Next Article in Journal
Technologies for Monoclonal Antibody Discovery and Development
Previous Article in Journal
Spirulina and Chlorella Dietary Supplements—Are They a Source Solely of Valuable Nutrients?
Previous Article in Special Issue
PtNiSnO2 Nanoframes as Advanced Electrode Modifiers for Ultrasensitive Detection of Trazodone in Complex Matrices
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nanostructured Ni-Zeolite Y and Carbon Nanohorns Electrode for Sensitive Electrochemical Determination of B-Group Vitamins

Department of Analytical Chemistry and Biochemistry, Faculty of Materials Science and Ceramics, AGH University of Krakow, al. A. Mickiewicza 30, 30-059 Kraków, Poland
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2025, 26(21), 10469; https://doi.org/10.3390/ijms262110469
Submission received: 5 October 2025 / Revised: 21 October 2025 / Accepted: 25 October 2025 / Published: 28 October 2025
(This article belongs to the Special Issue Electrochemistry: Molecular Advances and Challenges)

Abstract

This work presents the fabrication and analytical application of nanostructured Ni-zeolite (NiZY) and carbon nanohorns (CNHs) modified glassy carbon electrode (NiZY/CNHs-GCE) in the differential pulse voltammetric (DPV) determination of vitamin B2 (VB2) molecules. The synergistic combination of NiZY and CNHs significantly enhances the electrochemical performance of the sensor, as confirmed by structural, textural, morphological, and electrochemical studies. The redox behavior of VB2 on NiZY/CNHs-GCE was found to be adsorption-controlled, involving a two-electron, two-proton reversible reduction process. Under optimized conditions, the DPV response of NiZY/CNHs-GCEs in McIlvaine buffer solution (pH 3.4) exhibited a linearity in the VB2 concentration range of 0.01 to 0.20 mg L−1 (r = 0.9993) with a detection limit of 3.2 µg L−1 (8.6 × 10−9 mol L−1). Furthermore, well-resolved reduction peaks of vitamins B2 and B9 (VB9) enabled their simultaneous and selective detection, with linear ranges of 0.01 to 0.20 mg L−1 for VB2 and 0.01 to 0.16 mg L−1 for VB9. The proposed analytical method, characterized by high selectivity and robustness, was successfully applied in the determination of both vitamins in commercially available dietary supplements, achieving relative errors within −6.2% to 2.7%.

1. Introduction

The vitamin B2 molecule (VB2), commonly known as riboflavin, is an essential water-soluble micronutrient that plays a crucial role in cellular metabolism, acting as a precursor of the coenzymes flavin mononucleotide (FMN) and flavin adenine dinucleotide (FAD). These cofactors participate in various biochemical processes, including cell respiration, antioxidant defense, and the metabolism of fats, proteins, and carbohydrates [1,2]. Since the human organism cannot synthesize riboflavin, it must be obtained from dietary sources such as dairy products, meat, eggs, and dark-green leafy vegetables. Inappropriate riboflavin intake can lead to a variety of metabolic disorders and clinical manifestations collectively known as ariboflavinosis [3]. Given its essential role in maintaining human health, the accurate quantification of VB2 in biological, pharmaceutical, and food matrices is crucial for nutritional evaluation, quality control, and clinical diagnostics.
The chemical properties of the VB2 molecule, such as its water solubility, intrinsic fluorescence, and redox activity, present both opportunities and challenges for analytical determination. Interference from complex matrices, low analyte concentrations, and the need for rapid and cost-effective analysis have driven the development of diverse analytical methods. Classical approaches include fluorometry [4,5,6] and spectrophotometry [7,8,9], which exploit the compound’s intrinsic fluorescence and absorbance properties, respectively. High-performance liquid chromatography (HPLC), often coupled with fluorescence or mass spectrometric detection, remains standard in terms of sensitivity, selectivity, and applicability to complex matrices [10,11,12]. Capillary electrophoresis [13] and enzyme-linked assays [14] have also been employed. While these methods offer high analytical performance, they often require labor-intensive sample preparation, expensive instrumentation, or long analysis times. Electrochemical methods, by contrast, provide attractive alternatives owing to their low cost, high sensitivity, portability, and compatibility with miniaturization for on-site analysis. Among these, voltammetric techniques, such as cyclic voltammetry (CV), differential pulse voltammetry (DPV), and square-wave voltammetry (SWV), have received growing attention for VB2 determination, primarily due to its advantageous electrochemical properties imparted by the isoalloxazine ring system [15]. The redox activity of this molecule provides well-defined and reproducible electrochemical signals, thereby enabling sensitive and selective detection across a wide range of electrode platforms, including modified carbon paste electrodes (CPEs) [16,17,18,19,20,21], glassy carbon electrodes (GCEs) [22,23,24,25,26,27], pencil graphite electrodes (PGEs) [28,29,30,31], and screen-printed electrodes (SPEs) [32,33,34,35]. Advances in electrode materials, including carbon nanostructures [36,37], metal nanoparticles [38,39,40], and conducting polymers [41,42,43], have further enhanced the selectivity and sensitivity of voltammetric assays of VB2 molecules. These developments have allowed the successful application of voltammetric methods of its determination in diverse matrices, including pharmaceutical formulations, fortified foods, and biological samples, often with minimal pretreatment requirements.
In the area of electrode-modifying nanomaterials, a particularly distinctive position is occupied by zeolites, crystalline aluminosilicates with well-defined pore structures, high surface area, and exceptional ion-exchange capabilities. These materials have attracted significant attention due to their unique ability to selectively host target analytes [44,45] and catalyze redox reactions [46]. Unlike carbon-based nanomaterials, which primarily improve conductivity and provide a large surface, or metal nanoparticles, which offer excellent electrocatalytic activity but may suffer from aggregation or instability, zeolites stand out for their intrinsic molecular recognition, chemical stability, and tunable porosity [47,48]. Nevertheless, the inherently low electrical conductivity of zeolites necessitates their integration with conductive components, most commonly carbon-based materials, to achieve efficient electron transfer. Although graphite [49,50,51] has long been the predominant electron-conducting additive in zeolite-modified electrodes (ZMEs), recent studies increasingly explore carbon black [52], mesoporous carbon [53,54], and graphene oxide [55,56]. The combination of zeolites with advanced carbon-based nanomaterials results in hybrid electrodes with enhanced sensitivity, selectivity, and stability for a wide range of electroanalytical applications.
In this work, a voltammetric sensor was developed by modifying a glassy carbon electrode (GCE) with a nanostructured composite of Ni-exchanged synthetic zeolite Y (NiZY) and carbon nanohorns (CNHs) for the determination of water-soluble B-group vitamins. The synergistic combination of electrocatalytic NiZY and conductive CNHs creates a novel zeolite-modified composite electrode that offers superior electrochemical performance, confirmed by comprehensive morphological, structural, textural, and electrochemical measurements. Optimal electrochemical conditions for VB2 determination using NiZY/CNHs-GCEs were established, with selectivity confirmed against potential interferents. Based on CV and DPV studies, the mechanism of redox reaction of VB2 molecules on the surface of NiZY/CNHs-GCEs was also proposed. Furthermore, the possibility of VB2 analysis in the presence of vitamin B9 (VB9) was demonstrated and supported by their simultaneous DPV determination in dietary supplements.

2. Results and Discussion

2.1. Characterization of Electrode Modifiers

Structurally, zeolite Y belongs to the faujasite (FAU) framework type, for which [SiO4] and [AlO4] tetrahedra are arranged into secondary building units (SBUs) in the form of double six-membered rings (D6R). These units link sodalite cages and hexagonal prisms to form large supercages (~12 Å), which are interconnected through a 12-membered ring that defines pore apertures of ~7.4 Å [57]. This structural arrangement results in the presence of characteristic ring vibration bands, which can be clearly distinguished in the FT-IR spectrum of zeolite Y (Figure 1A). An important feature of the FAU framework is the band located at 578 cm−1, resulting from D6R vibrations connecting the sodalite cages and defining the overall topology of zeolite Y. The most intense bands appear at about 1140 and 1018 cm−1, assigned to asymmetric stretching of Si–O–Si and Si–O–Al bonds, respectively. Weaker bonds at 750–800 cm−1 correspond to symmetric stretching vibrations of the tetrahedral units, while the bands in the region of 450–500 cm−1 arise from the bending modes of T–O–T (T = Si, Al). Hydroxyl groups and adsorbed water contribute additional bands at 3000–4000 cm−1 and ca. 1640 cm−1 [58]. Comparison of the FT-IR spectra of zeolite Y before and after Ni(II) ion exchange (Figure 1A) shows that the positions of the aforementioned bands remain essentially unchanged, with only minor variations in their intensity. The observed changes, particularly in the T–O–T and pseudo-lattice regions, reflect the subtle degree of ring deformation caused by incorporation the Ni2+ ions into the zeolite Y structure [59].
Since IR spectroscopy is largely insensitive to nonpolar bonds, Raman spectroscopy was employed to characterize the sp2 carbon lattice vibrations of CNHs. The Raman spectrum of the CNHs (Figure 1B) exhibited characteristic D (disordered) and G (graphitic) bands with nearly equal scattering strengths located at approximately 1350 and 1580 cm−1, respectively. In CNHs, the D band, originating from defects, edges, or lattice distortions, mainly reflects the conical tips and curvature of the nanohorns, while the G band corresponds to the in-plane vibrations of sp2 carbon atoms in the graphitic lattice. Additionally, a broad 2D (G′) band is observed around 2700 cm−1 reflecting the multilayered and curved nature of the CNH walls [60].
Nitrogen adsorption–desorption analysis (Table 1) indicates that zeolite Y possesses over 2-fold-higher BET surface area and a predominantly microporous structure with an average pore diameter of 1.95 nm, whereas CNHs exhibit limited microporosity, with a relatively large external surface area (284.4 m2·g−1). These results emphasize that while zeolite Y offers a highly microporous framework, carbon nanohorns provide complementary external surface sites, making their combination particularly promising for the development of sensors with enhanced sensitivity and selectivity.
Morphological characterization of the electrode modifiers shows that zeolite Y (Figure 1C) exhibits a well-defined crystalline structure, with cubic particles characteristic of the FAU framework. The crystals are relatively uniform in size, ranging from 0.5 to 2 µm, and display smooth surfaces with sharp edges, indicative of high crystallinity and controlled synthesis conditions. Concomitantly, the EDS spectrum of zeolite Y after ion exchange with nickel cations (Figure 1D) shows distinct Ni peaks, providing clear evidence of the successful substitution of exchangeable cations within the zeolite structure. Quantitative analysis of nickel in the zeolite, performed on multiple distinct surface points, revealed average Ni atomic concentrations of 0.56 at.%.
The SEM observation of CNHs (Figure 1E) reveals their characteristic tendency to form large aggregates rather than exist as isolated nanostructures. These aggregates appear as spherical or flower-like assemblies composed of numerous nanostructures. As a result of the nanoscale dimensions of CNHs and the dense packing within aggregates, the outlines of individual nanohorns are difficult to distinguish. Finally, the SEM image of the NiZY/CNH composite (Figure 1F) shows the morphological characteristics of both materials. The modifying layer exhibits a heterogeneous microstructure in which the cubic crystals of zeolite Y are uniformly dispersed within a network of CNH aggregates. The CNHs form an interconnected network surrounding the zeolite crystals, resulting in a continuous and stable layer.

2.2. Electrochemical Characterization of Modified Electrodes

The impact of surface modifications on the performance of the developed sensors was examined through CV (Figure 2A) and EIS (Figure 2B) measurements conducted in a 0.1 mol L−1 solution containing 1 mmol L−1 of the [Fe(CN)6]4− model redox probe. To interpret the EIS data, the Nyquist plots were fitted with an equivalent electrical circuit corresponding to the Randles cell (Figure 2B), in which the double-layer capacitance was modeled by a Constant Phase Element (CPE) to reflect the porous and heterogeneous nature of the modified electrodes. On this basis, key parameters such as charge transfer resistance (Rct), Warburg impedance coefficient (σ), electroactive surface area (Ael), double-layer capacitance (Cdl), and the heterogeneous electron-transfer rate constant (ks) were established and summarized in Table 2. The Rct and σ value for the impedance spectra recorded at the formal potential (Ef) of the redox probe were determined according to Equations (1) and (2):
R c t = R T n 2 F 2 A k s · 2 ( D O D R ) 1 4 D O 1 2 · C O + D R 1 2 · C R
σ = R T n 2 F 2 A 2 · 4 D O 1 2 · C O + D R 1 2 · C R
where R indicates the universal gas constant (R = 8.314 J mol−1 K−1), T is the temperature [K], F is the Faraday constant (F = 96,485.3 C mol−1), n is the number of transferred electrons (n = 1), A is the electroactive surface area of the electrode, DO and DR are the diffusion coefficients of the oxidized and reduced forms (7.2 × 10−6 cm2 s−1), respectively, and CO and CR are their concentrations in the solution.
The heterogeneous rate constant (ks) can be determined from the impedance data by taking the ratio of Equations (1) and (2), which gives the expression presented in Equation (3):
k s = σ R c t · D O / R 2
where DO/R indicates the average diffusion coefficient of the oxidized and reduced forms.
As can be seen in Figure 2A, a bare GCE in the presence of a model redox probe shows a reversible electrochemical response characterized by well-defined anodic and cathodic peaks with a peak-to-peak separation (∆E) equal to 72 mV (Table 2), indicating efficient electron transfer. At the same time, EIS measurements reveal a high-frequency semicircle corresponding to charge transfer resistance and a low-frequency linear Warburg segment due to diffusion of the redox species. The observed double-layer capacitance value of 24.7 µF cm−2 reflects a clean electrode surface that is largely unaffected by adsorption or ion-exchange phenomena.
The electrochemical response of the GCE exhibits notable changes upon modification with the layer containing only zeolite or carbon nanohorns. For the NiZY-GCE, the substantially higher ∆E (261 mV) compared to the theoretical value indicates a quasi-reversible [Fe(CN)6]3−/4− redox process, arising from the insulating and porous nature of the zeolite framework, which restricts charge transfer between the redox probe and the electrode. The pronounced decrease in both the electroactive surface area (4.2 mm2) and the peak current (1.0 µA) further indicates the limited number of active sites for the redox probe. These observations are consistent with the high charge-transfer resistance (Rct = 70.7 kΩ) and the low heterogeneous rate constant (ks = 1.6 × 10−6 cm s−1), both of which confirm that electron transfer is strongly hindered by the zeolite film. Additionally, the elevated Warburg coefficient indicates diffusion limitations within the zeolite micropores. Together, these factors confirm that the electrochemical response of the NiZY-GCE is significantly influenced by the porous structural and resistive properties of zeolites. In the case of the CNHs-modified GCE, a notable asymmetry between the anodic and cathodic peak currents, along with the large peak separation (∆E = 360 mV) and the high Ipa/Ipc ratio (1.2), indicates a quasi-reversible electron-transfer behavior. The excellent electrical conductivity and the high surface area of CNHs lead to a substantial increase in the anodic peak current (Ipa = 21.0 µA) and the electroactive surface area (Ael = 9.7 mm2), thereby improving the accessibility of the redox probe to active sites. In contrast to an insulating zeolite-containing film, the CNH layer facilitates electron transfer, as evidenced by the relatively low Rct and higher ks values, reflecting faster electron-transfer kinetics. Moreover, the conical morphology of CNHs creates interconnected pathways that enhance diffusion of the redox probe, thereby reducing mass-transfer limitations, as indicated by the smaller Warburg coefficient compared to the NiZY-GCE.
Ultimately, the NiZY/CNHs-GCE exhibits enhanced electrochemical performance due to the complementary properties of zeolite and carbon nanohorns. The zeolite provides a well-structured microporous framework that can pre-concentrate the redox probe, facilitating its interaction with the electrode surface. At the same time, the CNHs offer high conductivity and form an interconnected pathway network that enables efficient electron transfer throughout the electrode. This synergy results in an increased electroactive surface area (Ael = 10.0 mm2) and reduced mass-transfer limitations (σ = 2.8 kΩ·s−1/2). By combining these materials, the composite electrode achieves a favorable balance between high surface area, number of active sites, and efficient electron-transfer pathways, illustrating the advantage of integrating zeolites and carbon nanohorns for the improved electrochemical performance of NiZY-CNHs.
The favorable electrochemical properties of the NiZY/CNHs-GCE can be further demonstrated by its improved analytical performance in VB2 determination, as confirmed by the enhanced current response compared to the bare GCE and electrodes modified with each component individually (Figure 2C). For the bare GCE, which served as the substrate for the deposition of the modifying layer, the DPV signal corresponding to the reduction of 50 µg L−1 VB2 was highly unrepeatable due to pronounced adsorption of the vitamin on the unmodified electrode surface. This poor reproducibility makes the GCE unsuitable for reliable analysis, emphasizing the need to modify its surface to achieve accurate electrochemical detection of VB2.
The deposition of the layer containing only zeolite nanoparticles led to very low electrochemical activity of NiZY-GCE toward VB2 molecules. This is especially apparent after background subtraction, which reveals only a minimal current response of NiZY-modified GCE (Figure 2D), likely due to the insulating nature and microporous structure of the zeolite Y. In contrast, modification of the GCE with a CNHs-containing layer allowed the recording of a stable and reproducible reduction signal for VB2, distinguished by its reversible character and well-defined shape. However, the corresponding peak current is nearly four times lower than that obtained for the NiZY/CNHs-GCE, underscoring the superior performance of the composite electrode. Furthermore, the observed shift of the peak potential toward a less negative value, from −0.332 V and −0.339 V for the CNHs-GCE and NiZY-GCE, respectively, to −0.249 V for the NiZY/CNHs suggests a catalytic effect, indicating that the composite facilitates the analyte redox process and improves reaction kinetics by lowering overpotential for VB2 reduction. Moreover, the DPV results (Figure 2C) show a clear increase in the capacitive background current in the following order: bare GCE < CNHs-GCE < NiZY-GCE < NiZY/CNHs-GCE. The higher background signal for NiZY-GCE compared to CNHs-GCE indicates a strong adsorption of analyte molecules by the zeolite layer, even though zeolite is electrically insulating. The highest response for the NiZY/CNHs-GCE results from the synergistic effect of the conductive CNHs and the high adsorption capacity of NiZY. However, the Cdl values from EIS (Table 2) decrease in the opposite order, which is caused by the strongly heterogeneous and porous nature of the modified layers. This interpretation is supported by the DPV results, which confirm that the actual electroactive area increases despite the lower apparent Cdl from EIS. These differences highlight the advantageous role of both composite components: while CNHs contribute high conductivity and efficient charge transport, the addition of Ni-exchanged zeolite introduces a microporous framework capable of pre-concentrating VB2 molecules near the electrode surface, enhancing both the number of active sites and the kinetics of electron transfer, thereby delivering superior electrochemical performance for the composite electrode.

2.3. Investigation of the Redox Behavior of VB2 on the NiZY/CNHs-GCE

The electrochemical behavior of VB2 molecules on the NiZY/CNHs-GCE was systematically investigated using CV and DPV measurements. VB2 at a concentration of 1 mg L−1 was subjected to CV analysis at scan rates ranging from 0.006 to 0.1 V s−1 (Figure 3A). The recorded voltammograms displayed well-defined anodic and cathodic peaks, characteristic of a reversible redox couple corresponding to the electron-transfer reaction of VB2. The near-unity ratio of anodic to cathodic peak currents (Ipa/Ipc = 1.01) and the unchanged peak positions over the range of scan rate indicate that the electron-transfer process is fully reversible and kinetically facile under the experimental conditions. The dependence of peak currents on scan rate provided further mechanistic insight. The obtained linear relationship between the peak currents and scan rate (inset in Figure 3A), rather than the square root of scan rate, indicated that the VB2 redox reaction is adsorption-controlled. For accurate determination of the number of electrons (n) involved in the electrochemical redox reaction of VB2 molecules in adsorption-controlled electrochemical systems, the difference between the cathodic peak potential (Eca) and the potential corresponding to half of the peak current (Eca/2) was carefully measured and found to be 0.032 V, according to Equation (4) [61]:
E p c E p c / 2 = 0.0565 n             [ V ]
The value corresponding to an electron-transfer number was calculated to be 1.7, indicating that the electrochemical redox reaction of VB2 occurs through a two-electron process.
The redox behavior of VB2 molecules was further analyzed in supporting electrolyte of varying pH values. For this purpose, DPV measurements were performed in a series of McIlvaine buffer solutions with pH values ranging from 2.6 to 3.6, and the corresponding signals of VB2 at a concentration of 0.2 mg L−1 were recorded for each solution (Figure 3B). The voltammograms revealed that both anodic and cathodic peak potentials shifted negatively with increasing pH, indicating the participation of protons in the electrochemical redox process. A linear relationship was observed between the VB2 cathodic peak potential and pH (Figure 3C) with a slope of (−0.060 ± 0.001) V per unit change in pH, close to theoretical value of 0.059 V pH−1, suggesting that the number of protons involved in the reaction is equal to the number of electrons transferred.
The results of this study demonstrate that the electrochemical redox reaction of VB2 molecules at the NiZY/CNHs-modified GCE proceeds through adsorption-controlled kinetics involving the exchange of two electrons and two protons, as illustrated in Figure 3D. The reversible reduction occurs at the isoalloxazine ring, where electron transfer converts VB2 between its oxidized quinone-like state and the reduced hydroquinone-like form. This behavior is consistent with previously reported mechanisms for flavin systems, confirming the proton-coupled, two-electron nature of the VB2 redox process [20,27,31,34].

2.4. Optimization of the Experimental Conditions

As shown in Figure 3C, the peak current of VB2 molecules depends on the pH of the McIlvaine buffer solution, with the highest response observed at pH 3.4 (green line), which was selected as the optimal condition for subsequent measurements. Concurrently, to achieve the most sensitive and accurate detection of VB2, the key parameters of the DPV technique were systematically optimized. These included step potential (Es), which was changed from 1 to 6 mV; pulse amplitude (dE), which was varied from 10 to 60 mV (both positive and negative modes); and current sampling time (ts) and waiting time (tw), which were modified from 5 to 40 ms. Each parameter was individually varied to assess its effect on the VB2 reduction signal in terms of peak height, shape, symmetry, and potential. In consequence, the optimal conditions for DPV determination of VB2 were chosen as follows: Es = 3 mV, dE = 40 mV, ts = 10 ms, and tw = 10 ms.

2.5. Analytical Performance

The sensitivity of VB2 determination on the NiZY/CNHs-GCE in McIlvaine buffer solution (pH 3.4) was evaluated under optimal conditions by DPV measurements. Each voltammogram, recorded at different concentrations of VB2 molecules, was background-corrected by subtracting the signal obtained in the supporting electrolyte without the depolarizer (dashed line in Figure 4), resulting in the color-marked curves. Linearity was assessed by considering both the cathodic and anodic peak currents, and the corresponding calibration plots of peak current versus VB2 concentration were constructed and shown in Figure 4A,B, respectively. In addition, the analytical performance of the VB2 sensor was also tested in the presence of vitamin B9 (VB9) to demonstrate the possibility of their simultaneous determination. DPV measurements were conducted under the same parameters as in the individual experiment, with the exception of the supporting electrolyte, which was McIlvaine buffer at pH 3.0. The cathodic responses of VB2 and VB9 molecules in the mixture were investigated when the concentrations of both vitamins were simultaneously varied in the range from 0.01 to 0.2 mg L−1. The resulting DVP curves, along with the corresponding calibration graphs, are presented in Figure 4C–E. The calculated calibration curve coefficients and analytical characteristics for the individual and simultaneous determination of VB2 and VB9 molecules are summarized in Table 3.
For individual determination of VB2, both cathodic (Figure 4A) and anodic (Figure 4B) peaks were observed at similar potentials of approximately −0.25 V. The well-developed and symmetric nature of the obtained signals confirms the reversible redox behavior of riboflavin on the NiZY/CNHs-GCE surface. In both directions, an excellent linear correlation was achieved between the peak current and the VB2 concentration in the range of 0.01 to 0.2 mg L−1, highlighting the reliable quantitative performance of the sensor. The limit of detection (LOD) and the limit of quantification (LOQ) were estimated using the equations LOD = 3.3·SD/b and LOQ = 10·SD/b, where SD is the standard deviation of the y-intercept of the regression line and b is the slope of the calibration curve. The results obtained for both anodic and cathodic responses were comparable (Table 3), with LOD values in the nanomolar range, demonstrating the suitability of the NiZY/CNHs-GCE for trace-level determination of VB2 molecules.
In the simultaneous determination of VB2 and VB9, a significant separation between their reduction peak potentials was observed (Figure 4C), allowing for distinct and interference-free electrochemical signals for each analyte. In the mixture solution containing both vitamins, the sensor exhibited a wide and linear range, extending from 0.01 to 0.20 mg L−1 for VB2 and from 0.01 to 0.16 mg L−1 for VB9. Although the presence of VB9 caused a slight reduction in the sensitivity to VB2 determination, this effect did not significantly compromise the overall analytical performance of the electrode. In fact, despite this signal attenuation, the proposed method achieved an LOD in the nanomolar range for both vitamins, underlining the high sensitivity of the NiZY/CNHs-GCE. It is worth emphasizing that the ability to simultaneously quantify VB2 and VB9 represents a significant analytical advantage of the proposed sensor, especially for applications involving complex biological or pharmaceutical matrices, where both vitamins may coexist at trace levels.
As shown in Table 4, various voltammetric sensors have been developed for the electrochemical determination of VB2 molecules, employing different electrode modification strategies to enhance analytical performance in terms of sensitivity, selectivity, and detection limits. Although they provide an overall analytical advantage, these modification strategies are limited by inherent material challenges. For example, carbon-based nanomaterials often require chemical activation or surface functionalization and tend to agglomeration, which can reduce reproducibility and long-term stability. Metal nanoparticles may suffer from instability, high reactivity, potential toxicity, safety concerns, and complex synthesis procedures involving hazardous reagents and high energy input. Conducting polymer can exhibit poor mechanical stability, dopant-dependent signal drift, non-uniform coatings, and limited stability in extreme conditions [62].
In contrast, NiZY/CNHs-GCE provides a simple, robust, and cost-effective alternative for the electrochemical sensing of VB2 molecules. The synergistic combination of zeolite Y and carbon nanohorns endows a uniquely structured electrode surface with a high surface area and abundant active sites that facilitate efficient electron transfer. Zeolite Y contributes excellent ion-exchange capacity and chemical stability, providing a well-defined microenvironment for the dispersion of Ni2+-based catalytic sites and enhanced selectivity toward VB2 molecules. Meanwhile, CNHs offer outstanding conductivity and mechanical strength without requiring any chemical pretreatment, which constitutes a distinct advantage over graphene or carbon nanotubes. This NiZY/CNHs-GCE not only achieves comparable or superior analytical performance to sensors based on noble metals or multifunctional polymers but does so with significantly lower cost, simpler preparation, and higher environmental safety. The fabrication procedure is straightforward, requiring only basic laboratory equipment and avoiding any hazardous reagents or sophisticated apparatus. The resulting suspension remains stable for days, enabling the preparation of multiple electrodes with consistent performance. Furthermore, sensors fabricated by a simple drop-casting method exhibit exceptional reproducibility, long-term operational stability, and mechanical durability, maintaining their electrochemical response over hundreds of measurements without special storage conditions. The last statement is further supported by the results of stability and reproducibility studies, which showed that ten consecutive voltammetric scans at VB2 concentrations of 50 µg L−1 and 100 µg L−1 recorded under identical conditions resulted in relative standard deviation (RSD) values of 2.4% and 1.3%, respectively. These results confirmed excellent short-term stability and repeatability with a consistent signal response during successive measurements. On the other hand, after 7, 14, and 21 days of sensor storage under laboratory conditions, the peak current for both VB2 concentrations remained above 87% of the initial value, indicating minimal signal drift and confirming robust long-term stability of NiZY/CNHs-GCE. Reproducibility was further verified by comparing the responses of five independently prepared electrodes under the same conditions, resulting in an RSD of 8.6%, demonstrating reliable electrode fabrication and uniform analytical performance.
In summary, the NiZY/CNHs-GCE combines the structural robustness of zeolites, the electrochemical efficiency of carbon nanostructures, and the cost-effectiveness of scalable fabrication. This unique combination positions it as a highly competitive alternative to more complex nanocomposite systems, offering an ideal balance between analytical performance, environmental sustainability, and practical usability for routine electrochemical applications.

2.6. Selectivity

The selectivity of NiZY/CNH-GCE was also evaluated by testing its response to VB2 in the presence of potentially interfering species representing other vitamins, common biomolecules, pharmaceutical excipients, and surface-active compounds. For this purpose, the solution containing VB2 at a concentration of 50 µg L−1 was combined with a 4- to 200-fold excess of vitamin B6 (VB6) and vitamin C (VC), a 20- to 1000-times excess of glucose and caffeine, and 10- to 200-times excess of surface-active compounds, including sodium dodecyl sulphate (SDS), cetrimonium bromide (CTAB), and Triton-X100. The effect of common excipients of dietary supplements (titanium dioxide, starch, magnesium stearate, and cellulose) was tested individually by adding an appropriate weighed amount of 2 to 10 mg of each substance directly to the voltammetric cell. DPV measurements conducted in both the absence and in the presence of each individual species were used to evaluate changes in VB2 peak current, potential, and shape, thus assessing possible interference effects. Tolerance limits were defined as the highest interferent concentration that causes no more than a ±10% change in peak current.
Of the vitamins evaluated, VB6 has a pronounced effect on the VB2 reduction signal, increasing the peak current by 50% at 200-fold excess, while the addition of vitamin C did not cause significant change in the peak current, shape, and position. The observed enhancement of the VB2 peak current in the presence of the highest excess of VB6 may be attributed to effects of adsorption or surface activation. Specifically, VB6 molecules may undergo a partial co-adsorption, which locally enhances the concentration of VB2 near the electrode surface and/or modify its physicochemical properties, such as surface roughness or hydrophilicity, thus improving the accessibility and electron transfer efficiency of VB2. In the case of common biomolecules, only caffeine exhibits an interference effect causing enhancement in the voltammetric response of NiZY/CNHs-GCE to 120% of the initial value of VB2 peak current at 1000-fold excess. A more significant impact is related to surface-active compounds, i.e., Triton X-100 (nonionic surfactant) and CTAB (cationic surfactant), the presence of which in the 200-fold excess is associated with a 22% and 50% decrease in VB2 reduction signal, respectively. Among dietary supplement excipients, a slight signal increase (less than 17%) is observed for the highest tested amount of magnesium stearate added to the measurement cell. On the other hand, cellulose demonstrated a suppressive effect on the VB2 signal, reducing it to approximately 87% of the initial value, which can be connected with the solid phase present in the supporting electrolyte that limits the electrode reaction. Nevertheless, minimizing the interference effect from tablet excipients can be achieved by implementation of appropriate sample preparation strategies (e.g., solid particulate filtration).
The results of the selectivity assessment demonstrate that although some compounds can influence the VB2 response at very high excess, the proposed NiZY/CNHs-GCE provides reliable performance and selectivity under typical sample conditions.

2.7. Analytical Application

The NiZY/CNHs-GCE sensor was further evaluated for its applicability in the analysis of commercial vitamin dietary supplements (Figure 5) containing VB2 and VB9. Sample preparation was carried out according to the procedure described in Section 3.3.2 and analyzed using the standard addition method, which allows for accurate quantification in complex matrices and minimizes matrix-related interferences. The results of the measurements are summarized in Table 5.
For both vitamins, a well-defined peak corresponding to their reduction was observed on the surface of the NiZY/CNHs-GCE. Furthermore, peak currents increased linearly with each addition of the standard, demonstrating the quantitative response of the sensor for both VB2 and VB9 molecules in complex sample matrices. The measured vitamin content (Table 5) corresponds closely to the manufacturer’s declared values, with a relative error (RE) ranging from −6.2% to 2.7%, clearly indicating that the sensor provides accurate determinations of both analytes. At the same time, the excellent precision with relative standard deviation (RSD) below 2.1% confirms reliable and precise quantification in complex supplement matrices. Importantly, all measured values fall within acceptable tolerance limits for dietary supplements, further validating the suitability of the method for routine quality control of dietary supplements.

3. Materials and Methods

3.1. Chemicals and Solutions

Zeolite Y (ZY) in the sodium form was acquired from Thermo Scientific (Waltham, MA, USA) (Alfa Aesar). Carbon nanohorns, as grown (CNHs), and polystyrene (average Mw~35,000) were purchased from Sigma-Aldrich (St. Louis, MO, USA). Organic solvent, i.e., dichloromethane (ACS Reagent), and an acetone solution suitable for HPLC (≥99.9%) were obtained from Honeywell Research Chemicals (Seelze, Germany) and Avantor Performance Materials Poland S.A. (Gliwice, Poland), respectively. Nickel(II) nitrate hexahydrate, 98%, vitamin B2 (riboflavin, Secondary Pharmaceutical Standard), and vitamin B9 (folic acid, meeting USP testing specifications) were acquired from Sigma-Aldrich. The standard stock solution of vitamin B2 (VB2), as well as vitamin B9 (VB9) at a concentration of 1000 mg L−1 with the addition of 0.02 mol L−1 NaOH, was prepared weekly and stored at 4 °C under light-protected conditions. Working solutions of lower concentrations were prepared daily by diluting the stock solutions of VB2 and VB9 with double-distilled water. The supporting electrolyte in the form of citrate-phosphate buffer (McIlvaine buffer) was prepared by mixing appropriate volumes of 0.1 mol L−1 citric acid (Avantor Performance Materials Poland S.A.) with 0.2 mol L−1 Na2HPO4 (Avantor Performance Materials Poland S.A.) in order to obtain a solution with a pH value in the range of 2.6 to 3.6. Tested interference compounds, i.e., vitamin B6 (pyridoxine, ≥98%), L-ascorbic acid (ACS reagent grade, ≥99%), caffeine, Triton X-100, sodium dodecyl sulfate (SDS), titanium dioxide, starch, cellulose, and magnesium stearate were acquired from Sigma-Aldrich, whereas glucose and cetrimonium bromide (CTAB) were purchased from Avantor Performance Materials Poland S.A. All reagents were of analytical grade and used without further purification.

3.2. Instrumentation

The infrared spectrum of zeolite Y was recorded in the mid-infrared region (4000–400 cm−1) using a Vertex 70v FT-IR spectrometer (Bruker, Billerica, MA, USA) with KBr pellets (256 scans, 4 cm−1 resolution). In turn, Raman spectrum of CNHs was collected using a WITec Alpha 300 M+ (WITec GmbH, Ulm, Germany) with a 600 groove mm−1 grating and a 532 nm laser with adjusted power to prevent sample degradation. A Zeiss Epiplan Neofluar (Carl Zeiss AG, Oberkochen, Germany) 20× long-working-distance objective (laser spot ~0.5 μm) was used. XY maps (40 × 40 μm2, 0.5 μm step size, 0.5 s per spectrum) were acquired, and data were processed using WITec Project Five 5.3 PRO software. The texture features of ZY and CNHs were established by multipoint N2 adsorption/desorption at 77 K using an ASAP 2010 (Micromeritics, Norcross, GA, USA). The samples were degassed at 623 K for 24 h prior to analysis. The specific surface area (SSA) and pore size distribution were calculated using the Brunauer–Emmett–Teller (BET) and Barrett–Joyner–Halenda (BJH) methods, respectively. The morphology of ZY, CNHs, and the modifying layer was examined using a Thermo Scientific Scios 2 DualBeam, an ultra-high-resolution analytical-focused ion beam scanning electron microscope (FIB-SEM). Elemental analysis of ZY before and after Ni(II) cation exchange was performed using an energy-dispersive X-ray spectroscopy (EDS) attachment.
Voltammetric measurements were performed using an M161 electrochemical analyzer with an M164D electrode stand (mtm-anko, Kraków, Poland) and EALab 2.1 software. A 10 mL three-electrode cell consisted of a bare or modified GCE (MF-2012, φ = 3 mm, BASi, Bioanalytical Systems, Inc., West Lafayette, IN, USA) as the working electrode, a double-junction Ag | AgCl | 3 M KCl reference electrode (MINERAL, Gliwice, Poland), and a Pt wire auxiliary electrode. Electrochemical impedance spectroscopy (EIS) was performed using a μAUTOLAB III analyzer (EcoChemie, Utrecht, The Netherlands) with NOVA 2.0 software for data acquisition. The solutions were stirred at ~200 rpm with a Teflon®-coated magnetic stir bar (Kartell, Noviglio, Italy) and deoxygenated by argon purging before measurements in the negative potential range. The pH value of McIlvaine buffer was adjusted using a SevenCompact S210 pH meter (Mettler Toledo, Greifensee, Switzerland).

3.3. Procedures

3.3.1. Zeolite Y Modification and NiZY/CNHs Electrode Fabrication

To enhance electrocatalytic activity, zeolite Y was converted to its nickel form (NiZY) by an ion exchange process carried out in the following steps: (1) 0.5 g of ZY was treated with 5 mL of 0.2 M Ni(NO3)2·6H2O solution and shaken (2500 rpm) for 24 h at 25 °C; (2) after centrifuging at 13,500 rpm (10 min), the procedure was repeated twice, each time using a fresh Ni(II) solution; (3) the resulting material was thoroughly rinsed with double-distilled water to remove surface adhering salts and dried at 40 °C for 24 h in a drying oven. The incorporation of nickel ions in the zeolite Y structure was confirmed by SEM/EDS analysis.
To prepare the modified sensor, the CNHs without any pretreatment and NiZY were ground together in an agate mortar in a 1:1 w/w ratio, after which 10 mg of homogenized mixture was transferred to an Eppendorf tube and suspended in 500 µL of organic solvents, i.e., acetone and dichloromethane (2:3 v/v ratio), containing 5 mg of polystyrene and homogenized for 30 min (1800 rpm) using a Vortex Multi Speed MSV-3500 (BioSan, Riga, Latvia). In the next step, 10 μL of the modifying suspension was carefully dropped onto a polished surface of GCE (0.3 μm alumina, Buehler Micropolish II, Lake Bluff, IL, USA) and left for the solvent to evaporate under laboratory conditions for 24 h. Thereafter, the obtained NiZY/CNHs-GCE was stored under ambient conditions and used directly in measurements without any further activation. Following the same approaches, single-component electrodes containing 10 mg of either NiZY or CNHs were fabricated, resulting in NiZY-GCE and CNHs-GCE, respectively.

3.3.2. Sample Preparation and Analysis

Dietary supplements in the form of tablets containing riboflavin Apteo Witamina B2 (3 mg VB2/tablet; Synoptis Pharma Sp. z o.o., Warsaw, Poland) and Panawit Witamina B2 (10 mg VB2/tablet; PANAWIT Sp. z o.o., Kraków, Poland), as well as folic-acid-containing dietary supplements, Olimp Labs Kwas foliowy (400 µg VB9/tablet; OLIMP LABORATORIES Sp. z o.o., Dębica, Poland) and ActiFolin (800 µg VB9/tablet; POLSKI LEK Sp. z o.o., Warsaw, Poland), were purchased from a local pharmacy. For their voltammetric analysis, a stock solution of each vitamin (10 mg L−1) was prepared by weighing and grinding five tablets of each brand in an agate mortar. An appropriate portion of the resulting powder was transferred to a 100 mL (VB2) or 25 mL (VB9) amber gold volumetric flask and dissolved in double-distilled water. The solutions were then homogenized by sonication for 15 min and filtered through a 0.22 μm Mixed Cellulose Ester (MCE) syringe filter (Alchem Grupa, Nowa Sól, Poland) to remove insoluble excipients.
For the DPV analysis of VB2, 20 µL of corresponding dietary supplement solution was added to 5 mL of McIlvaine buffer solution (pH 3.4), and measurements were carried out using the standard addition method. At each addition step, four consecutive DPV scans were recorded, followed by background correction, and the mean peak current was determined. The VB2 content in the analyzed tablets was calculated from the extrapolated intercept of the resulting calibration plot. Vitamin B9-containg dietary supplements were analyzed in the same manner, except that McIlvaine buffer at pH 2.8 was used.

3.3.3. Standard Electrochemical Procedure

The modified electrodes were electrochemically characterized by electrochemical impedance spectroscopy (EIS) and cyclic voltammetry (CV) measurements performed in a 0.1 mol L−1 KCl solution containing 1 mmol L−1 of the reversible redox probe in the form of [Fe(CN)6]4−. Sinusoidal signals of frequency ranging from 100 kHz to 25 mHz and 10 mV amplitude, superimposed on the formal potential of K4[Fe(CN)6], were used in EIS measurements. Electrochemical figures of merit were obtained by applying an equivalent electrical circuit (EEC) model to fit the experimentally recorded impedance spectra.
To establish the electrochemical behavior of VB2 on the surface of the NiZY/CNHs-GCE, both CV and DPV analysis were applied. Cyclic voltammograms were recorded in McIlvaine buffer at pH 3.4 containing 1 mg L−1 of VB2, in the potential range of −0.1 to −0.6 V with a scan rate (v) that changed from 0.006 to 0.5 V s−1. In turn, DPV measurements were conducted in McIlvaine buffer solutions over a pH range of 2.6 to 3.6 in the presence of 0.2 mg L−1 VB2. The analytical performance and quantitative analysis of VB2 were performed using DPV in both anodic and cathodic directions. Measurements were carried out under optimized experimental conditions, which included McIlvaine buffer at pH 3.4, step potential (Es) = 3 mV, pulse amplitude (dE) = 40 mV, waiting time (tw) = 10 ms, and current sampling time (ts) = 10 ms. For VB2 and VB9 simultaneous determination, the cathodic response of NiZY/CNHs-GCE was recorded in McIlvaine buffer at pH 3.0 under the same optimized instrumental parameters of the DPV technique. All measurements were performed in a 5 mL electrochemical cell, recording at least four consecutive voltammograms, and the background-corrected peak currents were averaged for the evaluation of the obtained data.

4. Conclusions

In this work, a novel zeolite-modified electrochemical sensor was developed by modifying the surface of GCEs with a composite of Ni-exchanged zeolite Y and carbon nanohorns, The resulting NiZY/CNHs-GCE demonstrated markedly improved electrochemical activity toward the reduction of VB2 molecules, arising from the synergistic effect of both materials, which provides high surface area, enhanced probe accessibility, and efficient electron-transfer pathways. Morphological, structural, and voltammetric analyses confirmed the effective integration of NiZY and CNHs, resulting in the formation of an active sensing layer with favorable electrochemical parameters. As indicated by CV and DPV measurements, the redox reaction of VB2 molecules at the NiZY/CNHs-GCE follows adsorption-controlled kinetics, involving the transfer of two electrons and two protons.
The proposed DPV-based voltammetric method for the sensitive quantification of VB2 molecules allows us to achieve LOD values in the nanomolar range (8.6 × 10−9 mol L−1), highlighting its suitability for trace analysis. At the same time, the proposed NiZY/CNHs-GCE offers a notable analytical advantage by enabling the simultaneous quantification of VB2 and VB9, which is particularly valuable in complex biological or pharmaceutical matrices where both vitamins may be present at low concentrations. The close correspondence of the measured vitamin levels with the declared values, expressed as an RE higher than −6.2%, confirms the accuracy determinations of both analytes. The findings highlight the promising applicability of the developed sensor in routine quality control of dietary supplements, providing a reliable and accurate platform for trace-level vitamin analysis.

Author Contributions

Conceptualization, K.F.; methodology, K.F.; validation, K.F. and J.N.; formal analysis, K.F. and J.S.; investigation, K.F. and J.N.; resources, K.F. and B.B.; writing—original draft preparation, K.F., J.S. and B.B.; writing—review and editing, K.F., R.P. and B.B.; visualization, K.F.; supervision, B.B.; project administration, K.F.; funding acquisition, K.F. and R.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Science Centre, Poland (Project No. 2024/08/X/ST4/00778).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Powers, H.J. Riboflavin (Vitamin B-2) and Health. Am. J. Clin. Nutr. 2003, 77, 1352–1360. [Google Scholar] [CrossRef] [PubMed]
  2. Thakur, K.; Tomar, S.K.; Singh, A.K.; Mandal, S.; Arora, S. Riboflavin and Health: A Review of Recent Human Research. Crit. Rev. Food Sci. Nutr. 2017, 57, 3650–3660. [Google Scholar] [CrossRef]
  3. Pan, F.; Zhang, L.L.; Luo, H.J.; Chen, Y.; Long, L.; Wang, X.; Zhuang, P.T.; Li, E.M.; Xu, L.Y. Dietary Riboflavin Deficiency Induces Ariboflavinosis and Esophageal Epithelial Atrophy in Association with Modification of Gut Microbiota in Rats. Eur. J. Nutr. 2021, 60, 807–820. [Google Scholar] [CrossRef] [PubMed]
  4. Duyvis, M.G.; Hilhorst, R.; Laane, C.; Evans, D.J.; Schmedding, D.J.M. Role of Riboflavin in Beer Flavor Instability: Determination of Levels of Riboflavin and Its Origin in Beer by Fluorometric Apoprotein Titration. J. Agric. Food Chem. 2002, 50, 1548–1552. [Google Scholar] [CrossRef]
  5. Zajicek, J.L.; Tillitt, D.E.; Brown, S.B.; Brown, L.R.; Honeyfield, D.C.; Fitzsimons, J.D. A Rapid Solid-Phase Extraction Fluorometric Method for Thiamine and Riboflavin in Salmonid Eggs. J. Aquat. Anim. Health 2005, 17, 95–105. [Google Scholar] [CrossRef]
  6. Alvarado, U.; Zamora, A.; Liu, J.; Saldo, J.; Castillo, M. Rapid Quantification of Riboflavin in Milk by Front-Face Fluorescence Spectroscopy: A Preliminary Study. Foods 2020, 9, 6. [Google Scholar] [CrossRef]
  7. Ghasemi, J.; Abbasi, B.; Niazi, A.; Nadaf, E.; Mordai, A. Simultaneous Spectrophotometric Multicomponent Determination of Folic Acid, Thiamine, Riboflavin, and Pyridoxal by Using Double Divisor-Ratio Spectra Derivative-Zero Crossing Method. Anal. Lett. 2004, 37, 2609–2623. [Google Scholar] [CrossRef]
  8. López-de-Alba, P.L.; López-Martínez, L.; Cerdá, V.; Amador-Hernández, J. Simultaneous Determination and Classification of Riboflavin, Thiamine, Nicotinamide and Pyridoxine in Pharmaceutical Formulations, by UV-Visible Spectrophotometry and Multivariate Analysis. J. Braz. Chem. Soc. 2006, 17, 715–722. [Google Scholar] [CrossRef]
  9. Bartzatt, R.; Wol, T. Detection and Assay of Vitamin B-2 (Riboflavin) in Alkaline Borate Buffer with UV/Visible Spectrophotometry. Int. Sch. Res. Not. 2014, 2014, 453085. [Google Scholar] [CrossRef]
  10. Capo-Chichi, C.D.; Guéant, J.L.; Feillet, F.; Namour, F.; Vidailhet, M. Analysis of Riboflavin and Riboflavin Cofactor Levels in Plasma by High-Performance Liquid Chromatography. J. Chromatogr. B Biomed. Sci. Appl. 2000, 739, 219–224. [Google Scholar] [CrossRef]
  11. Hardwick, C.C.; Herivel, T.R.; Hernandez, S.C.; Ruane, P.H.; Goodrich, R.P. Separation, Identification and Quantification of Riboflavin and Its Photoproducts in Blood Products Using High-Performance Liquid Chromatography with Fluorescence Detection: A Method to Support Pathogen Reduction Technology. Photochem. Photobiol. 2004, 80, 609–615. [Google Scholar] [CrossRef]
  12. Viñas, P.; Balsalobre, N.; López-Erroz, C.; Hernández-Córdoba, M. Liquid Chromatographic Analysis of Riboflavin Vitamers in Foods Using Fluorescence Detection. J. Agric. Food Chem. 2004, 52, 1789–1794. [Google Scholar] [CrossRef]
  13. Cataldi, T.R.I.; Nardiello, D.; Carrara, V.; Ciriello, R.; De Benedetto, G.E. Assessment of Riboflavin and Flavin Content in Common Food Samples by Capillary Electrophoresis with Laser-Induced Fluorescence Detection. Food Chem. 2003, 82, 309–314. [Google Scholar] [CrossRef]
  14. Ravi, G.; Venkatesh, Y.P. Immunoassays for Riboflavin and Flavin Mononucleotide Using Antibodies Specific to D-Ribitol and D-Ribitol-5-Phosphate. J. Immunol. Methods 2017, 445, 59–66. [Google Scholar] [CrossRef]
  15. Lovander, M.D.; Lyon, J.D.; Parr, D.L.; Wang, J.; Parke, B.; Leddy, J. Critical Review—Electrochemical Properties of 13 Vitamins: A Critical Review and Assessment. J. Electrochem. Soc. 2018, 165, G18–G49. [Google Scholar] [CrossRef]
  16. Nezamzadeh-Ejhieh, A.; Pouladsaz, P. Voltammetric Determination of Riboflavin Based on Electrocatalytic Oxidation at Zeolite-Modified Carbon Paste Electrodes. J. Ind. Eng. Chem. 2014, 20, 2146–2152. [Google Scholar] [CrossRef]
  17. Khaloo, S.S.; Mozaffari, S.; Alimohammadi, P.; Kargar, H.; Ordookhanian, J. Sensitive and Selective Determination of Riboflavin in Food and Pharmaceutical Samples Using Manganese (III) Tetraphenylporphyrin Modified Carbon Paste Electrode. Int. J. Food Prop. 2016, 19, 2272–2283. [Google Scholar] [CrossRef]
  18. Mehmeti, E.; Stanković, D.M.; Chaiyo, S.; Švorc, Ľ.; Kalcher, K. Manganese Dioxide-Modified Carbon Paste Electrode for Voltammetric Determination of Riboflavin. Microchim. Acta 2016, 183, 1619–1624. [Google Scholar] [CrossRef]
  19. Bártová, M.; Bartoš, M.; Švancara, I.; Sýs, M. Shungite Paste Electrodes: Basic Characterization and Initial Examples of Applicability in Electroanalysis. Chemosensors 2024, 12, 118. [Google Scholar] [CrossRef]
  20. Hareesha, N.; Manjunatha, J.G. Elevated and Rapid Voltammetric Sensing of Riboflavin at Poly(Helianthin Dye) Blended Carbon Paste Electrode with Heterogeneous Rate Constant Elucidation. J. Iran. Chem. Soc. 2020, 17, 1507–1519. [Google Scholar] [CrossRef]
  21. Varun, D.N.; Manjunatha, J.G.; Hareesha, N.; Sandeep, S.; Mallu, P.; Karthik, C.S.; Prinith, N.S.; Sreeharsha, N.; Asdaq, S.M.B. Simple and Sensitive Electrochemical Analysis of Riboflavin at Functionalized Carbon Nanofiber Modified Carbon Nanotube Sensor. Monatshefte Fur Chem. 2021, 152, 1183–1191. [Google Scholar] [CrossRef]
  22. Sá, É.S.; Da Silva, P.S.; Jost, C.L.; Spinelli, A. Electrochemical Sensor Based on Bismuth-Film Electrode for Voltammetric Studies on Vitamin B2 (Riboflavin). Sens. Actuators B Chem. 2015, 209, 423–430. [Google Scholar] [CrossRef]
  23. Madhuvilakku, R.; Alagar, S.; Mariappan, R.; Piraman, S. Green One-Pot Synthesis of Flowers-like Fe3O4/RGO Hybrid Nanocomposites for Effective Electrochemical Detection of Riboflavin and Low-Cost Supercapacitor Applications. Sens. Actuators B Chem. 2017, 253, 879–892. [Google Scholar] [CrossRef]
  24. Sriramprabha, R.; Divagar, M.; Ponpandian, N.; Viswanathan, C. Tin Oxide/Reduced Graphene Oxide Nanocomposite-Modified Electrode for Selective and Sensitive Detection of Riboflavin. J. Electrochem. Soc. 2018, 165, B498–B507. [Google Scholar] [CrossRef]
  25. Selvarajan, S.; Suganthi, A.; Rajarajan, M. A Facile Synthesis of ZnO/Manganese Hexacyanoferrate Nanocomposite Modified Electrode for the Electrocatalytic Sensing of Riboflavin. J. Phys. Chem. Solids 2018, 121, 350–359. [Google Scholar] [CrossRef]
  26. Derakhshan, M.; Shamspur, T.; Molaakbari, E.; Mostafavi, A.; Saljooqi, A. Fabrication of a Novel Electrochemical Sensor for Determination of Riboflavin in Different Drink Real Samples. Russ. J. Electrochem. 2020, 56, 181–188. [Google Scholar] [CrossRef]
  27. Jesu Amalraj, A.J.; Wang, S.F. An Effective Morphology Controlled Hydrothermal Synthesis of Bi2WO6 and Its Application in Riboflavin Electrochemical Sensor. Colloids Surf. A Physicochem. Eng. Asp. 2022, 648, 129183. [Google Scholar] [CrossRef]
  28. Wu, S.H.; Sun, J.J.; Lin, Z.B.; Wu, A.H.; Zeng, Y.M.; Guo, L.; Zhang, D.F.; Dai, H.M.; Chen, G.N. Adsorptive Stripping Analysis of Riboflavin at Electrically Heated Graphite Cylindrical Electrodes. Electroanalysis 2007, 19, 2251–2257. [Google Scholar] [CrossRef]
  29. Ensafi, A.A.; Heydari-Bafrooei, E.; Amini, M. DNA-Functionalized Biosensor for Riboflavin Based Electrochemical Interaction on Pretreated Pencil Graphite Electrode. Biosens. Bioelectron. 2012, 31, 376–381. [Google Scholar] [CrossRef]
  30. Nagarajan, S.; Vairamuthu, R. Electrochemical Detection of Riboflavin Using Tin-Chitosan Modified Pencil Graphite Electrode. J. Electroanal. Chem. 2021, 891, 115235. [Google Scholar] [CrossRef]
  31. Sedhu, N.; Jagadeesh Kumar, J.; Sivaguru, P.; Raj, V. Electrochemical Detection of Riboflavin in Pharmaceutical and Food Samples Using in Situ Electropolymerized Glycine Coated Pencil Graphite Electrode. J. Electroanal. Chem. 2023, 928, 117037. [Google Scholar] [CrossRef]
  32. Kadara, R.O.; Haggett, B.G.D.; Birch, B.J. Disposable Sensor for Measurement of Vitamin B2 in Nutritional Premix, Cereal, and Milk Powder. J. Agric. Food Chem. 2006, 54, 4921–4924. [Google Scholar] [CrossRef]
  33. Riman, D.; Avgeropoulos, A.; Hrbac, J.; Prodromidis, M.I. Sparked-Bismuth Oxide Screen-Printed Electrodes for the Determination of Riboflavin in the Sub-Nanomolar Range in Non-Deoxygenated Solutions. Electrochim. Acta 2015, 165, 410–415. [Google Scholar] [CrossRef]
  34. Muthusankar, G.; Rajkumar, C.; Chen, S.M.; Karkuzhali, R.; Gopu, G.; Sangili, A.; Sengottuvelan, N.; Sankar, R. Sonochemical Driven Simple Preparation of Nitrogen-Doped Carbon Quantum Dots/SnO2 Nanocomposite: A Novel Electrocatalyst for Sensitive Voltammetric Determination of Riboflavin. Sens. Actuators B Chem. 2019, 281, 602–612. [Google Scholar] [CrossRef]
  35. Papavasileiou, A.V.; Hoder, T.; Medek, T.; Prodromidis, M.I.; Hrbac, J. Sensitive Riboflavin Sensing Using Silver Nanoparticles Deposited onto Screen-Printed Electrodes via Controlled-Energy Spark Discharges. Talanta 2023, 258, 124409. [Google Scholar] [CrossRef]
  36. Bai, J.; Ndamanisha, J.C.; Liu, L.; Yang, L.; Guo, L. Voltammetric Detection of Riboflavin Based on Ordered Mesoporous Carbon Modified Electrode. J. Solid State Electrochem. 2010, 14, 2251–2256. [Google Scholar] [CrossRef]
  37. Sonkar, P.K.; Ganesan, V.; Sen Gupta, S.K.; Yadav, D.K.; Gupta, R.; Yadav, M. Highly Dispersed Multiwalled Carbon Nanotubes Coupled Manganese Salen Nanostructure for Simultaneous Electrochemical Sensing of Vitamin B2 and B6. J. Electroanal. Chem. 2017, 807, 235–243. [Google Scholar] [CrossRef]
  38. Sangili, A.; Veerakumar, P.; Chen, S.M.; Rajkumar, C.; Lin, K.C. Voltammetric Determination of Vitamin B 2 by Using a Highly Porous Carbon Electrode Modified with Palladium-Copper Nanoparticles. Microchim. Acta 2019, 186, 299. [Google Scholar] [CrossRef]
  39. Guler, M.; Meydan, I.; Seckin, H. Electrochemical Characterization of Vitamin B2 (Riboflavin) at Ag Nanoparticles Synthesized by Green Chemistry Dispersed on Reduced Graphene Oxide. Diam. Relat. Mater. 2023, 135, 109875. [Google Scholar] [CrossRef]
  40. Gomes-Junior, P.C.; de Lima Augusto, K.K.; Longatto, G.P.; de Oliveira Gonçalves, R.; Silva, T.A.; Cavalheiro, É.T.G.; Fatibello-Filho, O. Ultrasmall Platinum Nanoparticles Synthesized in Reline Deep Eutectic Solvent Explored towards the Voltammetric Sensing of Riboflavin in Beverages and Biological Fluids. Sens. Actuators B Chem. 2023, 395, 134489. [Google Scholar] [CrossRef]
  41. Hajian, A.; Rafati, A.A.; Afraz, A.; Najafi, M. Electrosynthesis of High-Density Polythiophene Nanotube Arrays and Their Application for Sensing of Riboflavin. J. Mol. Liq. 2014, 199, 150–155. [Google Scholar] [CrossRef]
  42. Manjunatha, J.G.; Raril, C.; Hareesha, N.; Charithra, M.M.; Pushpanjali, P.A.; Tigari, G.; Ravishankar, D.; Mallappaji, S.C.; Gowda, J. Electrochemical Fabrication of Poly (Niacin) Modified Graphite Paste Electrode and Its Application for the Detection of Riboflavin. Open Chem. Eng. J. 2020, 14, 90–98. [Google Scholar] [CrossRef]
  43. Abhayashri Kamath, K.; Manjunatha, J.G.; Girish, T.; Sillanpää, M.; TIGHEZZA, A.M.; Albaqami, M.D. Sensitive Electrochemical Determination of Riboflavin at Simple and Low-Cost Poly (Valine) Modified Graphite Paste Electrode. Inorg. Chem. Commun. 2022, 143, 109811. [Google Scholar] [CrossRef]
  44. AlSuhaimi, A.O.; Althumayri, K.; Alessa, H.; Sayqal, A.; Mogharbel, A.T.; Alsehli, B.R.; El-Metwaly, N.M. Zeolite Integrated Carbon Paste Sensors for Sensitive and Selective Differential Pulse Voltammetric Determination of Carprofen. Microchem. J. 2023, 195, 109525. [Google Scholar] [CrossRef]
  45. Thatikayala, D.; Noori, M.T.; Min, B. Zeolite-Modified Electrodes for Electrochemical Sensing of Heavy Metal Ions–Progress and Future Directions. Mater. Today Chem. 2023, 29, 101412. [Google Scholar] [CrossRef]
  46. Hashemi, H.S.; Nezamzadeh-Ejhieh, A.; Karimi-Shamsabadi, M. A Novel Cysteine Sensor Based on Modification of Carbon Paste Electrode by Fe(II)-Exchanged Zeolite X Nanoparticles. Mater. Sci. Eng. C 2016, 58, 286–293. [Google Scholar] [CrossRef]
  47. Yang, J.; Zhou, Y.; Wang, H.J.; Zhuang, T.T.; Cao, Y.; Yun, Z.Y.; Yu, Q.; Zhu, J.H. Capturing Nitrosamines by Zeolite A: Molecular Recognition in Subnanometer Space. J. Phys. Chem. C 2008, 112, 6740–6748. [Google Scholar] [CrossRef]
  48. Lupulescu, A.I.; Kumar, M.; Rimer, J.D. A Facile Strategy to Design Zeolite L Crystals with Tunable Morphology and Surface Architecture. J. Am. Chem. Soc. 2013, 135, 6608–6617. [Google Scholar] [CrossRef]
  49. Porada, R.; Fendrych, K.; Baś, B. The Mn-Zeolite/Graphite Modified Glassy Carbon Electrode: Fabrication, Characterization and Analytical Applications. Electroanalysis 2020, 32, 1208–1219. [Google Scholar] [CrossRef]
  50. Senthilkumar, S.; Saraswathi, R. Electrochemical Sensing of Cadmium and Lead Ions at Zeolite-Modified Electrodes: Optimization and Field Measurements. Sens. Actuators B Chem. 2009, 141, 65–75. [Google Scholar] [CrossRef]
  51. Kawde, A.; Ismail, A.; Al-Betar, A.R.; Muraza, O. Novel Ce-Incorporated Zeolite Modified-Carbon Paste Electrode for Simultaneous Trace Electroanalysis of Lead and Cadmium. Microporous Mesoporous Mater. 2017, 243, 1–8. [Google Scholar] [CrossRef]
  52. Porada, R.; Fendrych, K.; Baś, B. Electrochemical Sensor Based on Ni-Exchanged Natural Zeolite/Carbon Black Hybrid Nanocomposite for Determination of Vitamin B6. Microchim. Acta 2021, 188, 323. [Google Scholar] [CrossRef]
  53. Fendrych, K.; Porada, R.; Baś, B. Ultrasensitive and Highly Selective Electrochemical Determination of Mesalazine in Pharmaceutical, Biological and Environmental Samples with the Use of Co-Mordenite/Mesoporous Carbon Modified Glassy Carbon Electrode. Electrochim. Acta 2023, 468, 143209. [Google Scholar] [CrossRef]
  54. Fendrych, K.; Porada, R.; Baś, B. A Novel Voltammetric Strategy of Furazidin Determination with the Use of Co-Ferrierite/Mesoporous Carbon Modified Glassy Carbon Electrode. J. Electrochem. Soc. Soc. 2025, 172, 017523. [Google Scholar] [CrossRef]
  55. Porada, R.; Wenninger, N.; Bernhart, C.; Fendrych, K.; Kochana, J.; Baś, B.; Kalcher, K.; Ortner, A. Targeted Modification of the Carbon Paste Electrode by Natural Zeolite and Graphene Oxide for the Enhanced Analysis of Paracetamol. Microchem. J. 2023, 187, 108455. [Google Scholar] [CrossRef]
  56. Fekrya, A.M.; Abdel-Gawad, S.A.; Azab, S.M.; Walcarius, A. A Sensitive Electrochemical Sensor for Moxifloxacin Hydrochloride Based on Nafion/Graphene Oxide/Zeolite Modified Carbon Paste Electrode. Electroanalysis 2021, 33, 964–974. [Google Scholar] [CrossRef]
  57. Kaduk, J.A.; Faber, J. Crystal Structure of Zeolite Y as a Functon of Ion Exchange. Ridaku J. 1995, 12, 14–34. [Google Scholar]
  58. Król, M.; Mozgawa, W.; Jastrzbski, W.; Barczyk, K. Application of IR Spectra in the Studies of Zeolites from D4R and D6R Structural Groups. Microporous Mesoporous Mater. 2012, 156, 181–188. [Google Scholar] [CrossRef]
  59. Król, M.; Mozgawa, W.; Barczyk, K.; Bajda, T.; Kozanecki, M. Changes in the Vibrational Spectra of Zeolites Due to Sorption of Heavy Metal Cations. J. Appl. Spectrosc. 2013, 80, 644–650. [Google Scholar] [CrossRef]
  60. Pagona, G.; Mountrichas, G.; Rotas, G.; Karousis, N.; Pispas, S.; Tagmatarchis, N. Properties, Applications and Functionalisation of Carbon Nanohorns. Int. J. Nanotechnol. 2009, 6, 176–195. [Google Scholar] [CrossRef]
  61. Bard, A.J.; Faulkner, L.R. Electrochemical Methods: Fundamentals and Applications; Wiley: New York, NY, USA, 2001; ISBN 9780387331508. [Google Scholar]
  62. Narmaeva, G.; Aronbaev, S.; Aronbaev, D. Achievements and Problems of Electrode Modification for Voltammetry. Int. J. Res.-Granthaalayah 2018, 6, 368–381. [Google Scholar] [CrossRef]
Figure 1. (A) FT-IR spectra of zeolite Y before and after ion-exchange with Ni(II) cations in the middle infrared region. (B) Raman spectrum of CNHs in the range of 500–3500 cm−1. SEM image of (C) Ni-zeolite Y and the corresponding (D) EDS spectrum, (E) CNHs, and (F) NiZY/CNH layer deposited on the surface of the GCE.
Figure 1. (A) FT-IR spectra of zeolite Y before and after ion-exchange with Ni(II) cations in the middle infrared region. (B) Raman spectrum of CNHs in the range of 500–3500 cm−1. SEM image of (C) Ni-zeolite Y and the corresponding (D) EDS spectrum, (E) CNHs, and (F) NiZY/CNH layer deposited on the surface of the GCE.
Ijms 26 10469 g001
Figure 2. (A) The CV voltammograms recorded on bare GCE (black line), NiZY-GCE (pink line), CNHs-GCE (dark blue line), and NiZY/CNHs-GCE (green line) for redox probe of K4[Fe(CN)6] (1 mmol L−1) in 0.1 mol L−1 KCl solution (scan rate v = 0.05 V s−1) and (B) corresponding impedance spectra in the form of a Nyquist plot (inset: Randles cell equivalent circuit). (C) Comparison of DP curves obtained for 50 µg L−1 VB2 on the GCE, NiZY-GCE, CNH-GCE, and NiZY/CNHs-GCE (C) before and (D) after the background subtraction step. Dashed lines denote the background signal. DPV experimental conditions: Ep = −0.1 V, Ek = −0.5 V, Es = 3 mV, dE = 40 mV, ts = 10 ms, and tw = 10 ms. Supporting electrolyte: McIlvaine buffer solution (pH 3.4).
Figure 2. (A) The CV voltammograms recorded on bare GCE (black line), NiZY-GCE (pink line), CNHs-GCE (dark blue line), and NiZY/CNHs-GCE (green line) for redox probe of K4[Fe(CN)6] (1 mmol L−1) in 0.1 mol L−1 KCl solution (scan rate v = 0.05 V s−1) and (B) corresponding impedance spectra in the form of a Nyquist plot (inset: Randles cell equivalent circuit). (C) Comparison of DP curves obtained for 50 µg L−1 VB2 on the GCE, NiZY-GCE, CNH-GCE, and NiZY/CNHs-GCE (C) before and (D) after the background subtraction step. Dashed lines denote the background signal. DPV experimental conditions: Ep = −0.1 V, Ek = −0.5 V, Es = 3 mV, dE = 40 mV, ts = 10 ms, and tw = 10 ms. Supporting electrolyte: McIlvaine buffer solution (pH 3.4).
Ijms 26 10469 g002
Figure 3. (A) CV curves registered in McIlvaine buffer solutions (pH 3.4) in the presence of 1 mg L−1 of VB2 on the NiZY/CNHs-GCE with a scan rate from 0.006 to 0.5 V s−1 (arrows—scan direction; solid lines—the cathodic scan; dashed lines—the anodic scan). Inset: the relationships between the peak current Ip and the scan rate v. (B) DPVs recorded for the reduction of 0.2 mg L−1 VB2 in McIlvaine buffer solution at pH from 2.6 to 3.6 after background subtraction. (C) The cathodic peak potential (Epc) as a function of the pH of the supporting electrolyte. (D) The proposed mechanism of the VB2 redox reaction.
Figure 3. (A) CV curves registered in McIlvaine buffer solutions (pH 3.4) in the presence of 1 mg L−1 of VB2 on the NiZY/CNHs-GCE with a scan rate from 0.006 to 0.5 V s−1 (arrows—scan direction; solid lines—the cathodic scan; dashed lines—the anodic scan). Inset: the relationships between the peak current Ip and the scan rate v. (B) DPVs recorded for the reduction of 0.2 mg L−1 VB2 in McIlvaine buffer solution at pH from 2.6 to 3.6 after background subtraction. (C) The cathodic peak potential (Epc) as a function of the pH of the supporting electrolyte. (D) The proposed mechanism of the VB2 redox reaction.
Ijms 26 10469 g003
Figure 4. (A) Cathodic and (B) anodic DPV calibration voltammogram (scan direction indicated by arrow) registered on the NiZY/CNH-GCE for the increasing concentration of VB2 from 0 (dashed line) to 0.2 mg L−1 (black lines) with curves after subtracted background current (green lines). Inset: corresponding calibration plots. (C) DP cathodic voltammograms of mixture solution containing VB2 and VB9 at the concentration from 0 to 0.2 mg L−1. (D) VB2 and (E) VB9 curves obtained after baseline correction with corresponding calibration graphs. Experimental conditions as described in Section 3.3.3.
Figure 4. (A) Cathodic and (B) anodic DPV calibration voltammogram (scan direction indicated by arrow) registered on the NiZY/CNH-GCE for the increasing concentration of VB2 from 0 (dashed line) to 0.2 mg L−1 (black lines) with curves after subtracted background current (green lines). Inset: corresponding calibration plots. (C) DP cathodic voltammograms of mixture solution containing VB2 and VB9 at the concentration from 0 to 0.2 mg L−1. (D) VB2 and (E) VB9 curves obtained after baseline correction with corresponding calibration graphs. Experimental conditions as described in Section 3.3.3.
Ijms 26 10469 g004
Figure 5. Experimental voltammograms (black) and background-corrected curves (color marked) for the analysis of (A) Apteo Witamina B2, (B) Panawit Witamina B2, (C) Olimp Labs Kwas foliowy, and (D) ActiFolin. Inset: corresponding calibration plots. Experimental conditions as described in Section 3.3.3.
Figure 5. Experimental voltammograms (black) and background-corrected curves (color marked) for the analysis of (A) Apteo Witamina B2, (B) Panawit Witamina B2, (C) Olimp Labs Kwas foliowy, and (D) ActiFolin. Inset: corresponding calibration plots. Experimental conditions as described in Section 3.3.3.
Ijms 26 10469 g005
Table 1. Textural parameters of employed zeolite Y and CNH.
Table 1. Textural parameters of employed zeolite Y and CNH.
MaterialSurface Area
[m2 g−1]
Pore Volume
[cm3 g−1]
Average Pore Diameter
[nm]
SBETSmicroSextVmicro+mezoVmicro
Zeolite Y1001.81992.659.160.5110.4611.95
Carbon nanohorns460.65176.27284.380.1520.0750.14
SBET—BET surface area; Smicro—micropore area; Sext—external surface area; Vmicro+mezo—micropore + mezopore volume; Vmicro—micropore volume.
Table 2. The comparison of electrochemical properties of tested electrodes achieved from CV and EIS measurements.
Table 2. The comparison of electrochemical properties of tested electrodes achieved from CV and EIS measurements.
ParameterUnitWorking Electrode
GCENiZY-GCECNHs-GCENiZY/CNHs-GCE
Formal potential, EfmV138164152160
Peak separation, ∆EmV72261360536
Anodic peak current, IpaµA9.61.021.09.8
Anodic to cathodic peak current, Ipa/Ipc-0.990.91.21.4
Charge-transfer resistance, Rct1.470.712.248.0
Warburg coefficient, σkΩ s−1/23.156.702.902.80
Electroactive surface area, Ael mm28.94.29.710.0
Double-layer capacitance, Cdl µF cm−224.715.76.34.2
Heterogeneous rate constant, ksm s−14.3 × 1051.6 × 10−64.5 × 10−61.4 × 10−6
Table 3. Summary of analytical parameters for VB2 in individual and simultaneous determination with VB9.
Table 3. Summary of analytical parameters for VB2 in individual and simultaneous determination with VB9.
ParameterUnitVB2VB2
in the Presence
of VB9
VB9
in the Presence
of VB2
CathodicAnodic
Linear rangemg L−10.01–0.20.01–0.20.01–0.20.01–0.16
Intercept aµA0.008 ± 0.0050.014 ± 0.0060.027 ± 0.0060.012 ± 0.002
Slope bµA L mg−15.10 ± 0.045.51 ± 0.053.82 ± 0.061.21 ± 0.03
r-0.99930.99910.99790.9964
LODµg L−1 (nM)3.2 (8.6)3.6 (9.5)5.2 (13.8)5.4 (12.3)
LOQµg L−1 (nM)9.8 (26.1)10.9 (28.9)15.7 (41.8)16.5 (37.3)
Figure (A)(B)(D)(E)
Table 4. Analytical performances of voltammetric methods for determination of VB2 molecules.
Table 4. Analytical performances of voltammetric methods for determination of VB2 molecules.
ElectrodeTechniqueLinear Range
[µmol L−1]
LOD
[nmol L−1]
SamplesRef.
1 Co2+-Y/CPECV1.7–34710Multivitamin tablet[16]
2 MnTPP/CPEDPV0.01–108.0Food, pharmaceuticals[17]
3 MnO2/CPEDPV0.02–915Tablet[18]
4 ShPE/MnO2DPV0.1–10.027.4-[19]
5 PHLD- MCPECV60–15040.2Pharmaceutical[20]
6 FCNF/CNTPEDPV5.0–60.015.35B-complex capsule[21]
7 BiFESWAdSV0.3–0.8100Oral solution, syrup, tablets[22]
1.0–9.0
8 Fe3O4/rGO/GCEDPV0.030–189Pharmaceutical,
nutrition products
[23]
1–100
9 SnO2/RGO/GCESWV0.1–150 34Pharmaceutical,
energy drink
[24]
10 ZnO/MnHCNF/GCEDPV0.2–3.0 0.0101Pharmaceutical,
milk powder
[25]
11 GO/Au/polyEAmVS/GCEDPV1–100 7.2Non-alcoholic beverage, energy drink[26]
12 Bi2WO6(PVP + NaOH)/GCEDPV0.03–457 3.65Almond milk, soymilk[27]
13 HGCESWV0.01–0.07 5Multivitamin
tablets
[28]
0.07–1.0
14 DNA-PGEDPV1.86–133956.5Multivitamin
tablets
[29]
15 PPGEASDPV0.008–2.340.202Multivitamin
tablets
[29]
16 Sn/Cs/PGESWV0.01–1.2 5.56Tablets, milk powder[30]
17 PGl/PGESWV0.02–0.45 1.24Pharmaceutical, foods[31]
18 SPEDPV2.66–61.12390Foods[32]
19 sparked-BiSPEsSWV0.001–0.10.7Pharmaceutical[33]
20 N-CQD/SnO2/SPCEDPV0.05–306 8 Tablets, milk powder[34]
21 AgNP-SPEDPV0.0019–0.1 0.56Pharmaceutical,
energy drink
[35]
22 OMC/GCECV0.4–1.020Vitamin tablets[36]
23 GC/MWCNTs-MnIIIsalenDPV1.0–400730Injection, tablet[37]
24 Pd-Cu@NSC/SPCEDPV0.004–0.10.0076Tablets, milk powder[38]
0.02–9.0
25 Ag/rGO/GCEDPV0.002–2.20.6Pharmaceutical [39]
26 USPtNPs-DES/MWCNT/GCESWV0.02–1.21.8Energy drink, biological fluids[40]
27 PTN/GCEDPV0.01–653Human plasma[41]
28 PNNMGPELSV5.0–65.0782 Tablets[42]
29 EP(VLN)MGPECV2.0–40.0286.9Pharmaceutical[43]
NiZY/CNHs-GCEDPV0.027–0.538.6Dietary supplementsThis work
1 Co2+-Y-zeolite-modified CPE; 2 manganese tetraphenylporphyrin-modified CPE; 3 manganese dioxide-modified CPE; 4 shungite paste electrode modified with MnO2 redox mediator; 5 polymerized-helianthin-dye-modified CPE; 6 functionalized carbon nanofiber and carbon nanotube composite paste electrode; 7 bismuth-film electrode; 8 Fe3O4 anchored reduced graphene oxide-modified GCE;9 tin oxide/reduced graphene oxide-modified GCE; 10 ZnO-manganese hexacyanoferrate-modified GCE; 11 graphene oxides/Au-nanoparticles/ionic liquid polymer of vinyl sulfonic acid sodium salt and methyl imidazolium-modified GCE; 12 bismuth tungstate (NaOH and poly(vinyl pyrrolidone)-modified GCE; 13 electrically heated graphite cylindrical electrode; 14 DNA-modified PGE; 15 pretreated PGE; 16 tin incorporated chitosan polymer matrix coated PGE; 17 polyglycine-coated PGE; 18 screen-printed electrode; 19 sparked-bismuth oxide SPE; 20 nitrogen-doped carbon quantum-dot-modified screen-printed carbon electrode; 21 Ag-nanoparticle-modified graphite SPE; 22 ordered mesoporous carbon-modified GCE; 23 multiwalled carbon nanotubes coupled with manganese salen-modified GCE; 24 palladium–copper nanoparticle-modified highly porous carbon electrode; 25 silver nanoparticles/reduced graphene oxide-modified GCE; 26 ultrasmall platinum nanoparticles synthesized in reline deep eutectic solvent/multi-walled carbon nanotube-modified GCE; 27 polythiophene-nanotube-modified GCE; 28 poly (niacin)-modified graphite paste electrode; 29 electrochemically polymerized valine-modified graphite paste electrode. SWAdSV—square-wave adsorption stripping voltammetry; ASDPV—differential pulse adsorptive stripping voltammetry; SWV—square-wave voltammetry.
Table 5. Results of VB2 and VB9 determination in dietary supplements.
Table 5. Results of VB2 and VB9 determination in dietary supplements.
VitaminSampleAmount of VB2 [mg per Tablet]RE [%]RSD [%]
Declared Found   x ¯ ± S
VB2Apteo Witamina B232.90 ± 0.06−3.32.1
Panawit Witamina B2109.38 ± 0.12−6.21.3
VB9Olimp Labs Kwas foliowy0.40.411 ± 0.0082.71.9
ActiFolin0.80.813 ± 0.0031.60.4
x ¯ —mean value; s—standard deviation; RSD% = s/ x ¯ × 100%.
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

Fendrych, K.; Nyrka, J.; Smajdor, J.; Piech, R.; Baś, B. Nanostructured Ni-Zeolite Y and Carbon Nanohorns Electrode for Sensitive Electrochemical Determination of B-Group Vitamins. Int. J. Mol. Sci. 2025, 26, 10469. https://doi.org/10.3390/ijms262110469

AMA Style

Fendrych K, Nyrka J, Smajdor J, Piech R, Baś B. Nanostructured Ni-Zeolite Y and Carbon Nanohorns Electrode for Sensitive Electrochemical Determination of B-Group Vitamins. International Journal of Molecular Sciences. 2025; 26(21):10469. https://doi.org/10.3390/ijms262110469

Chicago/Turabian Style

Fendrych, Katarzyna, Justyna Nyrka, Joanna Smajdor, Robert Piech, and Bogusław Baś. 2025. "Nanostructured Ni-Zeolite Y and Carbon Nanohorns Electrode for Sensitive Electrochemical Determination of B-Group Vitamins" International Journal of Molecular Sciences 26, no. 21: 10469. https://doi.org/10.3390/ijms262110469

APA Style

Fendrych, K., Nyrka, J., Smajdor, J., Piech, R., & Baś, B. (2025). Nanostructured Ni-Zeolite Y and Carbon Nanohorns Electrode for Sensitive Electrochemical Determination of B-Group Vitamins. International Journal of Molecular Sciences, 26(21), 10469. https://doi.org/10.3390/ijms262110469

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