[go: up one dir, main page]

Next Article in Journal
Emission Ellipsometry Study in Polymeric Interfaces Based on Poly(3-Hexylthiophene), [6,6]-Phenyl-C61-Butyric Acid Methyl Ester, and Reduced Graphene Oxide
Next Article in Special Issue
Impact of Dispersive Solvent and Temperature on Supercapacitor Performance of N-Doped Reduced Graphene Oxide
Previous Article in Journal
Miscanthus-Derived Biochar as a Platform for the Production of Fillers for the Improvement of Mechanical and Electromagnetic Properties of Epoxy Composites
Previous Article in Special Issue
A Nitrogen/Oxygen Dual-Doped Porous Carbon with High Catalytic Conversion Ability toward Polysulfides for Advanced Lithium–Sulfur Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Novel Non-Enzymatic Efficient H2O2 Sensor Utilizing δ-FeOOH and Prussian Blue Anchoring on Carbon Felt Electrode

by
Karoline S. Nantes
1,
Ana L. H. K. Ferreira
1,
Marcio C. Pereira
1,
Francisco G. E. Nogueira
2 and
André S. Afonso
1,*
1
Institute of Science, Engineering, and Technology, Federal University of Jequitinhonha and Mucuri Valleys (UFVJM), Teófilo Otoni 39803-371, Brazil
2
Department of Chemical Engineering, Federal University of São Carlos (UFSCar), São Carlos 13565-905, Brazil
*
Author to whom correspondence should be addressed.
Submission received: 3 August 2024 / Revised: 29 August 2024 / Accepted: 4 September 2024 / Published: 9 September 2024
Figure 1
<p>(<b>A</b>) Potentiodynamic growth of PB nanoparticles on CF electrode in N<sub>2</sub>-saturated 0.1 M KCl solution at pH 2.0 with 5 mM K<sub>3</sub>Fe(CN)<sub>6</sub> and 22 mg of δ-FeOOH. The potential window was set from −0.2 V to 0.7 V at a scan rate of 50 mV s<sup>−1</sup>, applying 30 scans. The arrows indicate scans increase. (<b>B</b>) CV of FC/PB-FeOOH electrode in N<sub>2</sub>-saturated 0.1 M KCl at pH 2.0 at a scan rate of 50 mV s<sup>−1</sup>. The CF was subjected to the following treatments: (a) i, (b) ii, and (c) iii.</p> ">
Figure 2
<p>(<b>A</b>) Evaluation by the CV of PB synthesized in different pH levels of the precursor solution. (<b>B</b>) PB synthesized with (a) 44 mg, (b) 22 mg, or (c) 11 mg of δ-FeOOH. (<b>C</b>) Effect of the number of scans on the synthesis of PB: (a) background, (b) 15 scans, (c) 30 scans, (d) 50 scans, (e) 100 scans, and (f) 150 scans. N<sub>2</sub>-saturated 0.1 M KCl solution at pH 2.0 was the medium in which CVs were performed. The potential range was from −0.2 V to 0.7 V, and the scan rate was 50 mV s<sup>−1</sup>.</p> ">
Figure 3
<p>The XRD pattern for the PB-FeOOH film was deposited on FTO.</p> ">
Figure 4
<p>SEM micrographs of (<b>A</b>,<b>B</b>) CF electrode and (<b>C</b>,<b>D</b>) CF/PB-FeOOH electrode. (<b>E</b>) EDS spectrum of CF/PB-FeOOH electrode and (<b>F</b>) the weight % of EDS analysis.</p> ">
Figure 5
<p>(<b>A</b>) CV curves at 100 mV·s<sup>−1</sup> and (<b>B</b>) EIS plots at open-circuit potential using N<sub>2</sub>-saturated 0.1 M KCl solution at pH 2.0 containing 1 mM K<sub>3</sub>Fe(CN)<sub>6</sub> and 1 mM K<sub>4</sub>Fe(CN)<sub>6</sub>, (a) CF, and (b) CF/PB-FeOOH.</p> ">
Figure 6
<p>CV curves at a potential range from −0.2 V to 0.7 V at a scan rate of 100 mV·s<sup>−1</sup> in 0.1 M of PBS without or with 1.0 mM of H<sub>2</sub>O<sub>2</sub>. (<b>A</b>) CF/PB-FeOOH electrode and (<b>B</b>) CF/PB electrode.</p> ">
Figure 7
<p>(<b>A</b>) Amperometric response of the CF/PB–δ-FeOOH electrode in 0.1 M PBS at −0.2 V versus Ag|AgCl with successive additions of H<sub>2</sub>O<sub>2</sub> under stirring. (<b>B</b>) Corresponding calibration curve. The error bars show the standard deviation for <span class="html-italic">n</span> = 3. The arrow indicates the increase in concentration of H<sub>2</sub>O<sub>2</sub>.</p> ">
Figure 8
<p>Amperometric response of CF/PB-FeOOH electrode in 0.1 M PBS at −0.2 V using 80 μM of H<sub>2</sub>O<sub>2</sub> (initial and final additions), 400 µM of uric acid, 100 µM of ascorbic acid, and 40 µM of dopamine.</p> ">
Versions Notes

Abstract

:
In this study, an efficient H2O2 sensor was developed based on electrochemical Prussian blue (PB) synthesized from the acid suspension of δ-FeOOH and K3[Fe(CN)6] using cyclic voltammetry (CV) and anchored on carbon felt (CF), yielding an enhanced CF/PB-FeOOH electrode for sensing of H2O2 in pH-neutral solution. CF/PB-FeOOH electrode construction was proved by scanning electron microscopy (SEM), energy-dispersive X-ray spectroscopy (EDS), and X-ray diffraction (XRD), and electrochemical properties were verified by impedance electrochemical and CV. The synergy of δ-FeOOH and PB coupled to CF increases electrocatalytic activity toward H2O2, with the sensor showing a linear range of 1.2 to 300 μM and a limit of detection of 0.36 μM. Notably, the CF/PB-FeOOH electrode exhibited excellent selectivity for H2O2 detection in the presence of dopamine (DA), uric acid (UA), and ascorbic acid (AA). The calculated H2O2 recovery rates varied between 93% and 101% in fetal bovine serum diluted in PBS. This work underscores the potential of CF/PB-FeOOH electrodes in progressing electrochemical sensing technologies for various biological and environmental applications.

1. Introduction

Hydrogen peroxide is a potent oxidizing agent widely applied in the food, cosmetic, and textile industries [1]. Additionally, it is a vital molecule within the human body, implicated in immunological responses, inflammation, and cell signaling [2]. The scientific literature characterizes hydrogen peroxide as a reactive oxygen species, with a concentration exceeding 50 µM in biological fluids, signaling metabolic dysfunction to diseases like Alzheimer’s, Parkinson’s, and diabetes, among other health conditions [3,4]. In the environmental context, hydrogen peroxide concentration in aquatic ecosystems, including rivers, lakes, and oceans, indicates the health of fauna, flora, and overall water and air quality. In the food and beverages industry, hydrogen peroxide is a technological aid for diminishing the presence of bacteria [2]. Nonetheless, following processes such as pasteurization, sterilization, and packaging, the H2O2 content in the final product must not exceed 147 µM. Thus, it is crucial for health, safety, and the environment to ensure that the level of hydrogen peroxide remains within a safe range [2,5,6].
Considering the significance and wide-ranging applications of H2O2 in industrial and biological contexts, several techniques, including spectrophotometry, chemiluminescence, fluorescence, and chromatography, are employed to measure its concentration. Nonetheless, these techniques necessitate expert operators, rendering them unsuitable for point-of-care scenarios. Moreover, the extensive analysis duration and the equipment’s cumbersome nature preclude their use in remote locations [7].
Conversely, electrochemical methods present a promising alternative due to their numerous benefits, such as sensitivity, portability, rapid response time, cost-effectiveness, and an environmentally friendly analytical approach; however, they often lack specificity [8]. The electrochemical sensor is an accessible tool that non-specialist technicians can utilize for point-of-care detection and continuous monitoring. Since H2O2 is a reactive molecule, its detection is enhanced by its interaction with the working electrode’s surface. Modifying this surface with an electrocatalytic material can substantially increase the sensor’s specificity [9].
Electrochemical sensors can be constructed by immobilizing biological materials on working electrodes, such as enzymes or inorganic materials, including carbon nanomaterials, metal complexes, metal oxide hydroxides, etc. [10,11,12]. These materials act as catalysts for H2O2 reactions on surface electrodes. While enzymatic sensors exhibit selectivity and sensitivity, their catalytic activity is affected by pH, temperature, and the reaction medium, which impact sensor performance quality and limit these sensors’ shelf life [10]. In contrast, non-enzymatic sensors are more attractive because of their low cost and stability. Different groups of scientists have demonstrated various non-enzymatic sensors for H2O2 using organic molecules and other inorganic compounds. Ranni et al. demonstrated a sensor for H2O2 with gold nanoparticle-decorated copper cross-linked pectin on a glassy carbon electrode [13]. Yin et al. developed a biochar H2O2 sensor using crab shells as the carbon precursor through pyrolysis [14]. Wang et al. constructed a photoelectrochemical H2O2 sensor based on pillar [5] arene-functionalized Au nanoparticles and MWNTs hybrid BiOBr heterojunction [15]. An electrochemical sensor modified with cerium oxide nanocubes and carbon black as a highly active non-enzymatic H2O2 catalyst was proposed by Shen et al. [16]. Researchers have demonstrated that semiconductor δ-FeOOH, a material with good catalytic performance and practical application in photocatalysis [17,18], produces non-enzymatic sensors with excellent electrocatalytic activity and electronic conduction properties. For instance, Afonso and colleagues constructed an all-plastic disposable carbon electrochemical cell modified with δ-FeOOH and silver nanoparticles, which exhibited an electrocatalytic response for H2O2 with a limit of detection of 71 µM [19]. Further research demonstrated an improvement in the same system with the introduction of carbon black, achieving a limit of detection of 22 µM [20].
Another important inorganic compound used to modify electrodes for electrocatalytic purposes is Prussian blue (PB). PB is known as an artificial enzyme capable of catalyzing H2O2 reactions. Extensive research involving PB-based composite materials for electrochemical sensing of H2O2 has focused on synthesizing and depositing PB nanoparticles on various conductive surfaces to enhance their electrocatalytic and conductivity properties [21,22,23,24]. However, enhancing individual properties does not guarantee the material’s overall performance, necessitating better structural integration of the components.
Moreover, PB, as an electrocatalyst for H2O2 reaction, faces several limitations. The penetration of H2O2 and counter ions into PB structure can lead to volume changes and mechanical stress, resulting in reduced stability [25]. Side reactions caused by lower electrical conductivity and decomposition of PB structure in a medium with pH-neutral or basic are effects that reduce the stability of PB [26,27,28]. Therefore, it is essential to develop methods that improve the intimate contact between components of the sensing surface and enhance the physical and electrochemical stability.
Scholtz and colleagues demonstrated that Prussian blue (PB) strongly interacts with the goethite surface (α-FeOOH) due to the prevalence of positive charge groups in a medium with a pH below 9 [29]. This interaction facilitates PB binding to the surface’s negative charges and hexacyanoferrate anions, which provides remarkable stability for Prussian blue. Significantly, this property can be extended to other iron oxide hydroxides, making them effective traps for Prussian blue and its analogs. By integrating sensitive components for H2O2 detection with suitable conductive substrates, the resulting sensor’s sensitivity, selectivity, and practicality can be significantly enhanced.
Flexible electrodes, particularly carbon felt, stand out due to their adaptability to substrate shapes, robust mechanical properties, and superior electrical conductivity [30,31]. These characteristics are instrumental in elevating the efficacy of electrochemical sensors. Carbon felt is especially noteworthy among flexible electrodes for its extensive electrochemical surface area, chemical stability, ease of regeneration, swift charge and ion transport, and affordability [32]. While various electrochemical sensors employing carbon felt have been utilized to detect H2O2 and other analytes using diverse nanomaterials [23,33,34,35,36], the development of a non-enzymatic sensor that synergizes the properties of PB and δ-FeOOH remains an area not yet studied. δ-FeOOH is a relatively new material in sensor development, and its unique properties and potential applications are still being explored.
In this study, we have developed a novel non-enzymatic and accurate sensor for the electrochemical sensing of H2O2. PB was electrochemically synthesized and deposited on CF from a δ-FeOOH acid suspension, which also was adsorbed on CF, resulting in the formation of the CF/PB-FeOOH electrode. The synergistic effect of PB and δ-FeOOH enhances the electrode’s performance, as demonstrated by its excellent electrochemical reduction in H2O2, rapid response, high sensitivity, low limit of detection, and strong applicability in biological samples.

2. Materials and Methods

2.1. Reagents and Apparatus

The experiment used only chemicals of analytical grade. Ammonium iron (II) sulfate hexahydrate (NH4)2Fe(SO4)2·6H2O, uric acid, dopamine hydrochloride, ascorbic acid, and fetal bovine serum (FBS) were purchased from Sigma-Aldrich (St. Louis, MO, USA); potassium hexacyanoferrate (III) (K3[Fe(CN)6]), potassium hexacyanoferrate (II) trihydrate (K4[Fe(CN)6]), and iron chloride (III) (FeCl3) were purchased from A.C.S. scientific (Sumaré, SP, Brazil); potassium phosphate monobasic (KH2PO4) and potassium phosphate dibasic (K2HPO4) were obtained from Dinâmica Química (Indaiatuba, SP, Brazil); 37% v/v hydrogen chloride (HCl) was supplied by Labsynth (Diadema, SP, Brazil); sodium hydroxide (NaOH) was purchased from Anidrol (Diadema, SP, Brazil); 30% v/v hydrogen peroxide (H2O2) was purchased from Alphatec (Santo André, SP, Brazil); and potassium chloride was obtained from C.R.Q. (Diadema, SP, Brazil). Ultrapure water from a Millipore Direct Q® system (Billerica, MA, USA) was used in this research.
The electrochemical experiments were conducted on a potentiostat/galvanostat Autolab PGSTAT128N ( Utrecht, The Netherlands) connected to a computer running NOVA 2.0 software. The 50 mL one-compartment electrochemical cell was equipped with a working electrode made of carbon felt (CF) (Fuel Cell Store, TX, USA) cut into small ribbons with 2.5 cm2 areas, a platinum counter electrode, and Ag|AgCl immersed in 3 M KCl reference electrode. The morphology of the PB nanoparticles was evaluated by scanning electron microscopy using a Philips XL-30 FEG microscope equipped with an energy-dispersive X-ray spectrometer (EDS, Bruker, Singapore). The structural analysis of PB-FeOOH film was carried out by XRD measurements using an X-ray diffractometer with Rigaku Miniflex 600 diffractometer, which emitted Cu Kα radiation (λ = 1.54056 Å) at a scanning speed of 2° min−1 at 30 kV.

2.2. Synthesis of δ-FeOOH

The synthesis of δ-FeOOH followed the procedure reported by Pereira et al. [37]. δ-FeOOH was prepared by adding 200 mL of 2 M NaOH to 200 mL of 0.71 mM (NH4)2Fe(SO4)2·6H2O solution with stirring. After forming the green precipitate, 5 mL of H2O2 was added and magnetically stirred for 30 min. The reddish-brown solid formed indicated the attainment of the δ-FeOOH phase. The as-made material was washed with deionized water, centrifuged at 3600 rpm for 5 min, and dried under vacuum at room temperature.

2.3. Carbon Felt Electrode Preparation

We assessed three different approaches to preparing the CF surface for PB synthesis: (i) the CF was exposed to an air atmosphere in the oven, maintained at 100 °C for 2 h; (ii) the CF was immersed in 1 M H2SO4 solution for 2 h, followed by a thorough rinse with distilled water and subsequent drying in the oven at 100 °C for 2 h; (iii) we cleansed the CF with distilled water and left to dry at room temperature.

2.4. Synthesis of PB-FeOOH Film

PB-FeOOH film was formed by electrochemically synthesizing PB nanoparticles on CF using cyclic voltammetry at a scan rate of 50 mV·s−1 and a voltage range of –0.2 V to 0.7 V vs. Ag|AgCl. The synthesis was performed by placing CF in a precursor solution containing 5 mM of K3[Fe(CN)6], 22 mg of δ-FeOOH, and 0.1 M KCl as a supporting electrolyte, forming the PB-FeOOH film on CF. The quantity of δ-FeOOH in precursor solution was varied in 11, 22, and 44 mg of δ-FeOOH. The conditions for forming the PB-FeOOH film, such as the pH values of the precursor solution and the number of cyclic voltammetric scans, were also assessed. The effect of the pH value of the precursor solution from 2.0 to 6.0 was studied by adjusting the solution with 0.1 M HCl. Finally, CV scan numbers were evaluated using 15, 30, 50, 100, and 150 scans. For comparison, PB nanoparticles using a CF surface previously optimized were synthesized by applying 100 cyclic voltammetric in 2 mM FeCl3, 2 mM K3[Fe(CN)6], 0.1 M KCl, and 0.1 M HCl solution based on the modified method published by De Mattos [38].
The XRD analysis was performed using conducting fluorine-doped tin oxide (FTO)-coated glass where PB–FeOOH was deposited. Before deposition, the substrate was washed with acetone and deionized water in an ultrasonic bath for 15 min.

2.5. Electrochemical Measurement

CF/PB–FeOOH electrodes were electrochemically characterized in a 0.1 M KCl solution at pH 2.0 and a 1 mM equimolar mixture of K3[Fe(CN)6]/K4[Fe(CN)6 containing 0.1 M KCl as a supporting electrolyte by cyclic voltammetry. The solution used for amperometric measurement was 0.1 M PBS at pH 7.0, under stirring, with or without adding H2O2. All solutions were N2-saturated for 15 min. The electrochemical impedance spectroscopy was conducted in the hexacyanoferrate solution, applying an open-circuit potential with a 5 mV amplitude and a frequency range from 100 kHz to 0.01 Hz.

3. Results and Discussion

According to the previous literature, CF surface treatments using acid solutions or heating in an air atmosphere have been shown to improve the electrochemical activity of the CF surface [34,35]. Therefore, we evaluated three approaches to prepare the CF surface by studying the magnitude of redox peaks of PB nanoparticles on CF. Figure 1A shows typical CV curves of PB nanoparticle synthesis, which occurs during the potential scanning in 5 mM of K3[Fe(CN)6], 22 mg of δ-FeOOH, and 0.1 M KCl at pH 2.0, applying 30 cycles. We observed increasing redox peaks at 0.39 V and 0.12 V, along with a slight potential shift with the number of scans, suggesting the growth of stable PB nanoparticles [39] on the CF electrode. Furthermore, δ-FeOOH was also deposited on CF due to electrostatic interaction [29]. Therefore, the electrode was designated CF/PB-FeOOH. CV tests were conducted in a 0.1 M KCl solution pH 2.0 to evaluate the electrochemical properties of CF/PB-FeOOH electrodes (Figure 1B). The characterization was maintained in a solution with pH 2 to ensure the structure and properties of the PB. The CF/PB-FeOOH electrodes exhibit two pairs of sharp redox peaks related to Prussian white (PW)/PB near 0.3/0.2 V with a potential separation of 100 mV; this was observed for three different CF surfaces prepared, indicating a fast charge transfer in CF (Reaction (R1)). Among the approaches used to prepare the CF surface, a better signal was achieved by washing CF with distilled water and drying it at room temperature (curve c). Contrary to the literature, which claims that the electrochemical activity of CF electrodes improves with chemical and thermal treatments—treatments that increase the functional groups as hydroxyl and carbonyl—we did not observe this in our findings. The interaction of FeOOH with CF was more effective; therefore, we conducted the following experiments with CF prepared by method (iii).
Fe4III [FeII(CN)6]3 + 4K+ + 4e ⇌ K4Fe4II[FeII(CN)6]3
PB PW
The electrochemical synthesis of PB on electrodes involves several controllable factors to achieve PB deposited with good electrocatalytic activity, such as deposition time, concentration of reagents, and effect of solution pH [40]. Figure 2A shows the CV of CF/PB-FeOOH electrodes in 0.1 M KCl solution pH 2.0 constructed in precursor solution with pH ranging from 2 to 6. The CF/PB-FeOOH electrode constructed under pH 2.0 exhibited notably more significant two pairs of redox peaks than those constructed under other pH conditions, indicating the successful electrochemical formation of PB from δ-FeOOH in an acid solution. Subsequently, we assessed the effect of δ-FeOOH mass in the precursor solution. Figure 2B shows the electrochemical performance of the CF/PB-FeOOH electrode in 0.1 M KCl prepared with 11, 22, and 44 mg of FeOOH. The optimal amount of FeOOH was 22 mg, which was used in the subsequent experiments. To analyze the electrochemical behavior of the CF/PB-FeOOH electrode obtained with different deposition cycles, we obtained the CV curves of the CF/PB-FeOOH electrodes in 0.1 M KCl generated with 15, 30, 50, 100, and 150 cycles. The peak current enhanced as the number of scans increased until 100 cycles but decreased with 150 cycles, indicating that the electrochemical properties of PB deteriorated with a high number of scans (Figure 2C). Based on the best electrochemical performance of the optimized CF/PB-FeOOH electrode, we proceed with the upcoming experiments under these conditions.
The amount of PB deposited on the CF/PB-FeOOH electrode following 100 electrodeposition cycles was determined by CV collected in 0.1 M KCl using Equation (1).
Γ Γ = Q n F A
Q is the charge equivalent to the reduction peak at 0.16 V, n represents the number of electrons in the redox process (n = 1), F stands for the Faraday constant, A is the electroactive area (3.0 cm2), and Γ Γ is the amount of PB on the active surface. The quantity of PB deposited was calculated to be 4.4 × 10−9 mol/cm2.
The structural property of PB-FeOOH nanoparticles was evaluated by X-ray diffraction using FTO as substrate. Figure 3 shows the XRD pattern with only a characteristic strong peak (JCPDS#1-239) at 17.65° corresponding PB phase (100). This result suggests that the PB-FeOOH film has a predominantly amorphous structure [41], indicating no crystalline peaks other than the characteristic peak of PB.
Figure 4 shows the SEM images of the surface morphology of CF and CF/PB-FeOOH electrodes, respectively. The bare CF electrode (Figure 4A,B) has a fibrous and smooth surface, while the CF/PB-FeOOH electrode (Figure 4C,D) has a rough surface with nanometric cubic structures [40]. In addition, some irregular microstructures can be seen on the surface. The EDS analysis in Figure 4E,F shows C, O, Fe, N, K, and Cl elements, confirming the successful electrosynthesis of the PB on CF.
We use CV and EIS to assess the electrocatalytic activity of bare CF and CF/PB-FeOOH in a ferri-ferro cyanide solution. In Figure 5A, CV was carried out by cycling the potential between −0.2 and 0.7 V at a scan rate of 100 mV·s−1. The peak current in the redox probe at CF/PB-FeOOH (curve b) was higher than that at CF (curve a) due to the good electrochemical properties of the modified electrode. For the CF/PB-FeOOH electrode, the Ia/IC has an average of 0.86, and ΔEp is +0.21 V, showing a quasi-reversible electrochemical response. However, for the bare CF electrode, ΔEp is 0.12 V. This better value is probably due to the different thermodynamic conditions of the surface.
Furthermore, the effect of scan rate was studied. Figure S1 shows CVs obtained with different scan rates from 10 to 300 mV·s−1 for CF and CF/PB-FeOOH electrodes and the magnitudes of CV peak current plotted against the square root v1/2. The results exhibited a linear relationship for both anodic and cathodic current enhancement along with an increase in scan rate for the CF and CF/PB-FeOOH, indicating a diffusion-controlled process at the surface electrode. From these data, we experimentally determined the electroactive area for CF and CF/PB-FeOOH electrodes using the Randles–Ševčík equation. The electroactive areas were 3.0 ± 0.1 cm2 and 8.2 ± 0.4 cm2 for CF and CF/PB-FeOOH electrodes, respectively. CF modified with PB exhibited a large surface area and improved CF’s electroactive properties. These findings were aligned with EIS experiments. Figure 5B shows the Nyquist plot for CF (curve a) and CF/PB-FeOOH (curve b). The circuits used to obtain the EIS data were Rs(Rct[Cdl]) for CF and Rs(CPE[RctW]) for CF/PB-FeOOH, where Rs is solution resistance, Rct is electron transfer kinetics as charge transfer resistance, Cdl is double layer capacitance, CPE is constant phase element, and W is mass transfer element. The Rct for CF and CF/PB-FeOOH were 324.2 Ω and 1.28 Ω, respectively. From the EIS data, the value of the apparent heterogeneous electron transfer constant (kapp) was calculated using Equation (2).
k a p p = R T n 2 F 2 A R C t C
where R stands for the ideal gas constant (8.414 Joule/(mol·K)), T is the temperature absolute (298 K), n is the number of electrons transferred during reaction redox (n = 1), F is the constant of Faraday (96,485 C/mol), A stand for the active area of the electrode (cm2); Rct is electron transfer kinetics as charge transfer resistance (Ω); C is the concentration of the probe redox (mol/cm3). In the ferri-ferro solution, the Kapp calculated were 0.25 × 10−4 and 1 × 10−7 cm·s−1 for CF/PB-FeOOH and CF, respectively. The Kapp value obtained for CF/PB-FeOOH is 254 folds higher than that for CF; this improvement agrees with other Kapp values for modified electrodes reported in the literature [20,42].
Cyclic voltammograms were recorded for the CF/PB-FeOOH-modified electrode in a 0.1 M PBS solution to simulate the pH of the biology environment, with and without 1.0 mM of H2O2, at a scan rate of 50 mV/s. For comparison, we also evaluated the performance of PB deposited on CF using iron (III) chloride and ferricyanide as precursor solution (CF/PB). Figure 6A,B display CF/PB-FeOOH and FC/PB CVs, revealing redox peaks at 0.28 and 0.21 V and 0.46 and 0.43 V in PBS, respectively. The capacitive current of CF/PB was higher than that of CF/PB-FeOOH in PBS. This behavior is likely due to the larger CF/PB film thickness than CF/PB-FeOOH. PB thick films provide easier ion exchange between PB nanoparticles and the solution interface, improving the electroactivity of the electrode [43,44]. However, this property did not enhance PB’s electrochemical catalytic activity toward H2O2. After adding H2O2, CF/PB and CF/PB-FeOOH voltammetric profiles changed, especially for the CF/PB-FeOOH electrode. Evident alterations in current values were observed at 0.0 V, −0.1 V, and −0.2 V, amounting to 35%, 28%, and 29% for CF/PB, and 31%, 66%, and 85% for CF/PB-FeOOH, respectively. The difference between the current response in PBS and PBS with H2O2 was more pronounced for CF/PB-FeOOH, demonstrating greater sensitivity to H2O2. The H2O2 reduction by CF/PB-FeOOH can be explained by considering the synergistic effect of the Fenton-like reaction at the FeOOH (Reaction (2a–c)) surface and the PB electrocatalytic process according to Reaction (2d) [19,45]. Based on this comparison, we demonstrate the positive influence of FeOOH in the electrochemical properties of PB deposited on CF to the reduction in H2O2 and its potential application in electroanalytical. Therefore, we evaluated the analytical performance of the CF/PB-FeOOH electrode by amperometry at 0.0 V, −0.1 V, and −0.2 V, which were assessed for various H2O2 concentrations. The current response increased with H2O2 concentration. Figure S2 shows the calibration curve constructed taken from the three potentials. The linear regression equations were as follows: Y (μA) = 0.122 × C (μM) + 0.800 (R2 = 0.977); Y (μA) = 0.250 × C (μM) + 2.33 (R2 = 0.940); and Y (μA) = 0.164 × C (μM) + 0.150 (R2 = 0.998), and the limits of detection (LOD) calculated were 0.91 μM, 0.55 μM, and 0.36 μM (S/N = 3) for 0.0 V, −0.1 V, and −0.2 V, respectively. We observed that the current value at −0.2 V exhibited outstanding linearly and showed low LOD. The current CF/PB-FeOOH electrode values for H2O2 detection showed a linear range of 1.2 to 300 μM, and the limit of quantification (LOQ) was 1.19 μM (Figure 7A,B). Our results show that the CF/PB-FeOOH electrode exhibits a low limit of detection comparable to those previously reported. Additionally, it is constructed with inexpensive materials (see Table S1 [19,20,38,46,47,48,49,50]).
(a) FeOOH + H2O2 → (FeOOH)H2O2
(b) Fe(III)(surf)(H2O2) + e → Fe(II)(surf) + HOO• + H+
(c) Fe(II)(surf) + HOO• + H+ → Fe(III)(surf) + H2O + ½ O2
(d) H2O2 + 2e → 2OH
The reproducibility of the three independent CF/PB-FeOOH electrodes was tested by measuring the reduction current of 80 μM H2O2 under the same experimental conditions. Figure S3 shows no change in reduction current, and the relative standard deviation (RSD) is calculated to be 3.3%. This value typically represents the acceptable reproducibility of the CF/PB-FeOOH electrode. The selectivity of the CF/PB-FeOOH electrode was evaluated in PBS in the presence of electroactive species usually found in biological samples, such as dopamine (DA), uric acid (UA), and ascorbic acid (AA). According to the literature, (DA) is found in blood plasma in a concentration range of 0.24 to 23 µM [51], (UA) from 200 to 390 µM [52], and (AA) at 30 µM [53]. Therefore, we used 100 µM of ascorbic acid, 400 µM of uric acid, 40 µM of dopamine, and 80 µM of H2O2 in the selectivity assay. Figure 8 displays the amperometric response in PBS to adding H2O2 and electroactive interferents. A significant reduction current was observed upon introducing H2O2 at the initial and final stages of the experiment. In contrast, a weak current was detected upon adding AA despite the concentration being three times higher than usually found in blood plasma. Moreover, no discernible reduction current is detected for DA and UA. These results demonstrate that the CF/PB-FeOOH electrode possesses high selectivity.
The long-term stability of the CF/PB-FeOOH electrode was estimated by storing it at 25 °C in a room. After being stored and tested eighteen days later, the sensor’s current response maintained 93% of its initial response. This characteristic, combined with the steady electrochemical response of the CF/PB-FeOOH electrode, makes it useful in sensing H2O2 levels in biological samples.
Living cells can be induced to release H2O2 into a culture medium enriched with FBS, which supplies a range of nutrients essential for cell growth [54]. Different concentrations of H2O2 standard solution were spiked into PBS with 10% FBS to assess the practical application potential of the CF/PB-FeOOH electrode in biological samples. The calculated H2O2 recovery rates varied between 93% and 101%, as shown in Table 1. These results underscore the suitability of the CF/PB-FeOOH electrode for detecting H2O2 in biological samples.

4. Conclusions

We have developed a novel non-enzymatic electrochemical sensor utilizing PB and δ-FeOOH to determine H2O2. PB was successfully synthesized electrochemically from δ-FeOOH acid suspension and deposited on CF, yielding a CF/PB-FeOOH electrode. XRD analysis suggests that PB-FeOOH film has a predominantly amorphous structure, and SEM images revealed nanocubes deposited on CF. The synergic effect of PB and δ-FeOOH enhanced the electroactivity of the CF/PB-FeOOH electrode, showing sensitivity, selectivity, and reproducibility to determine H2O2. CF/PB-FeOOH achieved greater sensitivity to H2O2 than the CF/PB electrode, which was used for comparison. Moreover, the CF/PB-FeOOH electrode detected H2O2 in 10% fetal bovine serum. This work underscores the potential of CF/PB-FeOOH electrodes in progressing electrochemical sensing technologies for various biological and environmental applications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/c10030082/s1, Figure S1: CVs recorded at different scan rates (10–300 mV·s−1) and linear plot current versus square root of scan rate at CF (A, B) and CF/PB-FeOOH (C, D) in 1 mM K3Fe(CN)6, 1 mM K4Fe(CN)6 and 1 mM KCl solution at pH 2.0; Figure S2: Calibration curve plot of CF/PB-FeOOH at three different potentials with successive addition of H2O2 in 0.1 M PBS; Figure S3: Amperometric response of the three independent CF/PB-FeOOH electrodes at −0.2 V in 0.1 M PBS after adding 80 μM H2O2; Table S1: The detection of limit, applied potential and pH for different sensors for H2O2 sensing. Reference [55] is cited in the supplementary materials.

Author Contributions

K.S.N.: conceptualization, methodology, experimental work, data analysis, and writing—original draft; A.L.H.K.F.: methodology and experimental work; M.C.P.: methodology; F.G.E.N.: methodology; A.S.A.: conceptualization, supervision, methodology, data analysis, resources, and writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by FAPEMIG (APQ-00607-22).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

We thank FAPEMIG, CNPq, and FINEP/MCTI for their support. We also acknowledge the Institute of Science, Engineering, and Technology of the Federal University of Jequitinhonha and Mucuri Valleys for supporting this work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mattos, I.L.d.; Shiraishi, K.A.; Braz, A.D.; Fernandes, J.R. Peróxido de hidrogênio: Importância e determinação. Quim. Nova 2003, 26, 373–380. [Google Scholar] [CrossRef]
  2. Giaretta, J.E.; Duan, H.W.; Oveissi, F.; Farajikhah, S.; Dehghani, F.; Naficy, S. Flexible Sensors for Hydrogen Peroxide Detection: A Critical Review. ACS Appl. Mater. Interfaces 2022, 14, 20491–20505. [Google Scholar] [CrossRef]
  3. Cheung, E.C.; Vousden, K.H. The role of ROS in tumour development and progression. Nat. Rev. Cancer 2022, 22, 280–297. [Google Scholar] [CrossRef] [PubMed]
  4. Geraskevich, A.V.; Solomonenko, A.N.; Dorozhko, E.V.; Korotkova, E.I.; Barek, J. Electrochemical Sensors for the Detection of Reactive Oxygen Species in Biological Systems: A Critical Review. Crit. Rev. Anal. Chem. 2022, 54, 742–774. [Google Scholar] [CrossRef]
  5. Xing, L.J.; Zhang, W.G.; Fu, L.J.; Lorenzo, J.M.; Hao, Y.J. Fabrication and application of electrochemical sensor for analyzing hydrogen peroxide in food system and biological samples. Food Chem. 2022, 385, 132555. [Google Scholar] [CrossRef]
  6. Thatikayala, D.; Ponnamma, D.; Sadasivuni, K.K.; Cabibihan, J.-J.; Al-Ali, A.K.; Malik, R.A.; Min, B. Progress of Advanced Nanomaterials in the Non-Enzymatic Electrochemical Sensing of Glucose and H2O2. Biosensors 2020, 10, 151. [Google Scholar] [CrossRef]
  7. Thwala, L.N.; Ndlovu, S.C.; Mpofu, K.T.; Lugongolo, M.Y.; Mthunzi-Kufa, P. Nanotechnology-Based Diagnostics for Diseases Prevalent in Developing Countries: Current Advances in Point-of-Care Tests. Nanomaterials 2023, 13, 1247. [Google Scholar] [CrossRef]
  8. Baranwal, J.; Barse, B.; Gatto, G.; Broncova, G.; Kumar, A. Electrochemical Sensors and Their Applications: A Review. Chemosensors 2022, 10, 363. [Google Scholar] [CrossRef]
  9. Hasanzadeh, M.; Shadjou, N.; Guardia, M.d.l. Current advancement in electrochemical analysis of neurotransmitters in biological fluids. TrAC Trends Anal. Chem. 2017, 86, 107. [Google Scholar] [CrossRef]
  10. Chen, X.M.; Wu, G.H.; Cai, Z.X.; Oyama, M.; Chen, X. Advances in enzyme-free electrochemical sensors for hydrogen peroxide, glucose, and uric acid. Microchim. Acta 2014, 181, 689–705. [Google Scholar] [CrossRef]
  11. Zhao, C.L.; Zhang, H.F.; Zheng, J.B. Synthesis of silver decorated sea urchin-like FeOOH nanocomposites and its application for electrochemical detection of hydrogen peroxide. J. Mater. Sci. Mater. Electron. 2017, 28, 14369–14376. [Google Scholar] [CrossRef]
  12. Lim, H.C.; Cho, Y.J.; Han, D.H.; Kim, T.H. Enhancing electrocatalytic performance and Stability: A novel Prussian Blue-Graphene quantum dot nanoarchitecture for H2O2 reduction. Appl. Surf. Sci. 2024, 646, 158920. [Google Scholar] [CrossRef]
  13. Rani, K.K.; Liu, Y.X.; Devasenathipathy, R.; Yang, C.; Wang, S.F. Simple preparation of gold nanoparticle-decorated copper cross-linked pectin for the sensitive determination of hydrogen peroxide. Ionics 2019, 25, 309–317. [Google Scholar] [CrossRef]
  14. Yin, J.W.; Zhang, H.T.; Wang, Y.; Laurindo, M.B.J.; Zhao, J.F.; Hasebe, Y.; Zhang, Z.Q. Crab gill-derived nanorod-like carbons as bifunctional electrochemical sensors for detection of hydrogen peroxide and glucose. Ionics 2024, 30, 3541–3552. [Google Scholar] [CrossRef]
  15. Wang, J.; Zhou, Q.X.; Fan, C.; Guo, X.; Bei, J.L.; Chen, T.T.; Yang, J.; Yao, Y. Ultrasensitive and specific photoelectrochemical sensor for hydrogen peroxide detection based on pillar 5 arene-functionalized Au nanoparticles and MWNTs hybrid BiOBr heterojunction. Microchim. Acta 2024, 191, 266. [Google Scholar] [CrossRef] [PubMed]
  16. Shen, W.; Shi, J.C.; Wang, X.F.; Xu, P.C.; Li, X.X. A Non-Enzymatic Electrochemical Sensor Based on Cerium Oxide Nanocubes for the Rapid Detection of Hydrogen Peroxide Residues in Food Samples. IEEJ Trans. Electr. Electron. Eng. 2024, 19, 1573–1578. [Google Scholar] [CrossRef]
  17. Chen, Y.H.; Zhao, D.; Sun, T.Q.; Cai, C.R.; Dong, Y.M. The preparation of MoS2/8-FeOOH and degradation of RhB under visible light. J. Environ. Chem. Eng. 2023, 11, 110353. [Google Scholar] [CrossRef]
  18. Campos, P.T.B.; Vaiss, V.S.; Ramalho, T.C. The high potential of the pure and Nb-doped 6-FeOOH (001) surface in the adsorption and degradation of a neurotoxic agent. Surf. Sci. 2024, 746, 122491. [Google Scholar] [CrossRef]
  19. de Meira, F.H.A.; Resende, S.F.; Monteiro, D.S.; Pereira, M.C.; Mattoso, L.H.C.; Faria, R.C.; Afonso, A.S. A Non-enzymatic Ag/δ-FeOOH Sensor for Hydrogen Peroxide Determination using Disposable Carbon-based Electrochemical Cells. Electroanalysis 2020, 32, 2231–2236. [Google Scholar] [CrossRef]
  20. Melo, W.E.R.d.; Nantes, K.S.; Ferreira, A.L.H.K.; Pereira, M.C.; Mattoso, L.H.C.; Faria, R.C.; Afonso, A.S. A Disposable Carbon-Based Electrochemical Cell Modified with Carbon Black and Ag/δ-FeOOH for Non-Enzymatic H2O2 Electrochemical Sensing. Electrochem 2023, 4, 523–536. [Google Scholar] [CrossRef]
  21. Jin, E.; Lu, X.F.; Cui, L.L.; Chao, D.M.; Wang, C. Fabrication of graphene/prussian blue composite nanosheets and their electrocatalytic reduction of H2O2. Electrochim. Acta 2010, 55, 7230–7234. [Google Scholar] [CrossRef]
  22. Farah, A.M.; Shooto, N.D.; Thema, F.T.; Modise, J.S.; Dikio, E.D. Fabrication of Prussian Blue/Multi-Walled Carbon Nanotubes Modified Glassy Carbon Electrode for Electrochemical Detection of Hydrogen Peroxide. Int. J. Electrochem. Sci. 2012, 7, 4302–4313. [Google Scholar] [CrossRef]
  23. Han, L.J.; Tricard, S.; Fang, J.; Zhao, J.H.; Shen, W.G. Prussian blue @ platinum nanoparticles/graphite felt nanocomposite electrodes: Application as hydrogen peroxide sensor. Biosens. Bioelectron. 2013, 43, 120–124. [Google Scholar] [CrossRef] [PubMed]
  24. Husmann, S.; Nossol, E.; Zarbin, A.J.G. Carbon nanotube/Prussian blue paste electrodes: Characterization and study of key parameters for application as sensors for determination of low concentration of hydrogen peroxide. Sens. Actuator B Chem. 2014, 192, 782–790. [Google Scholar] [CrossRef]
  25. Ma, L.T.; Cui, H.L.; Chen, S.M.; Li, X.L.; Dong, B.B.; Zhi, C.Y. Accommodating diverse ions in Prussian blue analogs frameworks for rechargeable batteries: The electrochemical redox reactions. Nano Energy 2021, 81, 105632. [Google Scholar] [CrossRef]
  26. Pajerowski, D.M.; Watanabe, T.; Yamamoto, T.; Einaga, Y. Electronic conductivity in Berlin green and Prussian blue. Phys. Rev. B 2011, 83, 153202. [Google Scholar] [CrossRef]
  27. Yi, H.C.; Qin, R.Z.; Ding, S.X.; Wang, Y.T.; Li, S.N.; Zhao, Q.H.; Pan, F. Structure and Properties of Prussian Blue Analogues in Energy Storage and Conversion Applications. Adv. Funct. Mater. 2021, 31, 2006970. [Google Scholar] [CrossRef]
  28. Wang, Z.W.; Yang, H.K.; Gao, B.W.; Tong, Y.; Zhang, X.J.; Su, L. Stability improvement of Prussian blue in nonacidic solutions via an electrochemical post-treatment method and the shape evolution of Prussian blue from nanospheres to nanocubes. Analyst 2014, 139, 1127–1133. [Google Scholar] [CrossRef]
  29. Scholz, F.; Schwudke, D.; Stösser, R.; Bohácek, J. The interaction of Prussian blue and dissolved hexacyanoferrate ions with goethite (α-FeOOH) studied to assess the chemical stability and physical mobility of Prussian blue in soils. Ecotox. Environ. Saf. 2001, 49, 245–254. [Google Scholar] [CrossRef]
  30. Yang, Y.R.; Gao, W. Wearable and flexible electronics for continuous molecular monitoring. Chem. Soc. Rev. 2019, 48, 1465–1491. [Google Scholar] [CrossRef]
  31. Kim, J.; Campbell, A.S.; de Avila, B.E.F.; Wang, J. Wearable biosensors for healthcare monitoring. Nat. Biotechnol. 2019, 37, 389–406. [Google Scholar] [CrossRef] [PubMed]
  32. Moreira, F.C.; Boaventura, R.A.R.; Brillas, E.; Vilar, V.J.P. Electrochemical advanced oxidation processes: A review on their application to synthetic and real wastewaters. Appl. Catal. B Environ. 2017, 202, 217–261. [Google Scholar] [CrossRef]
  33. Guo, C.A.Y.; Chen, C.F.; Lu, J.Y.; Fu, D.; Yuan, C.Z.; Wu, X.L.; Hui, K.N.; Chen, J.R. Stable and recyclable Fe3C@CN catalyst supported on carbon felt for efficient activation of peroxymonosulfate. J. Colloid Interface Sci. 2021, 599, 219–226. [Google Scholar] [CrossRef]
  34. Sun, B.; Skyllas-Kazacos, M. Chemical modification of graphite electrode materials for vanadium redox flow battery application—Part II. Acid treatments. Electrochim. Acta 1992, 37, 2459–2465. [Google Scholar] [CrossRef]
  35. Sun, B.; Skyllas-Kazacos, M. Modification of graphite electrode materials for vanadium redox flow battery application—I. Thermal treatment. Electrochim. Acta 1992, 37, 1253–1260. [Google Scholar] [CrossRef]
  36. Zhao, Y.G.; Ying, M.; Fu, Y.B.; Chen, W. Improving Electrochemical Performance of Carbon Felt Anode by Modifying with Akaganeite in Marine Benthic Microbial Fuel Cells. Fuel Cells 2019, 19, 190–199. [Google Scholar] [CrossRef]
  37. Pereira, M.C.; Garcia, E.M.; da Silva, A.C.; Lorençon, E.; Ardisson, J.D.; Murad, E.; Fabris, J.D.; Matencio, T.; Ramalho, T.D.; Rocha, M.V.J. Nanostructured δ-FeOOH: A novel photocatalyst for water splitting. J. Mater. Chem. 2011, 21, 10280–10282. [Google Scholar] [CrossRef]
  38. de Mattos, I.L.; Gorton, L.; Ruzgas, T.; Karyakin, A.A. Sensor for hydrogen peroxide based on Prussian blue modified electrode: Improvement of the operational stability. Anal. Sci. 2000, 16, 795–798. [Google Scholar] [CrossRef]
  39. Vieira, T.A.; Souza, J.R.; Gimenes, D.T.; Munoz, R.A.A.; Nossol, E. Tuning electrochemical and morphological properties of Prussian blue/carbon nanotubes films through scan rate in cyclic voltammetry. Solid State Ion. 2019, 338, 5–11. [Google Scholar] [CrossRef]
  40. Matos-Peralta, Y.; Antuch, M. Review-Prussian Blue and Its Analogs as Appealing Materials for Electrochemical Sensing and Biosensing. J. Electrochem. Soc. 2019, 167, 037510. [Google Scholar] [CrossRef]
  41. Elshorbagy, M.H.; Ramadan, R.; Abdelhady, I. Preparation and characterization of spray-deposited efficient Prussian blue electrochromic thin film. Optik 2017, 129, 130–139. [Google Scholar] [CrossRef]
  42. Coros, M.; Varodi, C.; Pogacean, F.; Gal, E.; Pruneanu, S.M. Nitrogen-Doped Graphene: The Influence of Doping Level on the Charge-Transfer Resistance and Apparent Heterogeneous Electron Transfer Rate. Sensors 2020, 20, 1815. [Google Scholar] [CrossRef]
  43. Fu, Z.Y.; Wei, Y.X.; Liu, W.M.; Li, J.Y.; Li, J.M.; Ma, Y.B.; Zhang, X.F.; Yan, Y. Investigation of electrochromic device based on multi-step electrodeposited PB films. Ionics 2021, 27, 4419–4427. [Google Scholar] [CrossRef]
  44. Lu, S.Y.; Chen, Y.H.; Fang, X.F.; Feng, X. Hydrogen peroxide sensor based on electrodeposited Prussian blue film. J. Appl. Electrochem. 2017, 47, 1261–1271. [Google Scholar] [CrossRef]
  45. Karyakin, A.A.; Karyakina, E.E.; Gorton, L. On the mechanism of H2O2 reduction at Prussian blue modified electrodes. Electrochem. Commun. 1999, 1, 78–82. [Google Scholar] [CrossRef]
  46. Du, S.; Ren, Z.; Wu, J.; Xi, W.; Fu, H. Vertical α-FeOOH nanowires grown on the carbon fiber paper as a free-standing electrode for sensitive H2O2 detection. Nano Res. 2016, 9, 2260–2269. [Google Scholar] [CrossRef]
  47. Rattanopas, S.; Schulte, A.; Teanphonkrang, S. Prussian Blue/Carbon Nanotube Sensor Spread with Gelatin/Zein Glaze: A User-Friendly Modification for Stable Interference-Free H2O2 Amperometry. Anal. Chem. 2022, 94, 4919–4923. [Google Scholar] [CrossRef] [PubMed]
  48. Liu, X.L.; Zhang, X.J.; Zheng, J.B. One-pot fabrication of AuNPs-Prussian blue-Graphene oxide hybrid nanomaterials for non-enzymatic hydrogen peroxide electrochemical detection. Microchem. J. 2021, 160, 105595. [Google Scholar] [CrossRef]
  49. Uzunçar, S.; Özdogan, N.; Ak, M. Amperometric detection of glucose and H2O2 using peroxide selective electrode based on carboxymethylcellulose/polypyrrole and Prussian Blue nanocomposite. Mater. Today Commun. 2021, 26, 101839. [Google Scholar] [CrossRef]
  50. Li, N.; Zhou, H.Y.; Liu, Y.H.; Yu, X.J.; Cao, L.; Xu, Y.J.; Xi, L.X.; Zhao, G.; Ban, X.X. Prussian blue@Au nanoparticles/SiO2 cavity/ITO electrodes: Application as a hydrogen peroxide sensor. New J. Chem. 2024, 48, 1300–1306. [Google Scholar] [CrossRef]
  51. Phung, V.D.; Jung, W.S.; Nguyen, T.A.; Kim, J.H.; Lee, S.W. Reliable and quantitative SERS detection of dopamine levels in human blood plasma using a plasmonic Au/Ag nanocluster substrate. Nanoscale 2018, 10, 22493–22503. [Google Scholar] [CrossRef] [PubMed]
  52. Li, Q.; Yang, Z.; Lu, B.; Wen, J.; Ye, Z.; Chen, L.L.; He, M.; Tao, X.M.; Zhang, W.W.; Huang, Y.; et al. Serum uric acid level and its association with metabolic syndrome and carotid atherosclerosis in patients with type 2 diabetes. Cardiovasc. Diabetol. 2011, 10, 72. [Google Scholar] [CrossRef] [PubMed]
  53. Da Costa, L.A.; García-Bailo, B.; Borchers, C.H.; Badawi, A.; El-Sohemy, A. Association between the plasma proteome and serum ascorbic acid concentrations in humans. J. Nutr. Biochem. 2013, 24, 842–847. [Google Scholar] [CrossRef] [PubMed]
  54. van der Valk, J.J.S. Fetal bovine serum—A cell culture dilemma. Science 2022, 375, 143–144. [Google Scholar] [CrossRef]
  55. Jiang, T.; Zhan, D.P.; Chen, Y. Preparation of carbon nanotube arrays nanocomposites filled with Prussian blue and electrochemical sensing of hydrogen peroxide. Ferroelectrics 2021, 580, 42–54. [Google Scholar] [CrossRef]
Figure 1. (A) Potentiodynamic growth of PB nanoparticles on CF electrode in N2-saturated 0.1 M KCl solution at pH 2.0 with 5 mM K3Fe(CN)6 and 22 mg of δ-FeOOH. The potential window was set from −0.2 V to 0.7 V at a scan rate of 50 mV s−1, applying 30 scans. The arrows indicate scans increase. (B) CV of FC/PB-FeOOH electrode in N2-saturated 0.1 M KCl at pH 2.0 at a scan rate of 50 mV s−1. The CF was subjected to the following treatments: (a) i, (b) ii, and (c) iii.
Figure 1. (A) Potentiodynamic growth of PB nanoparticles on CF electrode in N2-saturated 0.1 M KCl solution at pH 2.0 with 5 mM K3Fe(CN)6 and 22 mg of δ-FeOOH. The potential window was set from −0.2 V to 0.7 V at a scan rate of 50 mV s−1, applying 30 scans. The arrows indicate scans increase. (B) CV of FC/PB-FeOOH electrode in N2-saturated 0.1 M KCl at pH 2.0 at a scan rate of 50 mV s−1. The CF was subjected to the following treatments: (a) i, (b) ii, and (c) iii.
Carbon 10 00082 g001
Figure 2. (A) Evaluation by the CV of PB synthesized in different pH levels of the precursor solution. (B) PB synthesized with (a) 44 mg, (b) 22 mg, or (c) 11 mg of δ-FeOOH. (C) Effect of the number of scans on the synthesis of PB: (a) background, (b) 15 scans, (c) 30 scans, (d) 50 scans, (e) 100 scans, and (f) 150 scans. N2-saturated 0.1 M KCl solution at pH 2.0 was the medium in which CVs were performed. The potential range was from −0.2 V to 0.7 V, and the scan rate was 50 mV s−1.
Figure 2. (A) Evaluation by the CV of PB synthesized in different pH levels of the precursor solution. (B) PB synthesized with (a) 44 mg, (b) 22 mg, or (c) 11 mg of δ-FeOOH. (C) Effect of the number of scans on the synthesis of PB: (a) background, (b) 15 scans, (c) 30 scans, (d) 50 scans, (e) 100 scans, and (f) 150 scans. N2-saturated 0.1 M KCl solution at pH 2.0 was the medium in which CVs were performed. The potential range was from −0.2 V to 0.7 V, and the scan rate was 50 mV s−1.
Carbon 10 00082 g002
Figure 3. The XRD pattern for the PB-FeOOH film was deposited on FTO.
Figure 3. The XRD pattern for the PB-FeOOH film was deposited on FTO.
Carbon 10 00082 g003
Figure 4. SEM micrographs of (A,B) CF electrode and (C,D) CF/PB-FeOOH electrode. (E) EDS spectrum of CF/PB-FeOOH electrode and (F) the weight % of EDS analysis.
Figure 4. SEM micrographs of (A,B) CF electrode and (C,D) CF/PB-FeOOH electrode. (E) EDS spectrum of CF/PB-FeOOH electrode and (F) the weight % of EDS analysis.
Carbon 10 00082 g004
Figure 5. (A) CV curves at 100 mV·s−1 and (B) EIS plots at open-circuit potential using N2-saturated 0.1 M KCl solution at pH 2.0 containing 1 mM K3Fe(CN)6 and 1 mM K4Fe(CN)6, (a) CF, and (b) CF/PB-FeOOH.
Figure 5. (A) CV curves at 100 mV·s−1 and (B) EIS plots at open-circuit potential using N2-saturated 0.1 M KCl solution at pH 2.0 containing 1 mM K3Fe(CN)6 and 1 mM K4Fe(CN)6, (a) CF, and (b) CF/PB-FeOOH.
Carbon 10 00082 g005
Figure 6. CV curves at a potential range from −0.2 V to 0.7 V at a scan rate of 100 mV·s−1 in 0.1 M of PBS without or with 1.0 mM of H2O2. (A) CF/PB-FeOOH electrode and (B) CF/PB electrode.
Figure 6. CV curves at a potential range from −0.2 V to 0.7 V at a scan rate of 100 mV·s−1 in 0.1 M of PBS without or with 1.0 mM of H2O2. (A) CF/PB-FeOOH electrode and (B) CF/PB electrode.
Carbon 10 00082 g006
Figure 7. (A) Amperometric response of the CF/PB–δ-FeOOH electrode in 0.1 M PBS at −0.2 V versus Ag|AgCl with successive additions of H2O2 under stirring. (B) Corresponding calibration curve. The error bars show the standard deviation for n = 3. The arrow indicates the increase in concentration of H2O2.
Figure 7. (A) Amperometric response of the CF/PB–δ-FeOOH electrode in 0.1 M PBS at −0.2 V versus Ag|AgCl with successive additions of H2O2 under stirring. (B) Corresponding calibration curve. The error bars show the standard deviation for n = 3. The arrow indicates the increase in concentration of H2O2.
Carbon 10 00082 g007
Figure 8. Amperometric response of CF/PB-FeOOH electrode in 0.1 M PBS at −0.2 V using 80 μM of H2O2 (initial and final additions), 400 µM of uric acid, 100 µM of ascorbic acid, and 40 µM of dopamine.
Figure 8. Amperometric response of CF/PB-FeOOH electrode in 0.1 M PBS at −0.2 V using 80 μM of H2O2 (initial and final additions), 400 µM of uric acid, 100 µM of ascorbic acid, and 40 µM of dopamine.
Carbon 10 00082 g008
Table 1. Spinking and recovery of H2O2 in PBS with 10% of FBS.
Table 1. Spinking and recovery of H2O2 in PBS with 10% of FBS.
SampleAdded (μM)Found (μM)Recovery (%)
154.6693
21514.3095
33030.28101
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

Nantes, K.S.; Ferreira, A.L.H.K.; Pereira, M.C.; Nogueira, F.G.E.; Afonso, A.S. A Novel Non-Enzymatic Efficient H2O2 Sensor Utilizing δ-FeOOH and Prussian Blue Anchoring on Carbon Felt Electrode. C 2024, 10, 82. https://doi.org/10.3390/c10030082

AMA Style

Nantes KS, Ferreira ALHK, Pereira MC, Nogueira FGE, Afonso AS. A Novel Non-Enzymatic Efficient H2O2 Sensor Utilizing δ-FeOOH and Prussian Blue Anchoring on Carbon Felt Electrode. C. 2024; 10(3):82. https://doi.org/10.3390/c10030082

Chicago/Turabian Style

Nantes, Karoline S., Ana L. H. K. Ferreira, Marcio C. Pereira, Francisco G. E. Nogueira, and André S. Afonso. 2024. "A Novel Non-Enzymatic Efficient H2O2 Sensor Utilizing δ-FeOOH and Prussian Blue Anchoring on Carbon Felt Electrode" C 10, no. 3: 82. https://doi.org/10.3390/c10030082

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