Next Article in Journal
A Hybrid Controller for a Soft Pneumatic Manipulator Based on Model Predictive Control and Iterative Learning Control
Next Article in Special Issue
Biomarkers and Corresponding Biosensors for Childhood Cancer Diagnostics
Previous Article in Journal
Design of Laboratory Stand for Displacement Measurement of IPMC Actuators
Previous Article in Special Issue
Investigation of GOx Stability in a Chitosan Matrix: Applications for Enzymatic Electrodes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

An Electrochemical Sensor for Sulfadiazine Determination Based on a Copper Nanoparticles/Molecularly Imprinted Overoxidized Polypyrrole Composite

by
Manahil Babiker Elamin
1,
Shazalia Mahmoud Ahmed Ali
1,
Houda Essousi
2,
Amani Chrouda
1,
Laila M. Alhaidari
1,
Nicole Jaffrezic-Renault
3,* and
Houcine Barhoumi
2
1
Department of Chemistry, Faculty of Science Al-Zulfi, Majmaah University, Majmaah 11952, Saudi Arabia
2
Laboratory of Interfaces and Advanced Materials, Faculty of Sciences, Monastir University, Monastir 5000, Tunisia
3
Institute of Analytical Sciences, University of Lyon, UMR CNRS 5280, 69100 Villeurbanne, France
*
Author to whom correspondence should be addressed.
Sensors 2023, 23(3), 1270; https://doi.org/10.3390/s23031270
Submission received: 4 January 2023 / Revised: 12 January 2023 / Accepted: 18 January 2023 / Published: 22 January 2023
(This article belongs to the Special Issue Feature Papers in Biosensors Section 2022)

Abstract

:
To protect consumers from risks related to overexposure to sulfadiazine, total residues of this antibacterial agent in animal-origin foodstuffs not exceed international regulations. To this end, a new electrochemical sensor based on a molecularly imprinted polymer nanocomposite using overoxidized polypyrrole and copper nanoparticles for the detection of sulfadiazine is elaborated. After optimization of the preparation of the electrochemical sensors, their differential pulse voltammetric signal exhibits an excellent stability and reproducibility at 1.05 V, with a large linear range between 10−9 and 10−5 mol L−1 and a low detection limit of 3.1 × 10−10 mol L−1. The produced sulfadiazine sensor was successfully tested in real milk samples. The combination of the properties of the electrical conduction of copper nanoparticles with the properties of the preconcentration of the molecularly imprinted overoxidized polypyrrole allows for the highly sensitive detection of sulfadiazine, even in real milk samples. This strategy is new and leads to the lowest detection limit yet achieved, compared to those of the previously published sulfadiazine electrochemical sensors.

1. Introduction

Sulfadiazine (4-amino-N-2-pyrimidinylbenzenesulfonamide, SDZ) is an antibacterial agent widely used in human and veterinary medicines to treat infectious diseases such as ear infections, urinary tract diseases, toxoplasmosis, pneumonia, and malaria [1,2]. The over exposure to SFX can cause physiological problems such as liver damage, antibiotic resistance, and gene damage. Thus, to protect consumers from risks related to SDZ residues, the European Union Council Regulations set a safe limit stating that the combined total residues of SDZ must not exceed 100 ng/g in animal-origin foodstuffs [3]. Following international regulation, the maximum allowable concentration of SDZ in milk is 0.07 ppm [4]. Therefore, the determination of trace level of SDZ in milk and other animal products used for human consumption has become an important and urgent task. Previously, various techniques have been used for SDZ detection, such as high-performance liquid chromatography, ultraviolet [5,6], UV-Visible [7,8], and fluorescence [9] spectroscopies, and other hyphenated techniques [10,11,12]. The principal limitations of these techniques are their high costs, the requirement of a long preparation time, and the necessity of sophisticated technical skills. However, electrochemical sensors based on molecular imprinting polymer (MIP) have gained more attention in recent years, due to their simple manipulation, high selectivity, and portability [13,14,15,16]. MIP is formed by the polymerization or electropolymerization of functional monomer in the presence of the target analyte (template molecule), followed by the extraction of the template molecule from the polymer matrix via washing steps, leaving specific cavities which are complementary to the template molecule [17,18,19,20]. MIP has been widely applied in pharmaceutical analysis [21,22], food analysis [23], and environmental analysis [24]. A molecularly imprinted polyacrylate-based sensor for SDZ detection was developed by Liu et al. [25]. The determination of SDZ through differential pulse voltammetry was obtained in the range of 4 to 50 × 10−6 M, with a detection limit of 0.68 × 10−6 M. The method was successfully applied to determine sulfadiazine in milk and milk powder. A carbon paste electrode combined with polyacrylate MIP was used for the determination of sulfadiazine in milk and human serum samples by Sadeghi et al. [26]; the obtained dynamic range was from 2.0 × 10−7 M to 1.0 × 1 0−4 M, and the detection limit was 1.4 × 10−7 M.
Conductive polymer polypyrrole (PPy) is widely employed in chemical sensors and biosensors [27,28]. PPy has many advantages, such as being easy to prepare through electropolymerization, having good redox properties, being environmentally stable, and exhibiting good conductivity and good biocompatibility [20]. PPy can be overoxidized in a sodium hydroxide (NaOH) solution during several scans of cyclic voltammetry. This process plays a positive role during the formation of MIP since it introduces oxygen groups (carboxyl, carbonyl, and hydroxyl) into the OPPy backbone [29,30], which could be suitable for the recognition of sulfadiazine positive groups. However, this process makes the polymer film less conductive. To compensate for this drawback, we propose the subsequent electrodeposition of copper nanoparticles (CuNPs) that will help to enhance the film conductivity. CuNPs exhibit good conductivity, high chemical and mechanical stability, and a high specific surface area that makes them suitable for electrochemical sensors. Some researchers have proven that copper nanoparticles electrochemically deposited on the MIP surface are useful in the sensitive detection of drugs and pollutants [31,32,33].
Many electrochemical sensors were designed for the detection of sulfadiazine [25,26,34,35,36,37]. Some sensors were based on direct electrochemical detection on different types of electrode materials: glassy carbon [38], boron-doped diamond [39], or bismuth film [38]. Different nanomaterials were used for enhancing the sensitivity of the detection of SDZ: rare earth compounds [36,40,41], phthalocyanines [42,43], manganese and zinc oxides [44,45,46,47], vanadium and bismuth sulfides [38,41], metallic nanoparticles [40], and carbonaceous nanomaterials [38,39,42,44,48,49,50,51]. In addition to the various MIP-based sensors designed, different polymers were also used for MIP fabrication: polyacrylate [25,26], PEDOT [34], and PPy [35]. PPy was overoxidized in an NaOH solution, and a graphene oxide@covalent organic framework (GO@COF) nanocomposite was added for signal amplification of the MIP film. In this work, the unique conductive properties of CuNPs were combined for the first time with the molecularly imprinted OPPy film for sulfadiazine electrochemical determination. The analytical performance of the optimized CuNPs/MIP-OPPy/GCE sensor was evaluated by differential pulse voltammetry and validated for its detection in real spiked milk samples.

2. Materials and Methods

2.1. Chemical Reagents

Pyrrole was purchased from Merck, Germany, and then purified through distillation. Copper sulfate, sodium nitrate, sulfadiazine, acetone, sulfuric acid, hydrochloric acid, dapsone, sulfamethoxazol, dopamine, ascorbic acid, and uric acid were obtained from Sigma Aldrich, Germany. Hexacyanoferrate (II/III) was obtained from FlukaChemika.

2.2. Instrumentation

The samples were characterized by Fourier transform infrared spectroscopy (FT-IR) using a Perkin Elmer 1600 FTIR spectrometer (Perkin Elmer Waltham, MA, USA) and scanned over the wave number range between 400 and 4000 cm−1.
The surface morphology of the modified electrodes was characterized by the use of a ZEISS EVO MA 25 (Carl Zeiss S.A.S. Rueil-Malmaison, France) scanning electron microscope (SEM). HR-TEM images were obtained using a JEM-2100 analytical electron microscope (JEOL Europe S.A.S., Croissy-sur-Seine, France) operating at an accelerating voltage of 200 kV. For HR-TEM images, the CuNPs/MIP-OPPy composite was transferred from the electrode surface to the TEM grid.
Electrochemical measurements were carried out using an Autolab electrochemical analyzer, model PGSTAT 302N (Eco Chemie, Utrecht, The Netherlands). A three-electrode system was employed, consisting of a glassy carbon working electrode (GCE, diameter 3 mm), an Ag/AgCl/KCl(3M) electrode used as a reference electrode, and a platinum electrode used as a counter electrode. Electrochemical experiments were controlled by using Nova version 1.5 software. Cyclic voltammetry was carried out at a scan rate of 100 mV/s. For the detection of SDZ in 0.1 M H2SO4 solution, the parameters for DPV (differential pulse voltammetry) were: increment—5 mV; amplitude—50 mV; sample period—100 ms; pulse time—10 ms.

2.3. Preparation of the SDZ Imprinted Polypyrrole Film

Before use, the GCE was cleaned according to the procedure presented in [52]. The modification process consisted of the electropolymerization of 0.1 M pyrrole monomer in the presence of sulfadiazine in a 0.1 M H2SO4 solution, using cyclic voltammetry at a scan rate of 100 mV/s, in the potential range from −0.2 to +1.4 V for 10 cycles. After electropolymerization, the sensor was rinsed with water and dried at room temperature. The recognition sites in the imprinted matrix were formed after removing the template from the imprinted film. For this purpose, the film was overoxidized by cyclic voltammetry (CV) in a 0.1 M NaOH solution, until current stabilization was reached, leaving imprints in the polymer film corresponding to the template shape. This process made the surface of the electrode less conductive. The loss of PPy electroactivity was compensated for by incorporating CuNPs into the film through cyclic voltammetry. CV curves of MIP-OPPy/GCE in the presence of CuSO4 were obtained, under different scanning cycles (from 1 to 30), in a 0.1 M Na2SO4 solution containing 10−3 M CuSO4. After the electrodeposition of the CuNPs on the modified electrode, it was rinsed with deionized water and dried. The procedure for the fabrication of CuNPs/MIP-OPPy/GCE is schematically presented in Scheme 1. Respectively, the non-imprinted polymer membrane (NIP) was also fabricated in the same way, but without SDZ.

2.4. Determination of Sulfadiazine in Real Samples

Reduced-fat milk samples were purchased from Zulfi market. Appropriate quantities of SDZ were spiked in these samples. A total of 25 mL of milk were centrifuged at 2000 rpm for 30 min to precipitate fat. A total of 10 mL of acetonitrile was added to the supernatant for its deproteinization. The mixture was then centrifuged at 4000 rpm for 15 min, after which the supernatant was then mixed with 25 mL of H2SO4 solution (pH 3) [53].

3. Results and Discussion

3.1. Elaboration of CuNPs/MIP-OPPy Sensor and NIP Polypyrrole Film

3.1.1. Electropolymerization of MIP and NIP Polypyrrole Film

The cyclic voltammograms obtained during the electropolymerization of the MIP polypyrrole film (in the presence of SDZ) and of the NIP polypyrrole film (in the absence of SDZ) are presented in Figure 1. It appears that the exchanged charge quantity is lower for MIP compared to that of NIP.
The SEM image of bare GCE (Figure 2A) is used as the control for comparison. The surface of NIP/GCE (Figure 2B) shows a uniform formation of a polypyrrole film. From Figure 2C, the thickness of MIP in the presence of SDZ was less important than that of NIP under identical conditions, in agreement with the exchanged charge quantity.

3.1.2. Template Extraction through Overoxidation of MIP Polypyrrole Film

Using organic reagents as the eluent for the template removal requires abundant organic reagents and is also time-consuming. In addition, it is very difficult to completely remove the template in a short time. The elimination of the cationic charge in the polymer backbone by further oxidation in high pH conditions (overoxidation) allows the template to be removed from the overoxidized PPy film, and this procedure does not require any organic reagents. CV is used for the overoxidation of the PPy film in the 0.1 M NaOH solution; the cyclic voltammogram is shown in Figure 3. Comparing the first cycle to the fourth, the peaks at −0.01 V and 0.85 V disappeared in the fourth cycle, which is related to the oxidation process of PPy. At the same time, the anodic peak ascribed to SDZ oxidation at 1.1 V, through the reaction presented in Scheme 2, also disappeared. After SDZ elution from the PPy film, through the oxidation reaction, a molecularly imprinted film is fabricated. The 6th and 7th cyclic voltammograms are almost identical, which proves that PPy is completely oxidized. After overoxidation, a porous structure appears on the surface of MIP-OPPy, as shown in the SEM image (Figure 2D).

3.1.3. Electrodeposition of Copper Nanoparticles

The cyclic voltammograms obtained during the electroreduction of copper are presented in Figure 4. The reduction peak of copper occurs at −0.25 V. This CV curve is similar to that obtained for the electrodeposition of CuNPs on pectin [54]. In this work, the authors obtained crystalline copper with (111), (200) and (220) diffraction peaks. The EDX spectra of the composite film is presented in Figure 5A, showing that copper is a major compound of this composite. The obtained copper nanoparticles were observed using HR-TEM, as presented in Figure 5B. Their diameter is between 3 and 4 nm.

3.2. FTIR Characterization

Figure 6 shows the FTIR spectra of the SDZ powder, PPy/SDZ, and MIP-OPPy matrices. The absorption peaks observed at 3351 and 3422 cm−1 for sulfadiazine powder are attributed to the bending vibration of primary NH2 symmetric and asymmetric stretching, respectively [55,56]. The peaks at 1327 and 1581 cm−1 are assigned to C=N and -CNS of the SDZ molecules, respectively (Figure 6a, Table 1) [57]. The peaks observed at 666 and 785 cm−1 are assigned to the asymmetric stretching vibration of the SO2 group [58]. Before the removal of the sulfadiazine molecule, the FTIR spectrum of PPy/SDZ (Figure 6b, Table 1) shows two weak adsorption peaks at 672 and 785 cm−1, which are attributed to the asymmetric vibration of SO2-N in the SDZ molecules [59]. It indicates that the template SDZ molecule was successfully entrapped in the polymer film through hydrogen bonds. The absorption bands of the template molecules disappeared in the MIP-OPPy spectrum, which can be attributed to the ejection of SDZ from the film (Figure 6c, Table 1). Thus, the FTIR characterization has been efficiently adapted to confirm that the SDZ molecules were successfully entrapped and removed from the polymer structure. On the other hand, the FTIR spectra of PPy/SDZ and MIP-OPPy possess the same absorption bands, differing only in intensity. The absorption peaks at 2910 and 2980 cm−1 are attributed to the symmetric and asymmetric N-H stretching of the amine group of the pyrrole group. The band at 1750 cm−1 is attributed to the symmetric stretching vibration of C-H.

3.3. Effect of the Experimental Conditions for SDZ Detection

To maximize the sensitivity of the CuNPs/MIP-OPPy/GCE sensor for the determination of SDZ, some experimental parameters needed to be optimized, including the number of electropolymerization scan cycles, the concentrations of sulfadiazine and the monomer, incubation time, and pH value. The DPV signal of SDZ in a 0.1 M H2SO4 solution (pH = 2) containing 10−6 mol L−1 SDZ was reported as the response of the MIP sensor to the different parameters.

3.3.1. Effect of Py Monomer and SDZ Template Concentrations

The Py monomer concentration during the electrodeposition process can influence the electrochemical response of the MIP sensor [60]. Different films were grown from solutions containing a constant concentration of SDZ (10−3 M) and with varied concentrations of the pyrrole monomer, between 1 mM and 200 mM. Figure 7A reports the variation in the MIP sensor response for the different pyrrole concentrations. The response increases when increasing the pyrrole concentration up to 0.1 mol L−1 and for higher concentrations, the response decreases. We can deduce that the optimum monomer concentration is about 0.1 mol L−1.
The effect of the SDZ concentration in the electropolymerization solution on the MIP sensor response was further examined. The concentrations of SDZ were between 10−4 mol L−1 and 10−2 mol L−1. As shown in Figure 7B, the MIP sensor response increases with the increase in the SDZ concentration between 10−4 mol L−1 and 10−3 mol L−1. When the template concentration was higher than 10−3 mol L−1, the current decreased. According to this result, the optimum template concentration was chosen to be 10−3 mol L−1.

3.3.2. Effect of the Number of Electropolymerization Cycles

Beyond the concentrations of the monomer and the template, another important parameter that may affect the electroanalytical performance is the MIP film thickness, controlled by the number of scan cycles. Figure 8 shows the effect of the number of scan cycles on the MIP sensor response. It can be seen that the current response first increases, reaching a maximum at 15 cycles, then decreases for a larger number of cycles. A total of 15 cycles were then selected to elaborate the SDZ MIP sensor.

3.3.3. Effect of the Incubation Time

The incubation process is an important parameter that influences the sensitivity of the MIP sensor [61]. Incubation times of the MIP sensor varied from 5 to 25 min. Figure 9 shows that the MIP sensor response increases gradually with the increasing incubation time, and when the incubation time is higher than 15 min, the MIP sensor response remains almost stable, suggesting that the SDZ adsorption equilibrium is reached. The optimal incubation time was then selected as 15 min.

3.3.4. Effect of the Scan Rate

The effect of the scan rate on the oxidation peak current of SDZ was investigated by cyclic voltammetry, in the presence of 10−6 mol L−1 SDZ in 0.1 mol L−1 H2SO4 (pH = 2), with scan rates ranging from 20 to 200 mV/s (Figure 10). The oxidation peak currents increased with an increase in the scan rate, while their oxidation peak potentials gradually shifted to positive values. A plot of Ipa versus the square root of the scan rate (v1/2) in the range of 20 to 200 mV/s yielded a straight line, with a good correlation coefficient (R2 = 0.99235). This relationship transduces a diffusion-controlled oxidation reaction.

3.3.5. Effect of the pH Value

Figure 11A shows the effect of pH on the response of the MIP sensor within the pH range of 2.0–7.0. The potential of the oxidation peak is shifted in the positive direction when the pH increases, which is in agreement with the oxidation reaction presented in Scheme 2. The current decreases when pH levels range from pH 2.0 to higher pH values (Figure 11B). Based on these results, a pH equal to 2.0 was chosen for the detection of SDZ. The best affinity of SDZ for the MIP-OPPy film is obtained in solutions with a pH value of 2, showing that the main interactions are hydrogen bonds [60].

3.4. Analytical Performance of the MIP Sensor

Figure 12A shows the differential pulse voltammetry (DPV) curves recorded for different SDZ concentrations in 0.1 mol L−1 H2SO4 (pH = 2). The DPV curves are recorded in the range of 10−9 to 10−5 mol L−1. It can be seen that the peak current increased with increasing SDZ concentrations. Figure 12B shows the calibration curve as the peak current versus the -log ([SDZ]) recorded for the MIP sensor and the NIP sensor. The detection limit LOD of the MIP sensor is 3.1 × 10−10 mol L−1, according to the formula 3σ/S, σ being the background of the blank, and S being the sensitivity. The linear relationship has been observed in the concentration range from 10−9 to 10−5 mol L−1, with a correlation coefficient of R2 = 0.98307 and R2 = 0.98574 for NIP and MIP sensors, respectively. A high analytical response was obtained with a sensitivity of 3.78 μA mol L−1 for the MIP electrode, in comparison with the NIP electrode, with a sensitivity of 0.77 μA mol L−1. The imprinting factor of the MIP sensor is 4.9. The sensitivity of MIP without CuNPs was 1.387 μA mol L−1. The integration of CuNPs into the MIP sensor enhances the conductivity of the MIP film, which in turn can dramatically improve the sensitivity of the imprinted sensor. Compared with previously published works on electrochemical SDZ sensors, the CuNPs/MIP-OPPy/GCE sensor designed in this study presents the lowest detection limit and the largest dynamic range (Table 2).
Four sensors were prepared independently, on the same day and under the same optimal conditions, to detect 10−6 mol L−1 SDZ in the H2SO4 solution (pH 2). The relative standard deviations (RSD, n = 4) of the DPV measurements was 2.1%, which indicates the excellent reproducibility of the CuNPs/MIP-OPPy/GCE-based sensor. The stability of the developed sensor was also tested. The DPV response of the MIP sensor was measured after an interval of five days. As shown in Figure 13, the MIP sensor exhibits only a slight decrease (2.4%) in the DPV signal after 25 days, demonstrating high stability and good repeatability.
The response of the MIP sensor and of the NIP sensor for different interfering substances of SDZ was measured. The MIP and NIP sensors were introduced into individual solutions of sulfadiazine (10−6 mol L−1, pH 2) and of some analog substances, such as sulfamethoxazol (SFD), dapsone (DDS), dopamine (DA), uric acid (UA), and ascorbic acid (AA) at 10−6 mol L−1 concentration, and the DPV signal was measured. The responses are reported in Figure 14. The SDZ DPV signal of the MIP sensor was 1.4 times higher than that of the SFD, and 3 times higher than that of the DDS, these molecules being analogs of SDZ; the specificity of the imprints in OPPy film is demonstrated. Nevertheless, due to its similarity of structure, SFD is partly recognized by the imprints in the MIP-OPPy film. DA, AU and AA are interfering substances in biological samples. The DPV signals of the MIP and NIP sensors are quite similar, showing that they present non-specific adsorption on both films.

3.5. Real Sample Analysis

Real sample analysis was performed by the MIP sensor in SDZ-spiked milk samples. The final concentrations in the spiked samples, after pretreatment, were first controlled using HPLC/MS/MS. The recoveries of the SDZ and RSD values for three determinations are summarized in Table 3. The unexpected results of an 83% recovery rate for one spiked milk sample can be explained by a fouling of the MIP surface by remaining proteins, which could hinder the adsorption of sulfadiazine on the MIP surface; thus, the concentration of sulfadiazine was underestimated. For two concentrations, the obtained recoveries indicate that the MIP sensor can be used successfully for the determination of SDZ in milk samples.

4. Conclusions

A novel and efficient electrochemical sensor based on a CuNPs/MIP-OPPy nanocomposite for SDZ detection has been designed. The performance of this sensor was evaluated by conducting a study of its electrochemical response to sulfadiazine in aqueous samples using differential pulse voltammetry. The MIP sensor offers a linear current response to SDZ in the range of 10−9 to 10−5 mol L−1, with a low limit of detection of 3.1 × 10−10 mol L−1. The results indicate that the incorporation of copper nanoparticles provides better sensitivity (3.78 μA mol L−1) for the MIP electrode, the sensitivity of the MIP sensor without CuNPs being 1.387 μA mol L−1. The MIP electrode exhibits good reproducibility, repeatability, and stability for more than 25 days for the determination of SDZ, and it was applied for SDZ quantification in spiked real milk samples.

Author Contributions

Conceptualization, H.B. and M.B.E.; validation, L.M.A.; investigation, H.E.; writing—original draft preparation, A.C.; writing—review and editing, N.J.-R.; project administration, S.M.A.A.; funding acquisition, A.C. All authors have read and agreed to the published version of the manuscript.

Funding

The authors extend their appreciation to the Deputyship for Research & Innovation, Ministry of Education, Saudi Arabia, for funding this research work through Project Number IFP-2022-45.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. García-Galán, M.J.; Díaz-Cruz, M.S.; Barceló, D. Identification and determination of metabolites and degradation products of sulphonamide antibiotics. TrAC Trends Anal. Chem. 2008, 27, 1008–1022. [Google Scholar] [CrossRef]
  2. Darde, M.L.; Fougere, E.; Buxeraud, J. Toxoplasmosis medication. Actual. Pharm. 2018, 57, 22–26. [Google Scholar] [CrossRef]
  3. Samanidou, V.F.; Tolika, E.P.; Papadoyannis, I.N. Chromatographic Residue Analysis of Sulfonamides in Foodstuffs of Animal Origin. Sep. Purf. Rev. 2008, 37, 327–373. [Google Scholar] [CrossRef]
  4. Long, A.R.; Short, C.R.; Barker, S.A. Method for the isolation and liquid chromatographic determination of eight sulfonamides in milk. J. Chromatogr. 1990, 502, 87–94. [Google Scholar] [CrossRef]
  5. Balakrishnan, V.K.; Terry, K.A.; Toito, J. Determination of sulfonamide antibiotics in wastewater: A comparison of solid phase microextraction and solid phase extraction methods. J. Chromatogr. A 2006, 1131, 1–10. [Google Scholar] [CrossRef] [PubMed]
  6. Zheng, M.M.; Zhang, M.Y.; Peng, G.Y.; Feng, Y.Q. Monitoring of sulfonamide antibacterial residues in milk and egg by polymer monolith microextraction coupled to hydrophilic interaction chromatography/mass spectrometry. Anal. Chim. Acta 2008, 625, 160–172. [Google Scholar] [CrossRef] [PubMed]
  7. Jacobsson, S.P.; Carlsson, M.; Jönsson, U.; Nilsson, G. Quantitative determination of sulfasalazine by near-infrared spectroscopy and multivariate analysis in reflectance mode with a fibre-optic probe. J. Pharm. Biomed. Anal. 1995, 13, 415–417. [Google Scholar] [CrossRef] [PubMed]
  8. Karimi, M.; Aboufazeli, F.; Zhad, H.R.L.Z.; Sadeghi, O.; Najafi, E. Determination of sulfonamides in chicken meat by magnetic molecularly imprinted polymer coupled to HPLC-UV. Food Anal. Methods 2014, 7, 73–80. [Google Scholar] [CrossRef]
  9. Zhang, Y.; Xu, X.; Qi, X.; Gao, W.; Sun, S.; Li, X.; Jiang, C.; Yu, A.; Zhang, H.; Yu, Y. Determination of sulfonamides in livers using matrix solid-phase dispersion extraction high-performance liquid chromatography. J. Sep. Sci. 2012, 35, 45–52. [Google Scholar] [CrossRef]
  10. Arroyo-Manzanares, N.; Gámiz-Gracia, L.; García-Campaña, A.M. Alternative sample treatments for the determination of sulfonamides in milk by HPLC with fluorescence detection. Food Chem. 2014, 143, 459–464. [Google Scholar] [CrossRef]
  11. Barfield, M.; Spooner, N.; Lad, R.; Parry, S.; Fowles, S. Application of dried blood spots combined with HPLC-MS/MS for the quantification of acetaminophen in toxicokinetic studies. J. Chromatogr. B 2008, 870, 32–37. [Google Scholar] [CrossRef] [PubMed]
  12. Chu, Q.; Jiang, L.; Tian, X.; Ye, J. Rapid determination of acetaminophen and p-aminophenol in pharmaceutical formulations using miniaturized capillary electrophoresis with amperometric detection. Anal. Chim. Acta 2008, 606, 246–251. [Google Scholar] [CrossRef] [PubMed]
  13. Zhu, L.; Cao, Y.; Cao, G. Electrochemical sensor based on magnetic molecularly imprinted nanoparticles at surfactant modified magnetic electrode for determination of bisphenol A. Biosens. Bioelectron. 2014, 54, 258–261. [Google Scholar] [CrossRef] [PubMed]
  14. Takeuchi, T.; Mukawa, T.; Matsui, J.; Higashi, M.; Shimizu, K.D. Molecularly Imprinted Polymers with Metalloporphyrin-Based Molecular Recognition Sites Coassembled with Methacrylic Acid. Anal. Chem. 2001, 73, 3869–3874. [Google Scholar] [CrossRef] [PubMed]
  15. Khalifa, M.E.; Abdallah, A.B. Molecular imprinted polymer based sensor for recognition and determination of profenofos organophosphorous insecticide. Biosens. Bioelectron. X 2019, 2, 100027. [Google Scholar] [CrossRef]
  16. Shang, H.; Xu, H.; Jin, L.; Wang, C.; Chen, C.; Song, T.; Du, Y. 3D ZnIn2S4 nanosheets decorated ZnCdS dodecahedral cages as multifunctional signal amplification matrix combined with electroactive/photoactive materials for dual mode electrochemical—Photoelectrochemical detection of bovine hemoglobin. Biosens. Bioelectron. 2020, 159, 112202. [Google Scholar] [CrossRef]
  17. Zang, X.; Peng, Y.; Bai, J.; Ning, B.; Sun, S.; Hong, X.; Yuanyuan, L.; Ying, L.; Gao, Z. A novel electrochemical sensor based on electropolymerized molecularly imprinted polymer and gold nanomaterials amplification for estradiol detection. Sens. Actuators B Chem. 2014, 200, 69–75. [Google Scholar] [CrossRef]
  18. Torkashvand, M.; Gholivand, M.B.; Malekzadeh, G. Construction of a new electrochemical sensor based on molecular imprinting recognition sites on multiwall carbon nanotube surface for analysis of ceftazidime in real samples. Sens. Actuators B Chem. 2016, 231, 759–767. [Google Scholar] [CrossRef]
  19. Essousi, H.; Barhoumi, H. Electroanalytical application of molecular imprinted polyaniline matrix for dapsone determination in real pharmaceutical samples. J. Electroanal. Chem. 2018, 818, 131–139. [Google Scholar] [CrossRef]
  20. Turco, A.; Corvoglia, S.; Pompa, P.P.; Malitesta, C. An innovative and simple all electrochemical approach to functionalize electrodes with a carbon nanotubes/polypyrrole molecularly imprinted nanocomposite and its application for sulfamethoxazole analysis. J. Colloid Interface Sci. 2021, 599, 676–685. [Google Scholar] [CrossRef]
  21. Özcan, L.; Sahin, Y. Determination of paracetamol based on electropolymerized-molecularly imprinted polypyrrole modified pencil graphite electrode. Sens. Actuators B Chem. 2007, 127, 362–369. [Google Scholar] [CrossRef]
  22. Jing, T.; Xia, H.; Guan, Q.; Lu, W.; Dai, G.; Niu, J.; Jae-Min, L.; Qiaoling, H.; Yong-Ill, L.; Yikai, Z.; et al. Rapid and selective determination of urinary lysozyme based on magnetic molecularly imprinted polymers extraction followed by chemiluminescence detection. Anal. Chim. Acta 2011, 692, 73–79. [Google Scholar] [CrossRef] [PubMed]
  23. Karthika, P.; Shanmuganathan, S.; Viswanathan, S.; Delerue-Matos, C. Molecularly imprinted polymer-based electrochemical sensor for the determination of endocrine disruptor bisphenol-A in bovine milk. Food Chem. 2021, 363, 130287. [Google Scholar] [CrossRef] [PubMed]
  24. Sadriu, I.; Bouden, S.; Nicolle, J.; Podvorica, F.I.; Bertagna, V.; Berho, C.; Amalric, L.; Vautrin-Ul, C. Molecularly imprinted polymer modified glassy carbon electrodes for the electrochemical analysis of isoproturon in water. Talanta 2020, 207, 120222. [Google Scholar] [CrossRef] [PubMed]
  25. Liu, B.; Ma, Y.; Zhou, F.; Wang, Q.; Liu, G. Voltammetric Determination of Sulfadiazine Based on Molecular Imprinted Electrochemical Sensor. Int. J. Electrochem. Sci. 2020, 15, 9590–9596. [Google Scholar] [CrossRef]
  26. Sadeghi, S.; Motaharian, A. Voltammetric sensor based on carbon paste electrode modified with molecular imprinted polymer for determination of sulfadiazine in milk and human serum. Mater. Sci. Eng. C 2013, 33, 4884–4891. [Google Scholar] [CrossRef]
  27. Mazzotta, E.; Turco, A.; Malitesta, C.; Cotvaglia, S. Molecularly Imprinted Polypyrrole for the Electrochemical Detection of Sulfadimethoxine: The Effect of Imprinting Parameters. In Proceedings of the IEEE Sensors 2014, Valencia, Spain, 2–5 November 2014; IEEE: New York, NY, USA, 2014; pp. 1835–1838. [Google Scholar] [CrossRef]
  28. Zhang, Z.; Li, Y.; Xu, J.; Wen, Y. Electropolymerized molecularly imprinted polypyrrole decorated with black phosphorene quantum dots onto poly(3,4-ethylenedioxythiophene) nanorods and its voltammetric sensing of vitamin C. J. Electroanal. Chem. 2018, 814, 153–160. [Google Scholar] [CrossRef]
  29. Afkhami, A.; Gomar, F.; Madrakian, T. Electrochemical Sensor for Dapsone Using Molecularly Imprinted Polypyrrole Membrane as a Recognition Element. J. Electrochem. Soc. 2015, 162, B109. [Google Scholar] [CrossRef]
  30. Radi, E.A.; El-Naggar, A.E.; Nassef, M.H. Determination of Coccidiostat Clopidol on an Electropolymerized-Molecularly Imprinted Polypyrrole Polymer Modified Screen Printed Carbon Electrode. Anal. Methods 2014, 6, 7967–7972. [Google Scholar] [CrossRef]
  31. Sooraj, M.P.; Nair, S.A.; Pillai, S.C.; Hinder, S.J.; Mathew, B. CuNPs decorated molecular imprinted polymer on MWCNT for the electrochemical detection of L-DOPA. Arab. J. Chem. 2020, 13, 2483–2495. [Google Scholar] [CrossRef]
  32. Zarejousheghani, M.; Rahimi, P.; Borsdorf, H.; Zimmermann, S.; Joseph, Y. Molecularly Imprinted Polymer-Based Sensors for Priority Pollutants. Sensors 2021, 21, 2406. [Google Scholar] [CrossRef] [PubMed]
  33. Sooraj, M.; Mathew, B. Fabrication of a Structure-Specific Molecular Imprinted Polymer–Based Electrochemical Sensor Based on CuNP-Decorated Vinyl-Functionalized Graphene for the Detection of Parathion Methyl in Vegetable and Fruit Samples. Food Anal. Methods 2019, 12, 1028–1039. [Google Scholar] [CrossRef]
  34. Tavares, A.P.; de Sá, M.H.; Sales, M.G.F. Innovative screen-printed electrodes on cork composite substrates applied to sulfadiazine electrochemical sensing. J. Electroanal. Chem. 2021, 880, 114922. [Google Scholar] [CrossRef]
  35. Sun, Y.; He, J.; Waterhouse, G.I.N.; Xu, L.; Zhang, H.; Qiao, X.; Xu, Z. A selective molecularly imprinted electrochemical sensor with GO@COF signal amplification for the simultaneous determination of sulfadiazine and acetaminophen. Sens. Actuators B Chem. 2019, 300, 126993. [Google Scholar] [CrossRef]
  36. Beigizadeh, H.; Ganjali, M.R.; Norouzi, P. FFTSW Voltammetric Sensor based on Eu2(CO3)3 Nanoparticles for Sulfadiazine Determination. Anal. Bioanal. Electrochem. 2022, 14, 846–859. [Google Scholar]
  37. Vilian, A.T.E.; Hwang, S.K.; Lee, M.J.; Park, B.; Huh, Y.S.; Han, Y.K. Gold nanoparticle decorated patronite on rGO for the quantification of sulfadiazine at nanomolar levels in contaminated water. Chem. Eng. J. 2022, 439, 135782. [Google Scholar] [CrossRef]
  38. Devi, R.K.; Ganesan, M.; Chen, T.W.; Chen, S.M.; Al-Onazi, W.A.; Al-Mohaimeed, A.M.; Elshikh, M.S.; Yu, Y.Y. 3D-nanocubes of N-doped carbon quantum dots adorned manganese oxide: A functional electrocatalyst for the sensitive detection of sulfadiazine. Colloids Surf. A Physicochem. Eng. Asp. 2022, 648, 129141. [Google Scholar] [CrossRef]
  39. Souza, C.D.; Braga, O.C.; Vieira, I.C.; Spinelli, A. Electroanalytical determination of sulfadiazine and sulfamethoxazole in pharmaceuticals using a boron-doped diamond electrode. Sens. Actuators B Chem. 2008, 135, 66–73. [Google Scholar] [CrossRef]
  40. Radha, A.; Wang, S.F. Bismuth sulfide microstructures decorated with functionalized boron nitride composite for electrochemical detection of sulfadiazine. Microchim. Acta 2022, 189, 429. [Google Scholar] [CrossRef]
  41. Lalmalsawmi, J.; Tiwari, D.; Lee, S.M.; Kim, D.J.; Kim, H. Efficient electrochemical sensor for trace detection of sulfamethazine in spring water: Use of novel nanocomposite material coated with Ag or Au nanoparticles. Microchem. J. 2022, 179, 107520. [Google Scholar] [CrossRef]
  42. Hong, X.; Ma, J. Sensitive Sulfadiazine Sensor Based on Multiwalled Carbon Nanotubes Wrapped with Polystyrene Sulfonate Polymer Chain. Int. J. Electrochem. Sci. 2017, 12, 6779–6787. [Google Scholar] [CrossRef]
  43. Hong, X.P.; Ma, J.Y. Electrochemical study of sulfadiazine on a novel phthalocyanine-containing chemically modified electrode. Chin. Chem. Lett. 2013, 24, 329–331. [Google Scholar] [CrossRef]
  44. Kokulnathan, T.; Kumar, E.A.; Wang, T.J.; Cheng, I.C. Strontium tungstate-modified disposable strip for electrochemical detection of sulfadiazine in environmental samples. Ecotoxicol. Environ. Saf. 2021, 208, 111516. [Google Scholar] [CrossRef]
  45. Yang, S.; Zheng, Y.; Zhang, X.; Ding, S.; Li, L.; Zha, W. Molecularly imprinted electrochemical sensor based on the synergic effect of nanoporous gold and copper nanoparticles for the determination of cysteine. J. Solid State Electrochem. 2016, 20, 2037–2044. [Google Scholar] [CrossRef]
  46. Maheshwaran, S.; Balaji, R.; Chen, S.M.; Tamilalagan, E.; Chandrasekar, N.; Ethiraj, S.; Samuel, M.S. Construction of graphene oxide wrapped gadolinium vanadate nanocomposites as an efficient electrocatalyst for the amperometric sensing of sulfadiazine. Process Saf. Environ. Prot. 2022, 168, 77–87. [Google Scholar] [CrossRef]
  47. Baby, J.N.; Sriram, B.; Wang, S.F.; George, M. Integration of samarium vanadate/carbon nanofiber through synergy: An electrochemical tool for sulfadiazine analysis. J. Hazard. Mater. 2021, 408, 124940. [Google Scholar] [CrossRef] [PubMed]
  48. Vinoth, S.; Govindasamy, M.; Wang, S.F.; Alothman, A.A.; Alshgari, R.A. Hydrothermally synthesized cubical zinc manganite nanostructure for electrocatalytic detection of sulfadiazine. Microchim. Acta 2021, 188, 131. [Google Scholar] [CrossRef]
  49. Kokulnathan, T.; Kumar, E.A.; Wang, T.J. Design and In Situ Synthesis of Titanium Carbide/Boron Nitride Nanocomposite: Investigation of Electrocatalytic Activity for the Sulfadiazine Sensor. ACS Sustain. Chem. Eng. 2020, 8, 12471–12481. [Google Scholar] [CrossRef]
  50. Ma, J.Y.; Hong, X.P. Simple Fabrication of Reduced Graphene Oxide—Ionic Liquid Composite Modified Electrode for Sensitive Detection of Sulfadiazine. Int. J. Electrochem. Sci. 2020, 15, 3729–3739. [Google Scholar] [CrossRef]
  51. Balasubramanian, P.; Settu, R.; Chen, S.M.; Chen, T.W. Voltammetric sensing of sulfamethoxazole using a glassy carbon electrode modified with a graphitic carbon nitride and zinc oxide nanocomposite. Microchim. Acta 2018, 185, 396. [Google Scholar] [CrossRef]
  52. Hong, X.P.; Zhu, Y.; Zhang, Y.Z. Electrocatalytic response of poly(cobalt tetraaminophthalocyanine)/multi-walled carbon nanotubes-Nafion modified electrode toward sulfadiazine in urine. J. Zhejiang Univ. Sci. B 2012, 13, 503–510. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Mani, V.; Devasenathipathy, R.; Chen, S.M.; Wang, S.F.; Devic, P.; Tai, Y. Electrodeposition of copper nanoparticles using pectin scaffold at graphene nanosheets for electrochemical sensing of glucose and hydrogen peroxide. Electrochim. Acta 2015, 176, 804–810. [Google Scholar] [CrossRef]
  54. Essousi, H.; Barhoumi, H.; Karastogianni, S.; Girousi, S.T. An Electrochemical Sensor Based on Reduced Graphene Oxide, Gold Nanoparticles and Molecular Imprinted Over-oxidized Polypyrrole for Amoxicillin Determination. Electroanalysis 2020, 34, 1546–1558. [Google Scholar] [CrossRef]
  55. Qin, L.; Liang, F.; Li, Y.; Wu, J.; Guan, S.; Wu, M.; Xie, S.; Manshi Luo, M.; Ma, D. A 2D Porous Zinc-Organic Framework Platform for Loading of 5-Fluorouracil. Inorganics 2022, 10, 202. [Google Scholar] [CrossRef]
  56. Campestrini, I.; de Braga, O.C.; Vieira, I.C.; Spinelli, A. Application of bismuth-film electrode for cathodic electroanalytical determination of sulfadiazine. Electrochim. Acta 2010, 55, 4970–4975. [Google Scholar] [CrossRef]
  57. De Braga, O.C.; Campestrini, I.; Vieira, I.C.; Spinelli, A. Sulfadiazine Determination in Pharmaceuticals by Electrochemical Reduction on a Glassy Carbon Electrode. J. Braz. Chem. Soc. 2010, 21, 813–820. [Google Scholar] [CrossRef] [Green Version]
  58. El Hassani, N.E.A.; Llobet, E.; Popescu, L.M.; Ghita, M.; Bouchikhi, B.; El Bari, N. Development of a highly sensitive and selective molecularly imprinted electrochemical sensor for sulfaguanidine detection in honey samples. J. Electroanal. Chem. 2018, 823, 647–655. [Google Scholar] [CrossRef]
  59. El-Feky, G.S.; El-Banna, S.T.; El-Bahy, G.S.; Abdelrazek, E.M.; Kamal, M. Alginate coated chitosan nanogel for the controlled topical delivery of Silver sulfadiazine. Carbohydr. Polym. 2017, 177, 194–202. [Google Scholar] [CrossRef] [PubMed]
  60. Zhang, Z.; Li, M.; Shen, F.; Ren, X. Direct fluorescence quantification of sulfadiazine from quenching of novel functional monomer based molecularly imprinted polymers. Anal. Methods 2015, 7, 5794–5800. [Google Scholar] [CrossRef]
  61. Oral, E.; Peppas, N.A. Dynamic studies of molecular imprinting polymerizations. Polymer 2004, 45, 6163–6173. [Google Scholar] [CrossRef]
Scheme 1. Electroanalytical strategy for SDZ detection on MIP-GCE/OPPy/CuNPs.
Scheme 1. Electroanalytical strategy for SDZ detection on MIP-GCE/OPPy/CuNPs.
Sensors 23 01270 sch001
Figure 1. Cyclic voltammograms of the electro-polymerization of pyrrole (0.1 M) in the absence (black) and presence (blue) of (10−3 M) SDZ on GCE in 0.1 M H2SO4 solution, with a scan rate of 100 mV/s and 15 scans.
Figure 1. Cyclic voltammograms of the electro-polymerization of pyrrole (0.1 M) in the absence (black) and presence (blue) of (10−3 M) SDZ on GCE in 0.1 M H2SO4 solution, with a scan rate of 100 mV/s and 15 scans.
Sensors 23 01270 g001
Figure 2. SEM images of the (A) GCE, (B) GCE/PPy, (C) GCE/PPy/SDZ, and (D) MIP-GCE/OPPy.
Figure 2. SEM images of the (A) GCE, (B) GCE/PPy, (C) GCE/PPy/SDZ, and (D) MIP-GCE/OPPy.
Sensors 23 01270 g002
Figure 3. Overoxidation of the imprinted PPy for the extraction of SDZ in 0.1 mol L−1 NaOH solution.
Figure 3. Overoxidation of the imprinted PPy for the extraction of SDZ in 0.1 mol L−1 NaOH solution.
Sensors 23 01270 g003
Scheme 2. Electrochemical oxidation of SDZ on the GCE.
Scheme 2. Electrochemical oxidation of SDZ on the GCE.
Sensors 23 01270 sch002
Figure 4. Cyclic voltammograms of the electrodeposition of CuNPs. Experimental conditions: 0.1 M Na2SO4, 10−3 M CuSO4.
Figure 4. Cyclic voltammograms of the electrodeposition of CuNPs. Experimental conditions: 0.1 M Na2SO4, 10−3 M CuSO4.
Sensors 23 01270 g004
Figure 5. (A) EDX spectra of the CuNPs/MIP-OPPy composite. (B) HR-TEM image of the CuNPs on the composite surface.
Figure 5. (A) EDX spectra of the CuNPs/MIP-OPPy composite. (B) HR-TEM image of the CuNPs on the composite surface.
Sensors 23 01270 g005
Figure 6. The FTIR spectra of the SDZ (a), PPy/SDZ (b), and MIP-OPPy (c) matrices.
Figure 6. The FTIR spectra of the SDZ (a), PPy/SDZ (b), and MIP-OPPy (c) matrices.
Sensors 23 01270 g006
Figure 7. Effect of pyrrole (A) and sulfadiazine (B) concentrations on the DPV peak current of sulfadiazine. Measuring conditions: 10−6 mol L−1 SDZ in 0.1 mol L−1 H2SO4 (pH = 2).
Figure 7. Effect of pyrrole (A) and sulfadiazine (B) concentrations on the DPV peak current of sulfadiazine. Measuring conditions: 10−6 mol L−1 SDZ in 0.1 mol L−1 H2SO4 (pH = 2).
Sensors 23 01270 g007
Figure 8. Effect of the number of electropolymerization cycles on the DPV peak current of sulfadiazine. Measuring conditions: 10−6 mol L−1 SDZ in 0.1 mol L−1 H2SO4 (pH = 2).
Figure 8. Effect of the number of electropolymerization cycles on the DPV peak current of sulfadiazine. Measuring conditions: 10−6 mol L−1 SDZ in 0.1 mol L−1 H2SO4 (pH = 2).
Sensors 23 01270 g008
Figure 9. Effect of the different incubation times on the DPV peak current of sulfadiazine. Measuring conditions: 10−6 mol L−1 SDZ in 0.1 mol L−1 H2SO4 (pH = 2).
Figure 9. Effect of the different incubation times on the DPV peak current of sulfadiazine. Measuring conditions: 10−6 mol L−1 SDZ in 0.1 mol L−1 H2SO4 (pH = 2).
Sensors 23 01270 g009
Figure 10. Effect of scan rate on the oxidation reaction of SDZ. Measuring conditions: 10−6 mol L−1 SDZ in 0.1 mol L−1 H2SO4 (pH = 2).
Figure 10. Effect of scan rate on the oxidation reaction of SDZ. Measuring conditions: 10−6 mol L−1 SDZ in 0.1 mol L−1 H2SO4 (pH = 2).
Sensors 23 01270 g010
Figure 11. (A) Variation of the DPV signal of SDZ for different values of pH. (B) Variation of the DPV signal of SDZ as a function of pH values.
Figure 11. (A) Variation of the DPV signal of SDZ for different values of pH. (B) Variation of the DPV signal of SDZ as a function of pH values.
Sensors 23 01270 g011aSensors 23 01270 g011b
Figure 12. (A) DPV response of the MIP sensor for concentrations of SDZ (10−9–10−5 mol L−1) in H2SO4 solution (pH 2), (B) the corresponding calibration curves for MIP and NIP sensors: DPV peak current versus the −logarithm of SDZ concentration.
Figure 12. (A) DPV response of the MIP sensor for concentrations of SDZ (10−9–10−5 mol L−1) in H2SO4 solution (pH 2), (B) the corresponding calibration curves for MIP and NIP sensors: DPV peak current versus the −logarithm of SDZ concentration.
Sensors 23 01270 g012
Figure 13. Responses obtained from the MIP-GCE/OPPy/CuNPs sensor in H2SO4 solution containing 10−6 mol L−1 of SDZ.
Figure 13. Responses obtained from the MIP-GCE/OPPy/CuNPs sensor in H2SO4 solution containing 10−6 mol L−1 of SDZ.
Sensors 23 01270 g013
Figure 14. DPV response for sulfadiazine and its interfering molecules by the MIP and NIP sensors.
Figure 14. DPV response for sulfadiazine and its interfering molecules by the MIP and NIP sensors.
Sensors 23 01270 g014
Table 1. FT-IR peak assignments for SDZ, PPy/SDZ, and MIP/OPPy.
Table 1. FT-IR peak assignments for SDZ, PPy/SDZ, and MIP/OPPy.
MatricesWave Number (cm−1)Peak Assignments
SDZ3351–3422Bending vibration of NH2
1327C=N
1581-CNS
666–785asymmetric SO2 stretching
PPy/SDZ672–785asymmetric SO2-N stretching
MIP/OPPy2910symmetric N-H stretching
2980asymmetric N-H stretching
1750symmetric C-H stretching
Table 2. Analytical performance of previously published electrochemical sensors for SDZ detection.
Table 2. Analytical performance of previously published electrochemical sensors for SDZ detection.
Recognition MaterialElectrochemical Technique
Electrode Type
Detection LimitDynamic RangeReference
MIP
PEDOT MIP on cork based SPEDPV
SPE
4.22 × 10−6 M8.0 × 10−6 M–186.0 × 10−6 M[34]
MWCNT-PA MIPDPV
GCE
0.68 × 10−6 M4 × 10−6 M–
50 × 10−6 M
[25]
PPy-MIP/GO@COFGCE
DPV
0.16 × 10−6 M0.5 × 10−6 M–200 × 10−6 M[35]
PA-MIP/CPEDPV
CPE
0.4 × 10−6 M2.0 × 10−7 M–1.0 × 10−4 M[26]
CuNPs/MIP-OPPyDPV
GCE
3.1 × 10−10 M10−9 M–10−5 MThis work
Nanomaterials
Eu2(CO3)3 NPsFFTSW voltammetry
CPE
4 × 10−8 M1.0 × 10−6 M–250.0 × 10−6 M[36]
AuNP/VS2-rGO compositesEIS
SPCE
4 × 10−10 M10 × 10−9 M–345 × 10−9 M[37]
N-CQD@Mn3O4 NCsDPV
SPCE
14 × 10−9 M0.5 × 10−6 M–663 × 10−6 M[38]
AgNP/TCBt, or AuNP/TCBtDPASV
GCE
0.097 × 10−6 M and 0.127 × 10−6 M0.89 × 10−6 M–107.7 × 10−6 M[39]
f-BN@Bi2S3 nanocompositeDPV
GCE
1.5 × 10−9 M0.01 × 10−6 M–62 × 10−6 M[40]
GdVO4-GO nanocompositeChronoamperometry
GCE
3.1 × 10−9 M0.02–145.5 × 10−6 M; 145.5–945.5 × 10−6 M[41]
SrWO4 nanoflakesDPV
SPCE
9 × 10−9 M0.05 × 10−6 M–235 × 10−6 M[42]
samarium vanadate/carbon nanofiber nanocompositeChronoamperometry
GCE
1.3 × 10−9 M0.009 × 10−6 M–445 × 10−6 M[43]
ZnMn2O4-NCDPV
GCE
2.1 × 10−9 M0.008 × 10−6 M–1264 × 10−6 M[44]
TiC/BN nanocompositesDPV
GCE
3 × 10−9 M0.01−44 × 10−6 M; 59−186 × 10−6 M[45]
rGO-(OPPF6) IL compositeChronoamperometry
GCE
0.07 × 10−6 M0.22 × 10−6 M–63.00 × 10−6 M[46]
g-C3N4/ZnO nanocompositeDPV
GCE
6.6 × 10−9 M20 × 10−9 M–1.1 × 10−6 M[47]
MWCNT–PSSChronoamperometry
GCE
0.6 × 10−6 M1.9 × 10−6 M–
160 × 10−6 M
[48]
CoPc-MWCNTChronoamperometry
GCE
____0.3 × 10−6 M–7.0 × 10−6 M[49]
CoIITAPc-MWCNTAmperometry
GCE
0.17 × 10−6 M0.5 × 10−6 M–43.5 × 10−6 M[50]
Bismuth filmDPV
GCE
2.1 × 10−6 M3.2–20.0 × 10−6 M; 20–97 × 10−6 M[51]
_____Electrochemical reduction
GCE
10.9 × 10−6 M62.7 × 10−6 M–
340 × 10−6 M
[52]
_____boron-doped diamond electrode
SWV
2.19 × 10−6 M8.01 × 10−6 M–1.19 × 10−4 M[54]
Table 3. Determination of sulfadiazine in spiked real milk samples (RSD n = 3).
Table 3. Determination of sulfadiazine in spiked real milk samples (RSD n = 3).
SamplesAdded (mol L−1)Found (mol L−1)Recovery (%)RSD (%)
54.70942.9
Milk108.32832.8
1514.0693.733.2
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

Elamin, M.B.; Ali, S.M.A.; Essousi, H.; Chrouda, A.; Alhaidari, L.M.; Jaffrezic-Renault, N.; Barhoumi, H. An Electrochemical Sensor for Sulfadiazine Determination Based on a Copper Nanoparticles/Molecularly Imprinted Overoxidized Polypyrrole Composite. Sensors 2023, 23, 1270. https://doi.org/10.3390/s23031270

AMA Style

Elamin MB, Ali SMA, Essousi H, Chrouda A, Alhaidari LM, Jaffrezic-Renault N, Barhoumi H. An Electrochemical Sensor for Sulfadiazine Determination Based on a Copper Nanoparticles/Molecularly Imprinted Overoxidized Polypyrrole Composite. Sensors. 2023; 23(3):1270. https://doi.org/10.3390/s23031270

Chicago/Turabian Style

Elamin, Manahil Babiker, Shazalia Mahmoud Ahmed Ali, Houda Essousi, Amani Chrouda, Laila M. Alhaidari, Nicole Jaffrezic-Renault, and Houcine Barhoumi. 2023. "An Electrochemical Sensor for Sulfadiazine Determination Based on a Copper Nanoparticles/Molecularly Imprinted Overoxidized Polypyrrole Composite" Sensors 23, no. 3: 1270. https://doi.org/10.3390/s23031270

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