; Siangproh, W.; Dungchai, W.; Motomizu, S.; Chailapakul, O. Determination of trace heavy metals in herbs by sequential injection analysis-anodic stripping voltammetry using screen-printed carbon nanotubes electrodes. Anal. Chim. Acta 2010, 668, 54–60. [Google Scholar] [CrossRef] [PubMed]
  16. Qian, L.; Durairaj, S.; Prins, S.; Chen, A. Nanomaterial-based electrochemical sensors and biosensors for the detection of pharmaceutical compounds. Biosens. Bioelectron. 2021, 112836. [Google Scholar] [CrossRef]
  17. Shitanda, I.; Mitsumoto, M.; Loew, N.; Yoshihara, Y.; Watanabe, H.; Mikawa, T.; Tsujimura, S.; Itagaki, M.; Motosuke, M. Continuous Sweat Lactate Monitoring System with Integrated Screen-Printed MgO-templated Carbon-Lactate Oxidase Biosensor and Microfluidic Sweat Collector. Electrochim. Acta 2021, 137620. [Google Scholar] [CrossRef]
  18. Finšgar, M.; Majer, D.; Maver, U.; Maver, T. Reusability of SPE and Sb-modified SPE Sensors for Trace Pb(II) Determination. Sensors 2018, 18, 3976. [Google Scholar] [CrossRef] [Green Version]
  19. Wang, J.; Lu, J.; Hocevar, S.B.; Farias, P.A.M.; Ogorevc, B. Bismuth-Coated Carbon Electrodes for Anodic Stripping Voltammetry. Anal. Chem. 2000, 72, 3218–3222. [Google Scholar] [CrossRef]
  20. Jovanovski, V.; Hočevar, S.B.; Ogorevc, B. Bismuth electrodes in contemporary electroanalysis. Curr. Opin. Electrochem. 2017, 3, 114–122. [Google Scholar] [CrossRef]
  21. Bedin, K.C.; Mitsuyasu, E.Y.; Ronix, A.; Cazetta, A.L.; Pezoti, O.; Almeida, V.C. Inexpensive bismuth-film electrode supported on pencil-lead graphite for determination of Pb(II) and Cd(II) Ions by anodic stripping voltammetry. Int. J. Anal. Chem. 2018, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Colozza, N.; Cacciotti, I.; Moscone, D.; Arduini, F. Effects of Humidity, Temperature and Bismuth Electrodeposition on Electroanalytical Performances of Nafion-coated Printed Electrodes for Cd2+ and Pb2+ Detection. Electroanalysis 2020, 32, 345–357. [Google Scholar] [CrossRef]
  23. Sánchez-Calvo, A.; Blanco-López, M.C.; Costa-García, A. Paper-Based Working Electrodes Coated with Mercury or Bismuth Films for Heavy Metals Determination. Biosensors 2020, 10, 52. [Google Scholar] [CrossRef] [PubMed]
  24. Zidarič, T.; Jovanovski, V.; Menart, E.; Zorko, M.; Kolar, M.; Veber, M.; Hočevar, S.B. Multi-pulse galvanostatic preparation of nanostructured bismuth film electrode for trace metal detection. Sens. Actuators B Chem. 2017, 245, 720–725. [Google Scholar] [CrossRef]
  25. Švancara, I.; Prior, C.; Hočevar, S.B.; Wang, J. A Decade with Bismuth-Based Electrodes in Electroanalysis. Electroanalysis 2010, 22, 1405–1420. [Google Scholar] [CrossRef]
  26. Shi, L.; Li, Y.; Rong, X.; Wang, Y.; Ding, S. Facile fabrication of a novel 3D graphene framework/Bi nanoparticle film for ultrasensitive electrochemical assays of heavy metal ions. Anal. Chim. Acta 2017, 968, 21–29. [Google Scholar] [CrossRef] [PubMed]
  27. Rojas-Romo, C.; Serrano, N.; Ariño, C.; Arancibia, V.; Díaz-Cruz, J.M.; Esteban, M. Determination of Sb(III) using an ex-situ bismuth screen-printed carbon electrode by adsorptive stripping voltammetry. Talanta 2016, 155, 21–27. [Google Scholar] [CrossRef] [Green Version]
  28. Hutton, E.A.; Hočevar, S.B.; Ogorevc, B. Ex situ preparation of bismuth film microelectrode for use in electrochemical stripping microanalysis. Anal. Chim. Acta 2005, 537, 285–292. [Google Scholar] [CrossRef]
  29. Pauliukaite, R.; Metelka, R.; Švancara, I.; Królicka, A.; Bobrowski, A.; Norkus, E.; Kalcher, K.; Vytřas, K. Screen-printed carbon electrodes bulk-modified with Bi2O3 OR Sb2O3 for trace determination of some heavy metals. Sci. Pap. Univ. Pardubic. Ser. A 2004, 10, 47–58. [Google Scholar]
  30. Jovanovski, V.; Hrastnik, N.I.; Hočevar, S.B. Copper film electrode for anodic stripping voltammetric determination of trace mercury and lead. Electrochem. Commun. 2015, 57, 1–4. [Google Scholar] [CrossRef]
  31. Cui, L.; Wu, J.; Ju, H. Electrochemical sensing of heavy metal ions with inorganic, organic and bio-materials. Biosens. Bioelectron. 2015, 63, 276–286. [Google Scholar] [CrossRef]
  32. Maczuga, M.; Economou, A.; Bobrowski, A.; Prodromidis, M.I. Novel screen-printed antimony and tin voltammetric sensors for anodic stripping detection of Pb(II) and Cd(II). Electrochim. Acta 2013, 114, 758–765. [Google Scholar] [CrossRef]
  33. Bobrowski, A.; Maczuga, M.; Królicka, A.; Konstanteli, E.; Sakellaropoulou, C.; Economou, A. Determination of Copper(II) Through Anodic Stripping Voltammetry in Tartrate Buffer Using an Antimony Film Screen-printed Carbon Electrode. Anal. Lett. 2017, 50, 2920–2936. [Google Scholar] [CrossRef]
  34. Bobrowski, A.; Putek, M.; Zarębski, J. Antimony Film Electrode Prepared In Situ in Hydrogen Potassium Tartrate in Anodic Stripping Voltammetric Trace Detection of Cd(II), Pb(II), Zn(II), Tl(I), In(III) and Cu(II). Electroanalysis 2012, 24, 1071–1078. [Google Scholar] [CrossRef]
  35. Finšgar, M.; Xhanari, K.; Petovar, B. A Comparison of Hydrochloric Acid and Acetate Buffer Media for Trace Metal Analysis Using Sb-Film Electrodes: A Detailed Electrochemical Impedance Spectroscopy Study. J. Electrochem. Soc. 2019, 166, H108–H118. [Google Scholar] [CrossRef]
  36. Finšgar, M.; Xhanari, K.; Petovar, B. Copper-film electrodes for Pb(II) trace analysis and a detailed electrochemical impedance spectroscopy study. Microchem. J. 2019, 147, 863–871. [Google Scholar] [CrossRef]
  37. Ashrafi, A.M.; Vytřas, K. Codeposited Antimony-Bismuth Film Carbon Paste Electrodes for Electrochemical Stripping Determination of Trace Heavy Metals. Int. J. Electrochem. Sci. 2013, 8, 2095–2103. [Google Scholar]
  38. Jovanovski, V.; Hrastnik, N.I. Insights into the anodic stripping voltammetric behaviour of copper film electrodes for determination of trace mercury. Microchem. J. 2019, 146, 895–899. [Google Scholar] [CrossRef]
  39. Zhu, W.W.; Li, N.B.; Luo, H.Q. Simultaneous determination of chromium(III) and cadmium(II) by differential pulse anodic stripping voltammetry on a stannum film electrode. Talanta 2007, 72, 1733–1737. [Google Scholar] [CrossRef] [PubMed]
  40. Czop, E.; Economou, A.; Bobrowski, A. A study of in situ plated tin-film electrodes for the determination of trace metals by means of square-wave anodic stripping voltammetry. Electrochim. Acta 2011, 56, 2206–2212. [Google Scholar] [CrossRef]
  41. Tian, Y.Q.; Li, N.B.; Luo, H.Q. Simultaneous Determination of Trace Zinc(II) and Cadmium(II) by Differential Pulse Anodic Stripping Voltammetry Using a MWCNTs–NaDBS Modified Stannum Film Electrode. Electroanalysis 2009, 21, 2584–2589. [Google Scholar] [CrossRef]
  42. Li, B.L.; Wu, Z.L.; Xiong, C.H.; Luo, H.Q.; Li, N.B. Anodic stripping voltammetric measurement of trace cadmium at tin-coated carbon paste electrode. Talanta 2012, 88, 707–710. [Google Scholar] [CrossRef]
  43. Korolczuk, M.; Tyszczuk, K.; Grabarczyk, M. Adsorptive stripping voltammetry of nickel and cobalt at in situ plated lead film electrode. Electrochem. Commun. 2005, 7, 1185–1189. [Google Scholar] [CrossRef]
  44. Tyszczuk-Rotko, K.; Maj, J. A Lead Film Electrode for Adsorptive Stripping Voltammetric Analysis of Ultratrace Tungsten(VI) in Acidic Medium. Electroanalysis 2012, 24, 101–106. [Google Scholar] [CrossRef]
  45. Tchounwou, P.B.; Yedjou, C.G.; Patlolla, A.K.; Sutton, D.J. Heavy Metal Toxicity and the Environment; Springer: Basel, Switzerland, 2012. [Google Scholar]
  46. Žurga, N. Analysis of Trace Heavy Metals in a Sample Droplet with Lab-On-A-Chip System. Ph.D. Thesis, University of Maribor, Maribor, Slovenia, 2020. [Google Scholar]
  47. Zhao, G.; Wang, H.; Liu, G.; Wang, Z. Box–Behnken response surface design for the optimization of electrochemical detection of cadmium by Square Wave Anodic Stripping Voltammetry on bismuth film/glassy carbon electrode. Sens. Actuators B Chem. 2016, 235, 67–73. [Google Scholar] [CrossRef]
  48. Plum, L.M.; Rink, L.; Haase, H. The Essential Toxin: Impact of Zinc on Human Health. Int. J. Env. Res. Public Health 2010, 7, 1342–1365. [Google Scholar] [CrossRef] [Green Version]
  49. Fosmire, G.J. Zinc Toxicity. Am. J. Clin. Nutr. 1990, 51, 225–227. [Google Scholar] [CrossRef]
  50. Petrović, S.; Guzsvány, V.; Ranković, N.; Beljin, J.; Rončević, S.; Dalmacija, B.; Ashrafi, A.M.; Kónya, Z.; Švancara, I.; Vytřas, K. Trace level voltammetric determination of Zn(II) in selected nutrition related samples by bismuth-oxychloride-multiwalled carbon nanotube composite based electrode. Microchem. J. 2019, 146, 178–186. [Google Scholar] [CrossRef]
  51. Ghorbani, M.; Pedramrad, T.; Aghamohammadhasan, M.; Seyedin, O.; Akhlaghi, H.; Afshar Lahoori, N. Simultaneous clean-up and determination of Cu(II), Pb(II) and Cr(III) in real water and food samples using a magnetic dispersive solid phase microextraction and differential pulse voltammetry with a green and novel modified glassy carbon electrode. Microchem. J. 2019, 147, 545–554. [Google Scholar] [CrossRef]
  52. Filice, S.; Mazurkiewicz-Pawlicka, M.; Malolepszy, A.; Stobinski, L.; Kwiatkowski, R.; Boczkowska, A.; Gradon, L.; Scalese, S. Sulfonated Pentablock Copolymer Membranes and Graphene Oxide Addition for Efficient Removal of Metal Ions from Water. Nanomaterials 2020, 10, 1157. [Google Scholar] [CrossRef] [PubMed]
  53. Finšgar, M.; Xhanari, K.; Otmačić Ćurković, H. Cyclic Voltammetry as an Electroanalytical Tool for Analysing the Reaction Mechanisms of Copper in Chloride Solution Containing Different Azole Compounds. Curr. Anal. Chem. 2020, 16, 465–474. [Google Scholar] [CrossRef]
  54. United Nations Office on Drugs, Crime, Laboratory, Scientific Section. Guidance for the Validation of Analytical Methodology and Calibration of Equipment used for Testing of Illicit Drugs in Seized Materials and Biological Specimens, A Commitment to Quality and Continuous Improvement; United Nations Publication: New York, NY, USA, 2009. [Google Scholar]
  55. Petovar, B.; Xhanari, K.; Finšgar, M. A detailed electrochemical impedance spectroscopy study of a bismuth-film glassy carbon electrode for trace metal analysis. Anal. Chim. Acta 2018, 1004, 10–21. [Google Scholar] [CrossRef] [PubMed]
  56. Finšgar, M.; Kovačec, L. Copper-bismuth-film in situ electrodes for heavy metal detection. Microchem. J. 2020, 154, 104635. [Google Scholar] [CrossRef]
  57. Finšgar, M.; Petovar, B.; Vodopivec, K. Bismuth-tin-film electrodes for Zn(II), Cd(II), and Pb(II) trace analysis. Microchem. J. 2019, 145, 676–685. [Google Scholar] [CrossRef]
Figure 1. (a) Voltammograms measured for the unmodified SPE at different γPb(II) for a single drop placed on the SPE holder, (b) the linear concentration range, and (c) the change in EPb(II) vs. γPb(II). The measurements were performed at Eacc = −1.5 V and tacc = 60 s in 0.1 M acetate buffer (without Bi(III)).
Figure 1. (a) Voltammograms measured for the unmodified SPE at different γPb(II) for a single drop placed on the SPE holder, (b) the linear concentration range, and (c) the change in EPb(II) vs. γPb(II). The measurements were performed at Eacc = −1.5 V and tacc = 60 s in 0.1 M acetate buffer (without Bi(III)).
Chemosensors 09 00038 g001
Figure 2. (a) Voltammograms measured for the BiFSPE at different γPb(II) for a single drop placed on the SPE holder, (b) the linear concentration range, and (c) the change in EPb(II) vs. γPb(II). The measurements were performed at Eacc = −1.5 V and tacc = 60 s in 0.1 M acetate buffer containing 0.5 mg/L Bi(III).
Figure 2. (a) Voltammograms measured for the BiFSPE at different γPb(II) for a single drop placed on the SPE holder, (b) the linear concentration range, and (c) the change in EPb(II) vs. γPb(II). The measurements were performed at Eacc = −1.5 V and tacc = 60 s in 0.1 M acetate buffer containing 0.5 mg/L Bi(III).
Chemosensors 09 00038 g002
Figure 3. Method of multiple standard addition using the BiFSPE for the determination of γPb(II) in a single drop of solution consisting of the 0.1 M acetate buffer spiked with 140.0 μg/L Pb(II).
Figure 3. Method of multiple standard addition using the BiFSPE for the determination of γPb(II) in a single drop of solution consisting of the 0.1 M acetate buffer spiked with 140.0 μg/L Pb(II).
Chemosensors 09 00038 g003
Figure 4. EIS Nyquist spectra measured for unmodified SPE and BiFSPE in 0.1 M acetate buffer containing 140.0 μg/L Pb(II).
Figure 4. EIS Nyquist spectra measured for unmodified SPE and BiFSPE in 0.1 M acetate buffer containing 140.0 μg/L Pb(II).
Chemosensors 09 00038 g004
Table 1. The influence of different ions on the Pb(II) stripping signal at Pb(II):interferent mass concentration ratio of 1:1.
Table 1. The influence of different ions on the Pb(II) stripping signal at Pb(II):interferent mass concentration ratio of 1:1.
InterferentChange in %
Cd(II)20.3
Zn(II)−17.8
Cu(II)3.3
Sn(IV)−57.9
Sb(III)49.4
Hg(II)22.8
Fe(III)−16.2
As(V)−15.8
K(I)15.3
Ca(II)−16.7
I−13.8
NO310.5
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Žurga, N.; Majer, D.; Finšgar, M. Pb(II) Determination in a Single Drop Using a Modified Screen-Printed Electrode. Chemosensors 2021, 9, 38. https://doi.org/10.3390/chemosensors9020038

AMA Style

Žurga N, Majer D, Finšgar M. Pb(II) Determination in a Single Drop Using a Modified Screen-Printed Electrode. Chemosensors. 2021; 9(2):38. https://doi.org/10.3390/chemosensors9020038

Chicago/Turabian Style

Žurga, Nika, David Majer, and Matjaž Finšgar. 2021. "Pb(II) Determination in a Single Drop Using a Modified Screen-Printed Electrode" Chemosensors 9, no. 2: 38. https://doi.org/10.3390/chemosensors9020038

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop