Next Article in Journal
Biodegradable Polyhydroxyalkanoates Formed by 3- and 4-Hydroxybutyrate Monomers to Produce Nanomembranes Suitable for Drug Delivery and Cell Culture
Previous Article in Journal
The U-Net Family for Epicardial Adipose Tissue Segmentation and Quantification in Low-Dose CT
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Electrochemical Detection of Furaltadone Antibiotic Drug by the Rare Earth Metal Tungstate Decorated Screen Printed Carbon Electrode

by
Sivaramakrishnan Vinothini
1,2,
Te-Wei Chiu
1,2,* and
Subramanian Sakthinathan
1,2
1
Department of Materials and Mineral Resources Engineering, National Taipei University of Technology, No. 1, Section 3, Zhongxiao E. Rd, Taipei 106, Taiwan
2
Institute of Materials Science and Engineering, National Taipei University of Technology, No. 1, Section 3, Zhongxiao E. Rd, Taipei 106, Taiwan
*
Author to whom correspondence should be addressed.
Technologies 2023, 11(4), 105; https://doi.org/10.3390/technologies11040105
Submission received: 31 May 2023 / Revised: 19 July 2023 / Accepted: 2 August 2023 / Published: 6 August 2023

Abstract

:
Furaltadone (FLD) is an antibiotic drug that is widely treated for coccidiosis, intestinal infection, and turkey blackhead. Moreover, excessive use of FLD may have some negative consequences for humans and domestic animals. Therefore, practical, sensitive, selective, and facile detection of FLD is still needed. In this exploration, a Eu2(WO4)3-nanoparticles-modified screen-printed carbon electrode was developed for the low-level detection of FLD. Hydrothermal techniques were used effectively to prepare the Eu2(WO4)3 complex. Scanning electron microscopy and X-ray diffraction investigations were used to confirm the Eu2(WO4)3. The results revealed that the Eu2(WO4)3 was well formed, crystalline, and uniformly distributed. Furthermore, the electrochemical behavior of the SPCE/Eu2(WO4) electrode was examined by differential pulse voltammetry and cyclic voltammetry studies. The SPCE/Eu2(WO4) electrode demonstrated improved electrocatalytic activity in the detection of FLD with a detection limit of 97 µM (S/N = 3), linear range of 10 nM to 300 µM, and sensitivity of 2.1335 µA µM−1 cm−2. The SPCE/Eu2(WO4) electrode detected FLD in the presence of 500-fold excess concentrations of other interfering pollutant ions. The practical feasibility of the SPCE/Eu2(WO4) electrode was tested on different antibiotic medicines and showed adequate recovery. Moreover, the SPCE/Eu2(WO4) electrode shows appreciable repeatability, high stability, and reproducibility.

Graphical Abstract

1. Introduction

Furaltadone (FLT) is a nitrofuran derivative in the family of synthetic antimicrobials, antibiotics, and chemotherapeutic agents. FLT antibiotics are widely used to treat coccidiosis, intestinal infections, and turkey blackhead [1,2]. However, excessive usage of FLT can cause side effects. Generally, FLT is found in animal products such as seafood, fish, poultry, shrimp, and cattle. FTD is commonly applied as a feed additive to promote growth evolution, bacterial infections, prevents protozoa in bee colonies, livestock, and aquaculture [3]. FTD is an antibiotic and a synthetic chemotherapeutic drug also found in poultry, fish, cattle, and shrimp. FLT and its derivatives have been linked to carcinogenic and mutagenic effects [4]. Therefore, many countries, such as the United States of America, Australia, China, the Philippines, Thailand, and Brazil, have banned FLT in food production and antibiotics medicine [5]. The European Union forbidden FLT use in the production of animal food and aquaculture products. Moreover, the World Health Organization (WHO) set the FLT residue level at 1 g/kg, and the Chinese Ministry of Agriculture classified an acceptable level of FLD in aquatic foods as 0.25 μg kg−1 [6]. Hence, the need to develop a technique to identify the quantity of FLT in foods and animal products is necessary. Furthermore, the persistent release of FLT into the green environment as a result of hospital or industrial disposal produces chronic toxicity in all living species. As a result, many academics and industrialists are concerned about the mineralization of FLT from industry. In addition to that, FLT can be metabolized rapidly in the human body, which contributes to its related diseases [7]. Therefore, a practical, facile, and sensitive technique for the detection of the aforementioned hazards of FLT is high importance [8,9]. It is even more important and difficult to prepare interface materials for the electrochemical detection of furaltadone because FLD has a short half-life, is unstable, easily breaks down into its metabolite and 3-amino-5-morpholinomethyl-2-oxazolidinone [10,11,12,13].
Different detection methods have been reported to determine the level of antibiotic drugs. They include liquid chromatography-UV photodiode array detection, high-performance liquid chromatography (HPLC), spectrophotometry, liquid chromatography-mass spectrometry (LCMS), fluorimetry, and enzyme-linked immunosorbent sensors [14]. The significant disadvantages of using these traditional methods are that they have a longer duration until results, entail high costs, need well-qualified technicians to operate, are sophisticated instrumentation, and require critical sample preparation [15]. To overcome these issues, the electrochemical method was used due to its low cost, rapid response, simplicity, reliability, excellent sensitivity, marvelous selectivity, and better reproducibility. In comparison to other methodologies, the electrochemical method is the most effective tool for analyzing FTD [16].
Rare earth metals have been used in metallurgy, optics, electrocatalysis, magnetic and other fields due to their unique properties. Rare earth metals have excellent properties and are widely applied in laser materials, superconductors, batteries, solar cells, supercapacitors, and electrochemical sensors due to their different electrical properties [17,18]. Rare earth elements are also fascinating due to the unusual features that result from their electrical structure. These qualities can be attributed to the contraction of 4f orbitals, which are altered by the 5s and 5p shells and have a significant impact on the chemical properties. These elements are very interesting for their excellent mechanical, thermal, chemical, optical, and electrical properties that can be applied in photonic devices and catalysts [19,20].
The rare earth element tungstate has the general formula RE2(WO4)3, where R stands for a rare earth element (RE = La2+, Ce2+, Pr2+, Er2+, etc.). Moreover, tungstate has demonstrated remarkable performance in ionic conductors, solid-state lasers, and catalysis. It is regarded as an ideal host material for doped RE ions due to the unique characteristics of the WO42− groups. Eu2(WO4)3 is a critical member of the tungstate family due to its excellent ionic characteristics and strong electrical conductivity [21]. Eu2(WO4)3 has very good catalytic properties in addition to its synergistic properties. Among all these tungstates, Eu2(WO4)3 is important for its high thermal stability, optical, electrical properties, supercapacitors, and Li-ion battery. Eu2(WO4)3 tends to form mixed phases by the solid-state preparation processes in high calcination temperatures. For the synthesis of rare earth metal tungstate, various methods have been developed, including sol-gel processing, solid-state reaction, microemulsion, molten-salt-assisted crystallization, and hydrothermal reaction. Surprisingly, one of the better alternative methods for nanomaterial synthesis is the hydrothermal method [20,21,22].
In this paper, we describe the hydrothermal synthesis of Eu2(WO4)3 as a simple, flexible, and straightforward method. In addition, the electrochemical performance of as-prepared Eu2(WO4)3 nanoparticles was evaluated using different characterization methods. The synthesized Eu2(WO4)3 was applied as a sensor material for the determination of FTD. The electrocatalytic behavior of the SPCE/Eu2(WO4)3 electrode used for FTD detection was very high compared to that of the unmodified SPCE. Furthermore, we find that the SPCE/Eu2(WO4)3 had excellent electrochemical activity for FTD detection and could be successfully applied to real sample analysis. The developed SPCE/Eu2(WO4)3 electrode demonstrated high sensitivity, selectivity, and the lowest detection limit for FTD detection. These electrochemical detections have multiple benefits, including ease of use, excellent selectivity, high sensitivity, low cost, and quick response times. At the same time, these detection techniques have several downsides, including the costly raw materials of the electrode, longer preparation time, and environmental pollution during the preparation of the electrode material. Additionally, the electrodes must be prepared using the right procedures to maintain consistent performance. The prepared SPCE/Eu2(WO4)3 electrochemical sensor also has outstanding electrochemical sensing capabilities for the detection of FT and demonstrates its potential catalytic sensing use in pharmaceutical samples.

2. Materials and Methods

Screen-printed carbon electrodes (0.08 cm2) were acquired from Zensor R&D Co., Ltd., Taipei, Taiwan. Europium (III) chloride (EuCl3·6H2O), methanol, sodium tungstate dihydrate (Na2WO4·2H2O), furmethonol (C13H16N4O6), and all other chemicals were received from Sigma-Aldrich Chemical Company, Taipei, Taiwan. Furthermore, sulfuric acid (H2SO4) chemical, and hydrochloric acid (HCl) chemicals were obtained from Merck Chemical Company. The supporting phosphate-buffered solution (PBS) electrolyte was made by using 0.05 mol L−1 of Na2HPO4 and 0.05 mol L−1 of monosodium phosphate (NaH2PO4). These FTD standard solutions were prepared by dissolving the appropriate amount of dimethylformamide (DMF) with FTD. All chemicals were made by analytical-reagent grade chemicals without further purification. All solutions were made using double distilled water, PBS (pH 7) was used to make a stock solution of furaltadone, and nitrogen purging was used to simultaneously deoxygenate the electrolyte.
The electrochemical evaluation by differential pulse voltammetry and cyclic voltammetry studies was conducted on CHI 1211B workstations, respectively. The CHI 1211B workstations were linked to a standard three-electrode cell that included an Ag/AgCl electrode as a reference electrode, an SPCE as a working electrode, and a Pt needle was used as counter electrodes, respectively. All the electrochemical studies were conducted at normal laboratory temperatures. Field emission scanning electron microscopy (FESEMEDX, JEOL JSM-7610F, Tokyo, Japan, and Hitachi Regulus 8100) combined with energy-dispersive X-ray spectroscopy was used to identify and assess the surface morphology with the elemental arrangement. Micro Raman Spectrum (Uni Nano Tech Co., Ltd., ACRON, Yongin, Republic of Korea) with an excitation source of 532 nm and a resolution of 1 m was used to identify the Raman spectra. An X-ray diffractometer (D2 Phaser, Bruker, Billerica, MA, USA) was used to study spectra under CuKa radiation (l = 1.5418 A).

2.1. Preparation of Eu2(WO4)3 Nanoparticles by Hydrothermal Methods

Eu2(WO4)3 nanoparticles were prepared according to previous reports with slight modification [18,20,22,23]. A hydrothermal process was used to create Eu2(WO4)3 nanoparticles. Under the ambient condition, analytical grade Eu(NO3)3·6H2O and Na2WO4·2H2O reagent solutions were continually mixed, and the suspension was transported to an autoclave and thermally treated at 180 °C in the oven for 12 h. After the hydrothermal process, white color fine powder was obtained and separated by centrifuging with water, washed with ethanol/water mixture, and finally dried in the oven. The resulting white powder was washed with deionized water and pure ethanol. The precipitate was dried in a hot air oven set to 60 °C. The prepared materials were crushed into white fine powders using a Hitech Agate mortar. These materials were calcined at 600 °C for 5 h in a muffle furnace to create crystalline Eu2(WO4)3 nanoparticles.

2.2. Fabrication of SPCE/Eu2(WO4)3 Modified Electrode by Simple Drop-Casting Methods

The SPCE was cleaned with ethanol and water before being dried at room temperature to remove the weakly adsorbed contaminants. The ASPCE was prepared using previously described methods [24]. In detail, SPCE was initially cleaned with ethanol/water by sonication to remove the adsorbed materials on the ASPCE surface. Then, the cleaned SPCE was shifted into an electrochemical cell containing PBS (pH 7)/KCl electrolytes mixture, followed by the applied potential 0 to 2.0 V up to 20 cycles for ASPCE preparation. After the ASPCE preparation, 2 mg of prepared Eu2(WO4)3 nanoparticles were dispersed in 200 µL of DD water and kept in an ultra-sonication bath for 15 min to form a homogeneous white color suspension. Later, 8 µL of Eu2(WO4)3 nanoparticles were coated on the surface of the activated SPCE and dried in a hot oven at 50 °C. The obtained Eu2(WO4)3/SPCE electrode was used for additional electrochemical studies.

3. Characterization Studies of Materials

XRD, SEM, and EDX Studies

The structural morphology and crystalline structure arrangement of the as-prepared Eu2(WO4)3 nanoparticles are shown in Figure 1. Figure 1A shows that the XRD diffraction pattern of Eu2(WO4)3 nanoparticles displayed a monoclinic phase. The Eu2(WO4)3 had diffraction peaks at 28.55°, 30.8°, 34.12°, 46.95° and 57.74° for the 101, 102, 111, 112, and 120 planes, respectively. This diffraction is concordant with the standard PDF file (PDF#22-0287) of the Eu2(WO4)3 monoclinic phase [25]. The results indicate a good formation of Eu2(WO4)3 nanoparticles.
Figure 2A,B shows the FESEM structural images of the Eu2(WO4)3 nanoparticles showing the coral shape structure and the particles typical size of approximately 300 nm. Figure 2C shows the EDX spectra and elemental mapping results used to study the chemical composition of the Eu2(WO4)3 nanoparticles. Additionally, the elemental mapping results of Eu2(WO4)3 nanoparticles revealed the presence of europium (Figure 2D), tungstate (Figure 2E), and oxygen (Figure 2F). Furthermore, Figure 2G exhibits the EDX spectra of Eu2(WO4)3 nanoparticles, showing the homogeneously distributed elements observed, such as europium (Eu), tungstate (W), and oxygen (O).

4. Electrochemical Studies of FTD on the ASPCE/Eu2(WO4)3 Electrode

4.1. Effects of Different Films and Effects of Different Concentrations

The electrocatalytic activity of FLD at different modified electrodes was identified using CV at a scan rate of 50 mVs−1 in the presence of 100 µM of FLD in 0.05 M PBS (pH = 7). Figure 3A exhibits the electrochemical detection response of FLD at the SPCE, ASPCE, and ASPCE/Eu2(WO4)3 electrodes. The SPCE electrode had a low peak current of 73 μA for the addition of 100 μM of FLD. The ASPCE electrode exhibited the detection response of 80 μA at the potential of −0.4 V. The SPCE/Eu2(WO4)3 electrode exhibited a FLD detection peak current response of −106.3 μA. The SPCE/Eu2(WO4)3 electrode reached the FLD detection peak current response of −155.5 μA, indicating that the reduction potential of FLD was higher than those of the SPCE, ASPCE, and SPCE/Eu2(WO4)3 electrodes. However, no characteristic peak was observed for ASPCE/Eu2(WO4)3 in the absence of 100 μM FLD (Figure 3B). This aforementioned result indicates the excellent catalytic behavior of the ASPCE/Eu2(WO4)3 electrode towards FLD. This suggests that the Eu2(WO4)3 coated on the ASPCE can enhance the oxidation signal and electron transfer capability of the ASPCE/Eu2(WO4)3 electrode.
Electrochemical detection mechanisms of FLD contain a nitro group (R-NO2), which is electrochemically reduced to form a hydroxylamine group (R-NHOH) with an equal number of electrons (+4e) and protons (+4H+) in the electron transfer process (Scheme 1). This irreversible reduction process of FLD at the surface of the ASPCE/Eu2(WO4)3 electrode is indicated by the electrochemical reduction of FLD. A weak redox peak indicates O1 for oxidation and R2 for reduction at potentials of +0.15 V and 0.07 V, respectively. This electrochemical process turns hydroxylamine (NH2OH) into a nitroso derivative by the two-electron and two-proton transfer reaction. The R1 peak signal observed for the reduction peak current (R1) of FLD is significantly larger than the redox couple. The ASPCE/Eu2(WO4)3 demonstrated outstanding electrochemical performance for FLD sensing due to the highly specialized surface area, which promotes fast electron conductivity. The Randles Sevcik equation and the hexacyanoferrate system (K3[Fe(CN)6]3−/4−) were used to determine the modified electrode active surface area. The ASPCE (0.052 cm2) and SPCE/Eu2(WO4)3 (0.135 cm2) electrodes have much lower active surface areas than the ASPCE/Eu2(WO4)3 (0.191 cm2) electrodes. Hence, the active surface area of the ASPCA/Eu2(WO4)3 electrode is one of the reasons for the high electrochemical activity toward the detection of FLD.
In addition, Figure 4A shows the CV performances of different concentrations (100–700 µM) of FLD in deoxygenated 0.05 M PBS. The FLD reduction peak currents linearly increased with increases in the concentration of FLD (Figure 4B). The corresponding linear regression equation was obtained as IP = 75.58v1/2 (Vs−1)1/2 + 10.60 with R2 = 0.998. Therefore, an excellent and fast electron transfer was observed at the ASPCE/Eu2(WO4)3 surface due to the good electrostatic interaction between the ASPCE and Eu2(WO4)3, which increased the higher electron flow rate and surface area of the composite matrix.

4.2. Effect of Different Scan Rates and Effect of Different pH Studies

Figure 5A reveals a noteworthy CV curve obtained at the ASPCE/Eu2(WO4)3 electrode for the different scan rates in 100 µM of FLD containing PBS (pH 7). These results indicated that the FLD reduction peak current response linearly increased with increases in scan rates from 10 to 160 mVs1. Figure 5B shows the related peak current (IP) vs. scan rate plot and the corresponding linear regression equation was obtained as IP = 75.58v1/2 (Vs−1)1/2 + 10.60 with R2 = 0.998.
This result shows that the reduction of FLD is a surface-controlled process over the ASPCE/Eu2(WO4)3. Moreover, with an increasing scan rate, the electrocatalytic reduction of FLD at an ASPCE/Eu2(WO4)3 electrode peak shifted to a negative potential. The electron transfer coefficient and the number of electrons transferred are involved in the rate-determining step. In addition, the number of electrons transferred involved in the reduction of FLD can be calculated by the surface-controlled reaction. This result indicates that the FLD reduction leads to the final product, and it is in good accordance with the previously published results. The overall reduction reaction can be expressed by the electrocatalytic reduction of FLD at the ASPCE/Eu2(WO4)3 electrode as a two-electron transfer reaction.
The CV studies showed the influence of electrolyte pH on the ASPCE/Eu2(WO4)3 electrode for the detection of FTD. The CV experiment was recorded at ASPCE/Eu2(WO4)3 electrodes at pH levels of 3 to 11 during the electrocatalytic reduction of 100 µM of FTD. As mentioned in Figure 6A, the reduction peak current of FTD increased as electrolyte pH values rose from 3 to 7, and afterward, it decrease at pH levels of 9 to 11 (Figure 6B). This result exhibited that, at lower pH, the FTD molecules are unstable and easily converted into reduced products. In the same experiment at higher pHs, proton deficiency occurred in the electrolytes. However, at pH 7, the electrochemical sensor exhibited higher electrocatalytic performance toward the detection of FTD. Therefore, in this work, we chose pH 7 for further electrochemical studies.

4.3. DPV Studies for Detection of FLD

DPV technique is an overall more sensitive and unique procedure than other voltammetry methods. For this reason, we used these techniques for the limit of detection, sensitivity and linear limit identification. Figure 7A revealed DPV responses of FLD reduction on ASPCE/Eu2(WO4)3 in 0.05 M PB solution containing different concentrations of FLD addition (0.1–450 μM). The reduction peak current was expanded with increasing the concentration of FLD. Figure 7B indicates the linear calibration plot amid the peak current (Ip) and FLD concentration. The linear concentration ranges of the ASPCE/Eu2(WO4)3 electrode were 0.1–125 μM and the limit of detection (LOD) was 7.4 nM. Additionally, the sensitivity of the ASPCE/Eu2(WO4)3 electrode was fixed at about 4.1084 μA μM−1 cm−2, which is an excessive value compared with that of the other electrodes. In consequence, the results demonstrate that the ASPCE/Eu2(WO4)3 electrode exhibited superior detection of FLD due to the higher electron transfer rate and good-catalytic behavior of the Eu2(WO4). Furthermore, the surface nature of ASPCE distributed a strong interaction between the ASPCE and Eu2(WO4)3. For that reason, the ASPCE/Eu2(WO4)3 modified electrode was used as a predominant electrode material for the electrochemical detection of FLD. The acquired performance results are compared with previously published FLD sensors and mentioned in Table 1. Moreover, the electrochemical detection of the carbendazim at the ASPCE/Eu2(WO4)3 electrode was lower than that of the other reported electrodes and methods, such as the FeVO/p-rGO NCs electrode [LOD = 138 nM] [26], AOZ/AMOZ electrode [LOD = 10 μg kg−1] [27], LC-MS/MSc method [LOD = 0.5 μg kg−1] [28], ELISA method [LOD = 0.3 μg kg−1] [29], and LC-MS/MS method [LOD = 0.2 μg/kg] [30]. According to the comparison table, the ASPCE/Eu2(WO4) electrode had good selectivity, a wide linear range, and a low LOD. Therefore, Eu2(WO4)3 is a promising candidate for the electrochemical detection of FLD due to its higher electron transfer rate.

4.4. Repeatability, Cyclic Stability, Reproducibility, and Operational Stability Studies of Eu2(WO4)3 Coated Electrodes

The repeatability and reproducibility performance of the SPCE/Eu2(WO4)3 electrodes were identified by the CV technique. Also, the stability study was conducted in the amperometric technique. The repeatability behavior of three SPCE/Eu2(WO4)3 electrodes was examined to determine the FLD detection. The CV curves showed the same cathodic detection peak currents of FLD with an RSD value of 2.4% (Figure 8A), which indicated that our prepared SPCE/Eu2(WO4)3 electrodes have outstanding repeatability. Reproducibility studies were performed with 50 µM FLD over five different SPCE/Eu2(WO4)3 electrodes consecutively for reproducibility analysis with an RSD value of 2.1% (Figure 8B). Figure 8C shows the cyclic stability responses of the ASPCE/Eu2(WO4)3 electrode in the presence of 100 µM FLD. The CV spectra of the first and 100th cycles had the same peak current values, indicating the excellent cyclic stability of the ASPCE/Eu2(WO4)3 electrode. Furthermore, operational stability studies of the RRDE/Eu2(WO4)3 electrode’s response were seen at increases of 100 µM FLD at 50 s intervals, and it preserved 96.5% of its original current up to the 2000 s (Figure 8D). These results prove the excellent operational stability of the RRDE/Eu2(WO4)3 electrode.

4.5. Real Sample Studies of FTD in Antibiotic Medicine

The CV studies approach was utilized to detect FLD in silverzine, sindine, and tetracycline medicine samples with the ASPCE/Eu2(WO4)3 electrode. In real sample analysis studies, the standard addition method was applied for the detection of FLD from antibacterial medicine. Real samples were prepared as follows. Antibacterial medicine was dissolved in water samples and applied for real sample detection. The FLD was then added to the real samples in concentrations of 20, 40, and 60 µM, as shown in Figure 9A–C. Finally, the real sample recovery percentages of the silverzine, sindine, and tetracycline medicine samples ranged from 95–98%, as listed in Table 2. The real sample analysis confirmed that our prepared ASPCE/Eu2(WO4)3 electrode had a good recovery percentage for the detection of FLD in antibacterial medicine samples.

5. Conclusions

An ASPCE/Eu2(WO4)3 electrode was prepared by non-covalent interaction. The Eu2(WO4)3 complex was identified using different characterization studies. The electrochemical performance of the ASPCE/Eu2(WO4)3 electrode was used to evaluate the cyclic voltammetry, differential pulse voltammetry, and amperometry technique. The ASPCE/Eu2(WO4)3 has excellent electrocatalytic activity concerning the detection of FLD at 0.19 V. Further, the ASPCE/Eu2(WO4)3 electrode had higher electrocatalytic activity towards the detection of FLD due to the high active surface area of Eu2(WO4)3 and fast electrons transport between the ASPCE/Eu2(WO4)3 electrode and FLD. Hence, the proposed ASPCE/Eu2(WO4) electrode had good electrocatalytic activity towards the reduction of FLD, with a low detection limit of 97 µM, long linear range of 20–300 µM, and high sensitivity of 2.1335 μA μM−1 cm−2. Moreover, the ASPCE/Eu2(WO4)3 electrode exhibited good stability, reproducibility, and repeatability. The ASPCE/Eu2(WO4)3 electrode exhibited practical detection in silverzine, sindine, and tetracycline medicine samples with admissible recoveries. Hence, the ASPCE/Eu2(WO4)3 electrode is favorably applicable to the selective, and sensitive detection of FLD.

Author Contributions

Conceptualization, S.V. and S.S.; methodology, S.V.; software, S.V.; validation, S.V., S.S. and T.-W.C.; formal analysis, S.V.; investigation, S.V.; resources, S.V. and S.S.; data curation, S.V.; writing—original draft preparation, S.V.; writing—review and editing, S.V. and S.S.; visualization, S.V.; supervision, T.-W.C.; project administration, T.-W.C.; funding acquisition, T.-W.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Ministry of Science and Technology (MOST 109–2221-E-027–059, 110–2221-E-027–041, NSTC 111–2221–E–027–104, and NSTC 112-2221-E-027-039).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are available on request.

Acknowledgments

The authors thank Abinaya Meenakshi, Functional Oxide Laboratory, National Taipei University of Technology (NTUT), Taiwan, for materials characterization help. The authors appreciate the Precision Research and Analysis Centre of the National Taipei University of Technology (NTUT) for providing the measurement facilities.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Diaz, T.G.; Cabanillas, A.G.; Valenzuela, M.I.A.; Correa, C.A.; Salinas, F. Determination of nitrofurantoin, furazolidone, and furaltadone in milk by high-performance liquid chromatography with electrochemical detection. J. Chromatogr. A 1997, 764, 243–248. [Google Scholar] [CrossRef] [PubMed]
  2. Mariappan, K.; Chen, T.W.; Chen, S.M.; Tseng, T.W.; Bian, Y.; Sun, T.T.; Jiang, J.; Yu, J. Fabrication of Hexagonal CuCoO2 Modified Screen-Printed Carbon Electrode for the Selective Electrochemical Detection of Furaltadone. Int. J. Electrochem. Sci. 2022, 17, 220644. [Google Scholar] [CrossRef]
  3. Vasu, D.; Keyan, A.K.; Sakthinathan, S.; Chiu, T.W. Investigation of electrocatalytic and photocatalytic ability of Cu/Ni/TiO2/MWCNTs Nanocomposites for detection and degradation of antibiotic drug Furaltadone. Sci. Rep. 2022, 12, 886. [Google Scholar] [CrossRef] [PubMed]
  4. Amalraj, A.J.J.; Murthy, U.N.; Fue, W.S. Ultrasensitive electrochemical detection of an antibiotic drug Furaltadone in fish tissue with a ZnO-ZnCo2O4 self-assembled nano-heterostructure as an electrode material. Microchem. J. 2021, 169, 106566. [Google Scholar] [CrossRef]
  5. Massa, A. Clinical trial of furaltadone, an antibacterial nitrofuran. Br. J. Clin. Pract. 1961, 15, 351–354. [Google Scholar] [CrossRef]
  6. Khodari, M.; Din, H.S.E.; Mersal, G.A.M. Electroreduction and quantification of furazolidone and furaltadone in different media. Mikrochim. Acta 2000, 135, 9–17. [Google Scholar] [CrossRef]
  7. Barbosa, J.; Freitas, A.; Moura, S.; Maria, J.L.M.; Silveira, I.N.; Ramos, F. Detection, accumulation, distribution, and depletion of furaltadone and nifursol residues in poultry muscle, liver, and gizzard. J. Agric. Food Chem. 2011, 59, 11927–11934. [Google Scholar] [CrossRef]
  8. Xu, Z.L.; Shen, Y.D.; Sun, Y.M.; Campbell, K.; Tian, Y.X.; Zhang, S.W.; Lei, H.T.; Jiang, Y.M. Novel hapten synthesis for antibody production and development of an enzyme-linked immunosorbent assay for determination of furaltadone metabolite 3-amino-5-morpholinomethyl-2-oxazolidinone (AMOZ). Talanta 2013, 103, 306–313. [Google Scholar] [CrossRef] [Green Version]
  9. Jester, E.L.E.; Abraham, A.; Wang, Y.; Steven, K.R.E.S.; Plakas, M. Performance evaluation of commercial ELISA kits for screening of furazolidone and furaltadone residues in fish. Food Chem. 2014, 145, 593–598. [Google Scholar] [CrossRef]
  10. Shi, S.; Cao, G.; Chen, Y.; Huang, J.; Tang, Y.; Jiang, J.; Gan, T.; Wan, C.; Wu, C. Facile synthesis of core-shell Co-MOF with hierarchical porosity for enhanced electrochemical detection of furaltadone in aquaculture water. Anal. Chim. Acta 2023, 1263, 341296. [Google Scholar] [CrossRef]
  11. Alawi, M.A. Analysis of furazolidone and furaltadone in chicken tissues and eggs using a modified HPLC/ELCD method. Fresenius Environ. Bull. 2000, 9, 508–514. [Google Scholar]
  12. Balamurugan, K.; Rajakumaran, R.; Chen, S.M.; Karthik, R.; Shim, J.J.; Shafi, P.M. Massive engineering of spinel cobalt tin oxide/tin oxide-based electrocatalyst for the selective voltametric determination of antibiotic drug furaltadone in water samples. J. Alloys Compd. 2021, 882, 160750. [Google Scholar] [CrossRef]
  13. Kokulnathan, T.; Wang, T.J.; Ahmed, F.; Alshahrani, T. Synthesis of 3D flower-like zinc-chromium layered double hydroxides: A functional electrode material for furaltadone detection. Process Saf. Environ. Prot. 2023, 176, 889–897. [Google Scholar] [CrossRef]
  14. Keyan, A.K.; Sakthinathan, S.; Vasu, D.; Yu, C.L.; Vinothini, S.; Chiu, T.W. Gadolinium molybdate decorated graphitic carbon nitride composite: Highly visualized detection of nitrofurazone in water samples. RSC Adv. 2022, 12, 34066–34079. [Google Scholar] [CrossRef]
  15. Keyan, A.K.; Vasu, D.; Sakthinathan, S.; Chiu, T.W.; Lee, Y.H.; Lin, C.C. Facile synthesis of silver-doped copper selenide composite for enhanced electrochemical detection of ecological toxic nitrobenzene. Electrocatalysis 2023, 14, 448–462. [Google Scholar] [CrossRef]
  16. Sakthinathan, S.; Keyan, A.K.; Vasu, D.; Vinothini, S.; Nagaraj, K.; Mangesh, V.L.; Chiu, T.W. Graphitic carbon nitride incorporated europium molybdate composite as an enhanced sensing platform for electrochemical detection of carbendazim in agricultural products. J. Electrochem. Soc. 2022, 169, 127504. [Google Scholar] [CrossRef]
  17. Enayat, M.J. Controllable fabrication of europium tungstate nanoparticles in the presence of new capping agents and investigation of their photocatalytic properties. J. Mater. Sci. Mater. Electron. 2017, 28, 16444–16449. [Google Scholar] [CrossRef]
  18. Nasrabadi, M.R.; Pourmohamadian, V.; Karimi, M.S.; Naderi, H.R.; Karimi, M.A.; Didehban, K.; Ganjali, M.R. Assessment of supercapacitive performance of europium tungstate nanoparticles prepared via hydrothermal method. J. Mater. Sci. Mater. Electron. 2017, 28, 12391–12398. [Google Scholar] [CrossRef]
  19. Templeton, D.H.; Zalkin, A. Crystal Structure of Europium Tungstate. Acta Cryst. 1963, 16, 762. [Google Scholar] [CrossRef] [Green Version]
  20. Naderi, H.R.; Nadri, P.; Sorouri, A.M.; Ganjali, M.R.; Ehrlich, H.; Nasrabadi, M.R.; Ahmadi, F. Evaluation of electrodes composed of europium tungstate/reduced graphene oxide nanocomposite for use as supercapacitors. Surf. Interfaces 2022, 31, 102002. [Google Scholar] [CrossRef]
  21. Kodairaa, C.A.; Britoa, H.F.; Maltab, O.L.; Serrac, O.A. Luminescence and energy transfer of the europium (III) tungstate obtained via the Pechini method. J. Lumin. 2003, 101, 11–21. [Google Scholar] [CrossRef]
  22. Karthikeyan, D.; Vijayakumar, K.; Suhasini, P.; Dhanusha, A.; Sabari, G.T.C. Third-order nonlinear optical responses of Cr3+ ions on La2(WO4)3 nanoparticles for optical limiting applications. J. Photochem. Photobiol. A Chem. 2023, 436, 114377. [Google Scholar] [CrossRef]
  23. Yun, F.; Cao, L.; Huang, J.; Wu, J. Effects of pH on the microstructures and optical property of FeWO4 nanocrystallites prepared via hydrothermal method. Ceram. Int. 2013, 39, 4133–4138. [Google Scholar]
  24. Kim, A.H.; Kang, J.H.; Hong, A.R.; Park, Y.J.; Jang, H.S.; Kang, Y.M.; Kim, D.H. Sputter-grown Eu-doped WO3-Eu2(WO4)3 composite red phosphor thin films. Opt. Mater. 2021, 122, 111721. [Google Scholar] [CrossRef]
  25. Kubendhiran, S.; Sakthinathan, S.; Chen, S.M.; Lee, C.M.; Lou, B.S.; Sireesha, P.; Su, C. Electrochemically activated screen printed carbon electrode decorated with nickel nano particles for the detection of glucose in human serum and human urine sample. Int. J. Electrochem. Sci. 2016, 11, 7934–7946. [Google Scholar] [CrossRef]
  26. Rajakumaran, R.; Babulal, S.M.; Chen, S.M.; Sukanya, R.; Karthik, R.; Shafi, P.M.; Shim, J.J.; Shiuan, C.Y. Ingenious design of iron vanadate engulfed 3D porous reduced graphene oxide nanocomposites as a reliable electrocatalyst for the selective amperometric determination of furaltadone in aquatic environments. Appl. Surf. Sci. 2021, 569, 151046. [Google Scholar] [CrossRef]
  27. Horne, E.; Cadogan, A.; Keeffe, M.O.; Boom, L.A.P.H. Analysis of Protein-bound metabolites of Furazolidone and furaltadone in pig liver by high-performance liquid chromatography and liquid chromatography–mass spectrometry. Analyst 1996, 121, 1463–1468. [Google Scholar] [CrossRef]
  28. Yan, C.; Teng, J.; Liu, F.; Yao, B.; Xu, Z.; Yao, L.; Chen, W. Signal amplified enzyme-linked immunosorbent assay with gold nanoparticles for sensitive detection of trace furaltadone metabolite. Microchem. J. 2020, 159, 105414. [Google Scholar] [CrossRef]
  29. Barbosa, J.; Freitas, A.; Mourao, J.L.; Silveira, M.I.N.; Ramos, F. Determination of furaltadone and nifursol residues in poultry eggs by liquid chromatography-electrospray ionization tandem mass spectrometry. J. Agric. Food Chem. 2012, 60, 4227–4234. [Google Scholar] [CrossRef]
  30. Bock, C.; Gowik, P.; Stachel, C. Matrix-comprehensive in-house validation and robustness check of a confirmatory method for the determination of four nitrofuran metabolites in poultry muscle and shrimp by LC-MS/MS. Biomed Life Sci. 2007, 856, 178–189. [Google Scholar] [CrossRef]
Figure 1. (A) XRD spectra of Eu2(WO4)3 and (B) crystal structure of Eu2(WO4)3.
Figure 1. (A) XRD spectra of Eu2(WO4)3 and (B) crystal structure of Eu2(WO4)3.
Technologies 11 00105 g001
Figure 2. (A,B) FESEM image of Eu2(WO4)3. (C) Eu2(WO4)3 elemental mapping results of (D) Eu, (E) W, and (F) O. (G) shows the EDX spectra of Eu2(WO4)3.
Figure 2. (A,B) FESEM image of Eu2(WO4)3. (C) Eu2(WO4)3 elemental mapping results of (D) Eu, (E) W, and (F) O. (G) shows the EDX spectra of Eu2(WO4)3.
Technologies 11 00105 g002
Figure 3. (A) CV response of (a) SPCE, (b) ASPCE, (c) SPCE/Eu2(WO4)3, (d) ASPCE/Eu2(WO4)3 electrodes in 0.05 M PBS (pH 7) with 100 μM of FTD and (e) ASPCE/Eu2(WO4)3 in absence of FTD. (B) Bar diagram of the peak current of SPCE, ASPCE, SPCE/Eu2(WO4)3, and ASPCE/Eu2(WO4)3 electrodes.
Figure 3. (A) CV response of (a) SPCE, (b) ASPCE, (c) SPCE/Eu2(WO4)3, (d) ASPCE/Eu2(WO4)3 electrodes in 0.05 M PBS (pH 7) with 100 μM of FTD and (e) ASPCE/Eu2(WO4)3 in absence of FTD. (B) Bar diagram of the peak current of SPCE, ASPCE, SPCE/Eu2(WO4)3, and ASPCE/Eu2(WO4)3 electrodes.
Technologies 11 00105 g003
Scheme 1. Possible electrochemical detection (reduction) mechanism of FLT.
Scheme 1. Possible electrochemical detection (reduction) mechanism of FLT.
Technologies 11 00105 sch001
Figure 4. (A) Cyclic voltammogram of ASPCE/Eu2(WO4)3 in 0.05 M PBS (pH 7) containing different concentrations of FLD at a scan rate of 50 mVs−1 (Different colors are mentioned for different concentration of FLD peak currents). (B) Correspondences to the calibration plot for FLD concentrations (100–700 µM) vs. peak current.
Figure 4. (A) Cyclic voltammogram of ASPCE/Eu2(WO4)3 in 0.05 M PBS (pH 7) containing different concentrations of FLD at a scan rate of 50 mVs−1 (Different colors are mentioned for different concentration of FLD peak currents). (B) Correspondences to the calibration plot for FLD concentrations (100–700 µM) vs. peak current.
Technologies 11 00105 g004
Figure 5. (A) Effect of different scan rates at ASPCE/Eu2(WO4)3 electrode on the reduction of FLD (100 μM) present in PBS (pH 7) (Different colors are mentioned for different scan rates peak currents). (B) Calibration graph for the reduction of FLD between scan rate and peak current.
Figure 5. (A) Effect of different scan rates at ASPCE/Eu2(WO4)3 electrode on the reduction of FLD (100 μM) present in PBS (pH 7) (Different colors are mentioned for different scan rates peak currents). (B) Calibration graph for the reduction of FLD between scan rate and peak current.
Technologies 11 00105 g005
Figure 6. (A) Cyclic voltammetry performance of FTD reduction at different pH from 5 to 11. (B) FTD reduction peak current vs. pH (Different colors are mentioned for different pH peak currents). (C) Effects of pH on the FTD reduction peak potential vs. pH.
Figure 6. (A) Cyclic voltammetry performance of FTD reduction at different pH from 5 to 11. (B) FTD reduction peak current vs. pH (Different colors are mentioned for different pH peak currents). (C) Effects of pH on the FTD reduction peak potential vs. pH.
Technologies 11 00105 g006
Figure 7. (A) DPV response of ASPCE/Eu2(WO4)3 electrode for different concentrations of FLD in PBS (pH 7) (Different colors are mentioned for different concentration of FLD peak currents). (B) calibration plot of the different concentrations of FLD vs. reduction peak current.
Figure 7. (A) DPV response of ASPCE/Eu2(WO4)3 electrode for different concentrations of FLD in PBS (pH 7) (Different colors are mentioned for different concentration of FLD peak currents). (B) calibration plot of the different concentrations of FLD vs. reduction peak current.
Technologies 11 00105 g007
Figure 8. (A) Repeatability study of SPCE/Eu2(WO4) electrodes, (B) reproducibility study of SPCE/Eu2(WO4) electrodes, (C) cyclic stability study of the RRD/Eu2(WO4) electrode, (D) operational stability study of the RRD/Eu2(WO4)3 electrode toward FTD detection in 0.05 M PBS (pH 7) up to 2000 s.
Figure 8. (A) Repeatability study of SPCE/Eu2(WO4) electrodes, (B) reproducibility study of SPCE/Eu2(WO4) electrodes, (C) cyclic stability study of the RRD/Eu2(WO4) electrode, (D) operational stability study of the RRD/Eu2(WO4)3 electrode toward FTD detection in 0.05 M PBS (pH 7) up to 2000 s.
Technologies 11 00105 g008
Figure 9. Real sample analysis of FLD in (A) silverzine, (B) sindine, and (C) tetracycline medicine samples.
Figure 9. Real sample analysis of FLD in (A) silverzine, (B) sindine, and (C) tetracycline medicine samples.
Technologies 11 00105 g009
Table 1. Comparison of the analytical performance of ASPCE/Eu2(WO4)3 electrode with other FTD sensors.
Table 1. Comparison of the analytical performance of ASPCE/Eu2(WO4)3 electrode with other FTD sensors.
Electrode and TechniqueLinear RangeLimit of DetectionReferences
FeVO/p-rGO NCs0.584138 nM[26]
AOZ/AMOZ0.05–2.010.0 μg kg−1[27]
LC-MS/MSc1–8000.5 μg kg−1[28]
ELISA0.9–105.30.3 μg kg−1[29]
LC-MS/MS0.98–0.990.2 μg/kg[30]
ASPCE/Eu2(WO4)3 10 nM–300 µM97 µMThis work
Table 2. Determination of FTD using different antibiotics medicine by ASPCE/Eu2(WO4)3 electrode.
Table 2. Determination of FTD using different antibiotics medicine by ASPCE/Eu2(WO4)3 electrode.
Real SamplesAddedFoundRecovery (%)
Silverzine medicine10990
2018.592.5
3029.297.3
Sindine medicine108.585
201995
3029.498
Tetracycline medicine109.393
2019.396.5
3029.598.3
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Vinothini, S.; Chiu, T.-W.; Sakthinathan, S. Electrochemical Detection of Furaltadone Antibiotic Drug by the Rare Earth Metal Tungstate Decorated Screen Printed Carbon Electrode. Technologies 2023, 11, 105. https://doi.org/10.3390/technologies11040105

AMA Style

Vinothini S, Chiu T-W, Sakthinathan S. Electrochemical Detection of Furaltadone Antibiotic Drug by the Rare Earth Metal Tungstate Decorated Screen Printed Carbon Electrode. Technologies. 2023; 11(4):105. https://doi.org/10.3390/technologies11040105

Chicago/Turabian Style

Vinothini, Sivaramakrishnan, Te-Wei Chiu, and Subramanian Sakthinathan. 2023. "Electrochemical Detection of Furaltadone Antibiotic Drug by the Rare Earth Metal Tungstate Decorated Screen Printed Carbon Electrode" Technologies 11, no. 4: 105. https://doi.org/10.3390/technologies11040105

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