Next Article in Journal
Copper-Containing Films Obtained by the Simple Citrate Sol–Gel Route for NO2 Detection: Adsorption and Kinetic Study
Previous Article in Journal
Effect of Unmanned Aerial Vehicles on the Spatial Distribution of Analytes from Point Source
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Film Carbon Veil-Based Electrode Modified with Triton X-100 for Nitrite Determination

Departament of Physics and Chemistry, Research and Innovation Center of Sensor Technologies, Ural State University of Economics, 8 Marta St., 62, Yekaterinburg 620144, Russia
*
Author to whom correspondence should be addressed.
Submission received: 15 July 2020 / Revised: 21 August 2020 / Accepted: 26 August 2020 / Published: 28 August 2020

Abstract

:
A film carbon veil-based electrode (FCVE) modified with non-ionic surfactant Triton X-100 (TrX100) has been developed for nitrite determination. A new simple and producible technique of hot lamination (heat sealing) has been used for the FCVE manufacturing. The paper presents the findings of investigating the FCVE and the TrX100/FCVE by using voltammetry, chronoamperometry, and scanning electron microscopy. Modification of the electrode with TrX100 improves the hydrophilic property of its surface, which results in a larger electrode active area and higher sensitivity. Optimal conditions for nitrite determination with the use of the TrX100/FCVE have been identified. The linear range (LR) and the limit of detection (LOD) are 0.1–100 μM and 0.01 μM, respectively. The relative standard deviation (RSD) does not exceed 2.3%. High selectivity of the sensor ensures its successful application for the analysis of real samples (sausage products and natural water). The obtained results accord well with the results of the standard spectrophotometric method.

Graphical Abstract

1. Introduction

Nitrous acid salts (nitrites) are common contaminants of environmental objects (water, soil) and some food products. Sources of nitrite in the ecological environment are household wastes, animal wastes, atmospheric emissions from enterprises, as well as nitrogen-containing fertilizers used in agriculture to improve yields. Sodium nitrite/potassium nitrite (E250/E249) is one of the most commonly used food additives in food industry to preserve meat and fish products recognizable appearance. Nitrite accumulates in products of vegetable and animal origin and enters the human body with food.
Nitrite belongs to toxic substances and can cause inorganic contamination to human health. Excessive nitrite in the blood vascular system can reduce the blood capacity to transport oxygen and cause oxygen deprivation [1]. Besides, nitrite could interact with amides/amines to form harmful N-nitrosamine compounds, resulting in cancer [2,3]. This is the reason for controlling the concentration of nitrite in water and food products. Nitrite level in drinking water should not exceed 3 mg/L [4]. Sausage products may contain 50 mg of nitrite per kilogram [5].
Thus, exploiting effective means for nitrite detection in food products, biological fluids and water is essential for human health and environmental protection. Hence, different physical-chemical methods for nitrite detection have been proposed, such as chromatography [1,6], titration [7,8] spectrophotometry [9,10] and electrochemistry [11,12]. Along with obvious advantages they have some shortcomings. The chromatographic approach is labor intensive and complicated. The titration method lacks sensitivity. Spectrophotometry, though highly sensitive, requires the use of toxic and/or cancer-causing reagents [13]. Electrochemical methods have been favored over other methods of nitrite sensing due to their high sensitivity and selectivity, simplicity, low cost, and feasibility for on-site analysis.
In electrochemical methods, the analytical signal is dependent on the state of the surface of the working electrode. It is known that the transition from a macro to a micro and nanostructured surface is accompanied by an increase in sensitivity and selectivity of measurements, and a decrease in the detection limit of the substance [14]. In order to engineer the electrode surface and improve its analytical and operational characteristics, the electrode surface is modified [15]. Various nanomaterials with special catalytic, adsorption, and energy characteristics can be applied as modifiers of electrodes intended for nitrite determination. Among them are grapheme [16,17], carbon nanotubes [15], metal [18,19] and oxide [20] nanoparticles. Sometimes in order to increase the selectivity of NO2 determination, macrocyclic compounds [21], enzymes [21] and combinations of different nanomaterials [22,23] can be added to sensitive layers of sensors. The disadvantages of the existing NO2 sensors include the use of complex and lengthy operations for processing and modifying their surface. In some cases, this process can last from 60 min [24] to 24 h [25]. Sometimes high temperature (400–700 °C) pre-treatment of the electrode is needed [26,27]. In addition, the existing nitrite sensors do not always possess good analytical characteristics; for example, the linear range of some sensors does not exceed the same order of magnitude [16,18,28].
Electroanalysis may exploit sensors of different designs and shapes. Planar sensors seem to offer potential for usage in field conditions, combined with portable equipment. Most often, they are obtained applying the screen-printed technology and using various conductive pastes and composites [29,30]. Despite relative simplicity and scalability of the screen-printed technology, its application is limited by the need to manufacture a special stencil and the use of organic solvents needed to wash it for reuse. The hot lamination technology does not have these shortcomings. It is the process of bonding two flat surfaces, one of which has a special adhesive layer to glue to the other surface. The technology of hot lamination was successfully applied by the authors to manufacture a planar sensor for ascorbic acid determination [31]. A suitable material for use in hot lamination is carbon veil, which is chemically inert, has good heat resistance, high electrical conductivity, is strongly adhesive to film polymer materials and is not costly. To date, in the academic literature we could not find any reference to carbon veil electrodes/sensors for nitrite sensing manufactured by hot lamination. Currently the existing single sensors based on carbon felt (non-woven material, similar to carbon veil, but of higher density and thickness) for nitrite detection [24,25,28,32,33] are manufactured applying a complex technology that includes labor-intensive operations of fixing felt in the electrode body and ensuring an electrical contact [24,25,33]. In addition, for sensor fabrication carbon felt has to be pretreated by repeated washing with various solvents [25], ultrasonic processing [25], and electrochemical activation [24].
The aim of this work is to develop a planar electrochemical sensor based on carbon veil with the application of the hot lamination technology for highly sensitive and rapid detection of nitrite in various objects. For that purpose a number of tasks have to be solved: (1) to manufacture the carbon veil planar electrode using hot lamination, (2) to choose a modifier to improve the analytical characteristics of the sensor, (3) to study electrochemical behavior of nitrite on the modified electrode, (4) to choose the optimal conditions for nitrite determination, (5) to test the developed sensor on real objects.

2. Materials and Methods

2.1. Reagents and Materials

All chemicals were of chemical grade purity and used as received: standard of sodium nitrite (LenReaktiv JSC, St. Petersburg, Russia), sodium nitrite (JSC Vekton, St. Petersburg, Russia), orthophosphoric acid (NevaReaktiv Ltd., St. Petersburg, Russia), acetic acid (NevaReaktiv Ltd., St. Petersburg, Russia), boric acid (Samara pharmaceutical factory, Samara, Russia), hydrochloric acid (SigmaTec, Khimki, Russia), sodium borate (AO Reachim Ltd., Moscow, Russia), potassium ferricyanide (AO Reachim Ltd., Moscow, Russia), zinc acetate (AO Reachim Ltd., Moscow, Russia), sodium hydrate (JSC ChemReactivSnab, Ufa, Russia), sulphanilamide (Merck KGaA, Darmstadt, Germany), N-1 (1-naphthyl)-1,2-diaminoethandichloride (JSC LenReaktiv, St. Petersburg, Russia), TritonX-100 (TrX100) (Panreac Quimica SLU, Barcelona, Spain), cetyltrimethylammonium bromide (CTAB) (Sigma-Aldrich Co., St. Louis, MO, USA), sodiumdodecyl sulfate (SDS) (Sigma-Aldrich Chemie GmbH, Steinheim, Germany), citric acid (JSC ChemReactivSnab, Ufa, Russia), ascorbic acid (Sigma-Aldrich Co, St. Louis, MO, USA), glucose (NevaReaktiv Ltd., St. Petersburg, Russia), ferric chloride (III) (AO Reachim Ltd., Moscow, Russia), copper sulphate(AO Reachim Ltd., Moscow, Russia), cobalt acetate (AO Reachim Ltd., Moscow, Russia), potassium bromate (JSC LenReaktiv, St. Petersburg, Russia), Cementit universal (Merz+Benteli AG, Niederwangen, Switzerland), acetone (Ecos-1, Moscow, Russia). Deionized water with a resistivity of 18 MΩ cm was used as the solvent.
Carbon veil with surface density of 30 g/m2 and surface electrical resistivity of 8–10 Ω (M-Carbo Ltd., Minsk, Belarus), as well as polyethylene terephthalate sheets (303 × 216 × 0.125 mm) (Fellows, Vietnam) were used for electrode fabrication.

2.2. Instruments

A laminator LM-260iD (Rayson Electrical MFG., Ltd., Foshan, GuangDong, China) was used for manufacturing carbon veil electrodes. Scanning electron microscopy (SEM) measurements were performed on Auriga Cross Beam microscope (Carl Zeiss NTS, Oberkochen, Germany) at 2 kV accelerating voltage. Deionized water with a resistivity of 18 MΩ cm was obtained on an Akvalab-UVOI-MF-1812 installation (JSC RPC Mediana-filter, Moscow, Russia). A magnetic stirrer with controlled heating, RCT basic (IKA-Werke, Staufen, Germany), was used to prepared aqueous extracts of sausage products. Electrochemical (voltammetric and chronoamperometric) studies were conducted on IVA-5 analyzer (IVA Ltd., Yekaterinburg, Russia). Spectrophotometric measurements were performed on ECO-VIEW UV 1200 spectrophotometer (Shanghai Mapada Instruments Co., Ltd., Shanghai, China).

2.3. Procedures

2.3.1. Phytosynthesis of Gold Nanoparticles (Phyto-Au)

An extract from strawberry leaves was used as source of antioxidants (reducers) for synthesis of gold nanoparticles [34]. To 5 mL of boiling 1 mM HAuCl4 solution, 1 mL of freshly prepared extract from strawberry leaves was added with vigorous stirring (pH 11). Gold sol was synthesized for 5 min and then cooled to ambient temperature. The obtained nanoparticles were separated from the supernatant and resuspended in the initial volume of deionized water. The diameter of spherical phyto-Au was 14 nm [35].

2.3.2. Manufacturing of the Sensor

The sensor manufacturing procedure is shown in Figure 1. A piece of carbon veil was placed on a sheet of polyethylene terephthalate and moved between the laminator rolls heated to 140 °C. The obtained film coated with carbon veil was cut into 35 × 3 mm strips. A 293 × 206 mm section of carbon veil may result in 575 electrodes sized 35 × 3 mm. The middle part of the electrode separating the working and contact zones was covered with a mixture of cementite-acetone in a ratio of 1:5 by volume. The geometric area of the working zone of the electrode was 15 mm2 (5 × 3 mm). The resulting electrode was called a film carbon veil-based electrode (FCVE).
Modification of the FCVE working zone involved the process of drop-casting of surfactant solutions (TrX100, CTAB, SDS) and phytosynthesized sol of gold nanoparticles (phyto-Au) on the electrode. When applying 10 µL of 0.1 mM surfactant to the FCVE working area, the surfactant/FCVE was obtained. The modified phyto-Au/FCVE was obtained by applying 10 µL phyto-Au sol to the FCVE; phyto-Au/TrX100/FCVE was obtained by applying 10 µL of 0.1 mM TrX100 and 10 µL phyto-Au sol. After modification, the electrodes were dried under a regular incandescent lamp.

2.3.3. Electrochemical Measurements

Electrochemical studies were performed by cyclic and linear sweep voltammetry, as well as chronoamperometry in a three-electrode cell, including a silver–silver chloride reference electrode (Ag/AgCl/KCl, 3.5 M) (Gomel, Belarus), a carbon rod as an auxiliary electrode, and working electrodes FCVE and TrX100/FCVE. Working electrodes were washed with deionized water before application. Electrochemical measurements were carried out at (24 ± 1) °C.
Cyclic voltammograms of NO2 were recorded in the potential range from 0.3 to 1.5 V on FCVE and TrX100/FCVE at a potential scan rate 0.05 Vs−1.
Cyclic voltammograms 1.0 mM [Fe(CN)6]3–/4– were recorded in the potential range from −0.5 to +1.2 V on FCVE and TrX100/FCVE at a potential scan rate 0.05 Vs−1 in 0.1 M KCl.
Chronoamperometric measurements were carried out at potential +0.9 V in the solution containing 1.0 mM K4[Fe(CN)6] + 0.1 M KCl and at a potential +1.1 V in BRB pH 3 containing 0.02 mM NO2.
Linear sweep voltammograms of NO2 were recorded at anodic potential scanning in the range from 0.3 to 1.5 V. The potential scanning rate was varied within the range of 0.02–0.25 Vs−1.

2.3.4. Sample Preparation

Samples of sausage products from different manufacturers were bought in the nearest supermarket. The sample preparation was carried out in accordance with the prescribed procedure [36,37]: 20 g of sausage was ground on a grater to particles of 1–2 mm in size, placed in a heat-resistant glass and filled with 40 mL of distilled water. Then, the contents of the glass were heated in a water bath at 80 °C for 40 min while stirring. After cooling to 20 °C, the mixture was filtered through a blue ribbon filter. The filtrate was used for electrochemical analysis.
Samples of environmental water were tap water and water taken from wells. Water sample preparation was not performed.

2.4. Statistical Analysis and Data Treatment

All measurements were carried out three times, and the results were calculated for a confidence level of 0.95. The results are presented as X ± ΔX, where X is the average value and ΔX is the confidence interval. The recovery of NO2 was calculated according to IUPAC recommendations [38]. Limits of detection (LOD) and quantification (LOQ) were calculated as 3SD/b and 10SD/b, respectively, where SD is the standard deviation of the response and b is the slope of the calibration graphic. F-and t-tests were used to compare the results of NO2 determination in water samples and sausage extracts obtained on the developed TrX100/FCVE sensor and the reference spectrophotometric method.

3. Results

3.1. Electrochemical Behavior of NO2 on FCVE and Modified FCVE

The electrochemical behavior of nitrite on the FCVE and modified FCVE was studied in the Britton–Robinson buffer (BRB) background electrolyte. As can be seen in Figure 2, the cyclic voltammogram registered on the FCVE (curve 1) and TrX100/FCVE (curve 3) in the BRB solution in the absence of nitrite exhibits no anodic and cathodic signals, which indicates the electrochemical inertia of the carbon veil and TrX100. The introduction of 0.02 mM NO2 in BRB leads to the appearance of nitrite oxidation current at 1.18 V (Figure 2, curve 2 and 4). At the same time no signals are observed on the cathodic branch of the cyclic voltammogram, which indicates irreversible oxidation of nitrite on the FCVE and TrX100/FCVE. There is a three-fold increase in the oxidation current of 0.02 mM NO2 on the TrX100/FCVE as compared to the FCVE (Figure 2, curve 4). The reason for this effect is that the carbon veil is a poorly wettable material, and the modification with the surfactant increases its hydrophilicity, which results in a larger active surface area of the electrode (the surface area estimate is given in Section 3.3).
Gold nanoparticles are widely used as electrode modifying agents that improve analytical and sensor properties of the electrode. In such a way, modification of the electrode surface with gold nanoparticles (AuNPs) can lead to better heterogeneous charge transfer ability [25]; higher conductivity [39]; larger surface area [18]; exhibition of enhanced catalytic activity [40]; and, in general, to higher sensitivity to nitrite detection. Synthesis of gold nanoparticles has a considerable impact on properties of modified electrodes. The use of gold nanoparticles synthesized with plant extracts as a reducing and stabilizing agent ensures higher sensitivity and lower limit of ascorbic acid detection in comparison with gold nanoparticles, synthesized by the Turkevich method [31]. In this respect, the study of the electrochemical behavior of nitrite on the FCVE modified with phyto-synthesized gold nanoparticles is attracting widespread interest.
Figure 3 shows cyclic voltammograms registered on the phyto-Au/FCVE and the phyto-Au/TrX100/FCVE in BRB in the presence and in the absence of NO2. In the absence of nitrite in BRB, one oxidation current of gold nanoparticles is recorded at 1.26 V on the phyto-Au/FCVE and at 1.23 V on the phyto-Au/TrX100/FCVE; and one reduction current of Au(III) ions at 0.63 V and 0.50 V, respectively (Figure 3, curves 1 and 3, a dashed line). Addition of 0.02 mM NO2 in BRB does not result in additional signals on cyclic voltammograms (Figure 3, curves 2 and 4, a solid line), but causes a stronger and broader anodic signal registered in BRB in the absence of nitrite. It is evident that a broad anodic signal recorded in nitrite-containing solutions on the phyto-Au/FCVE and the phyto-Au/TrX100/FCVE is the consequence of the overlap of two processes: nitrite oxidation and gold nanoparticles oxidation. The comparison of Figure 2 and Figure 3 shows that the anodic signal recorded on the phyto-Au/FCVE and the phyto-Au/TrX100/FCVE in BRB containing NO2 is not of the catalytic nature. The absence of the catalytic effect in the processes of nitrite oxidation on the gold nanoparticles was reported by Brainina et al. [41]. Since modification of the FCVE and TrX100/FCVE with gold nanoparticles did not result in a strong and clear individual nitrite signal, these electrodes were not used in further work.

3.2. Choice of Surfactant for FCVE Modification

Considering the positive effect of modifying the FCVE with a surfactant (Figure 2), we tested the impact of three different surfactant types—cationic (CTAB), anionic (SDS), and nonionic (TrX100)—on the nitrite electrooxidation. The range 0.01–1.0 mM of the studied concentrations correlates with the critical micellar concentrations of each surfactant at 20–25 °C which was claimed by the producer: 0.92 mM for CTAB; 7–10 mM for SDS; and 0.2–0.9 mM for TrX100. Figure 4 illustrates the dependence of the nitrite oxidation peak current (Ip) on the logarithm of surfactant concentration used for the FCVE modification.
As can be identified from Figure 4, SDS concentration on the FCVE surface does not affect the nitrite oxidation peak current. It might be due to the fact that SDS generates negative charge on the FCVE surface and repels charges of nitrite ions. The dependence Ip = f (Csurfactant) for CTAB and TrX100 can be broken into two sections. When CTAB and TrX100 concentrations grow from 0.01 to 0.1 mM, the nitrite oxidation current increases. When CTAB and TrX100 concentrations grow from 0.1 to 1.0 mM, the nitrite oxidation current stays almost the same. An increase in the nitrite oxidation current in the first section of the dependence Ip = f (Csurfactant) can be explained by a better wettability property of the FCVE caused by CTAB and TrX100 adsorption and, as a consequence, larger active electrode area. In the second section the adsorption layer of the surfactant on the FCVE surface saturates, which results in stabilization of the nitrite oxidation current. It is worth mentioning that on approaching the critical micellar concentration, the spread in nitrite oxidation current values widens by each consequent measure, which is most evident for TrX100 (Figure 4, curve 3). Since the maximum nitrite oxidation current and minimum spread in its values are observed for 0.1 mM TrX100, our choice was made in favor of the TrX100/FCVE.

3.3. Characterization of the FCVE and the TrX100/FCVE

Figure 5 presents SEM-images of the FCVE and the TrX100/FCVE surfaces. The FCVE surface (Figure 5a) looks like a fiber consisting of thin, randomly intertwined threads that are held together by a binder. The diameter of the fiber is from 5 to 10 microns. The comparison of the SEM-images shows that the SEM-images of the TrX100/FCVE (Figure 5b) do not visually differ much from the FCVE (Figure 5a).
In order to select the right equation for calculating the effective surface area of the electrode, we studied electrochemical behavior of K4[Fe(CN)6] on the TrX100/FCVE and FCVE by applying the method of cyclic voltammetry. Figure 6 illustrates cyclic voltammograms on the FCVE and the TrX100/FCVE in 0.1 M KCl containing 1 mM K4[Fe(CN)6]. Compared to the FCVE, the increase in red-ox currents of [Fe(CN)6]3–/4– on the TrX100/FCVE is observed. The ratio of anodic and cathodic peak currents (Impa/Impc) is 2.35 for the FCVE and 1.34 for the TrX100/FCVE, which indicates that the rates of the anodic and cathodic processes on the TrX100/FCVE are close, in contrast to the FCVE, where the rate of the anodic process is higher than the cathodic one. The potential difference between the cathodic and anodic peaks in [Fe(CN)6]3−/4− is 1.00 V and 1.13 V for the TrX100/FCVE and the FCVE, respectively, which confirms the irreversible nature of the processes. Following the results of the experiment the Cottrell equation [42] was used to measure the effective surface area of the electrodes:
I = nFAC D π t
where I is current (A), n is the number of electrons (n=1), F is Faraday’s Constant (96 500 C mol−1), A is the electrochemical active area of the electrode (cm2), D is the diffusion coefficient (D = 7.6 × 10−6 cm2 s−1) [43], C is the bulk concentration of K4[Fe(CN)6] (mol cm−3) and t is time (s).
As is given in the Cottrell Equation (1), the dependence slope I = f (t−1\2) can be used to measure the active surface area of the electrode. To obtain the dependence I = f (t−1\2), chronoamperograms of K4[Fe(CN)6] oxidation were recorded at the controlled potential. Then t−1\2 was calculated for each current value on the chronoamperogram, and I = f (t−1\2) was plotted.
Figure 7 shows chronoamperograms of K4[Fe(CN)6] oxidation at a potential of 0.9 V on the FCVE and the TrX100/FCVE (Figure 7a), as well as the dependences I = f (t−1/2) (Figure 7b). As can be seen from Figure 7b, the slope of the linear relationship I = f (t−1/2) for the TrX100/FCVE is 2.5 times higher than for the FCVE. According to the performed calculations, the area of the FCVE appeared to be 17 mm2 and 40 mm2 for the TrX100/FCVE.

3.4. Effect of pH

The effect of the background electrolyte pH on voltammetric characteristics of nitrite oxidation on the TrX100/FCVE was measured in the range of pH from 2 to 7. As can be seen in Figure 8, the nitrite oxidation on the TrX100/FCVE proceeds with the least overvoltage (Ep) at pH 3. In the range of pH from 3 to 7 the equation of linear regression (2) is as follows:
Ep (V) = 0.0087pH + 1.1041, R2 = 0.9084
and shows that nitrite oxidation on the TrX100/FCVE within this pH range is almost proton-independent. Non effect of pH solution on the peak potential of nitrite oxidation was discussed by Afkhami et al. [44].
The effect of the peak current (Ip) of nitrite oxidation on solution pH (Figure 8) accords well with the literature on the subject [25,44,45,46]. Oxidation current has the highest value at pH 3. A sharp decline of nitrite oxidation current at pH < 3 may be caused by nitrite instability in highly acidic media due to conversion of NO2 to NO and NO3 [46,47], by Equation (3):
2H+ + 3NO2 → 2NO + NO3 + H2O
A gradual decline of nitrite signal at pH > 3 might result from the formation of oxide layers on the electrode surface in basic solutions, which hinders nitrite oxidation [44]. Taking into consideration the evidence supporting participation of one electron in the nitrite oxidation process in acidic media [25], the electrode reaction is proposed as follows Equation (4):
NO2 → NO2 + ē
As can be seen from this reaction, the nitrite ion loses one electron and generates nitrogen dioxide molecules that can accumulate on the electrode surface in due course. Brainina et al. [41] reported that nitrite ion electrooxidation on marco- and nanostructured electrode surfaces is not of catalytic nature, but is complicated by passivation of the electrode with an adsorbed product of an electrochemical reaction. The further studies were performed using BRB with pH 3.

3.5. Effect of Potential Scan Rate

The effect of the potential scan rate in the range from 20 to 250 mVs−1 on the peak current and the peak potential of nitrite electrooxidation are given in Figure 9. As the scan rate grows, the peak current of nitrite oxidation increases and the peak potential shifts to the anodic region (Figure 9), which indicates the irreversible electrode process [43]. Dependence of the peak potential on the natural logarithm of the potential scan rate is expressed by Equation (5):
Ep (V) = (1.57 ± 0.07) + (0.13 ± 0.01) lnν, R2 = 0.9756.
The process of nitrite electrooxidation on the TrX100/FCVE is diffusion-controlled, which is confirmed by the linear dependence of the nitrite oxidation peak current on the square root of the scan rate, which is expressed by Equation (6):
Ip (μA) = (45.58 ± 0.95)ν1/2 (Vs−1) − (0.85 ± 0.04), R2 = 0.9979.
This conclusion also confirms the dependence of the natural logarithm of the nitrite oxidation peak current on the natural logarithm of the potential scan rate, expressed by Equation (7):
lnIp = (3.81 ± 0.11) + (0.52 ± 0.02)lnν, R2 = 0.9969.
The slope of the dependence lnIp = f (lnν) is 0.52 and is close to the theoretical value of 0.5, which is characteristic of a diffusion-controlled process [48].
The nitrite diffusion coefficients for the FCVE and TrX100/FCVE were calculated using the Cottrell Equation (1) based on a chronoamperogram recorded at E = 1.1 V in BRB (pH 3) containing 0.02 mM NO2 and the corresponding dependence I = f (t−1/2) (Figure 10). The value of the nitrite diffusion coefficient for the TrX100/FCVE was found as 2.24 × 10−5 cm2 s−1, which accords with the same order of magnitude given at modified carbon paste electrode in other studies [11,49]. For the FCVE the value of D was found 1.52 × 10−7 cm2 s−1.

3.6. Analytic Characteristics of TrX100/FCVE

Figure 11 shows voltammograms of nitrite oxidation on the TrX100/FCVE at different concentrations of NO2 in BRB and the corresponding dependence of the peak current on nitrite concentration. It was found that the dependence Ip = f (CNO2) is linear in the range from 0.1 to 0.9 µM and from 0.9 to 100 µM and is expressed by Equations (8) and (9), respectively:
Ip = (3.46 ± 0.15)CNO2− − (0.13 ±0.01), R2 = 0.9963
Ip = (0.26 ± 0.01)CNO2− + (2.96 ± 0.09), R2 = 0.9957
The existence of two slopes of the dependence Ip = f (CNO2) may be linked with NO2 adsorption on the electrode surface with increase in concentration, which makes the diffusion of the analyte more difficult, as reported previously in the literature [16,20,50].
The determined sensitivity for these ranges of NO2 concentrations was 3.46 and 0.26 μAμM−1, respectively. LOD and LOQ were 0.01 and 0.03 μM, respectively. The relative standard deviation (RSD) of 0.2 mM NO2 response is 1.1%. The proposed sensor provides stability of the analytical signal for a month, over this period it decreases by 6–9%.
While measuring the TrX100/FCVE selectivity, we studied the impact of those various foreign species that could be present in real samples of sausage products and water, on the response from 5 μM of nitrites. Herewith, the concentration of foreign species that could change the nitrite response by ±5%, was considered non-interfering. The results of checking selectivity of the electrode toward nitrites are presented in Table 1. It can be seen that a 200-fold excess of glucose, Fe(III) ions, chloride and sulfate anions; a 100-fold excess of acetate ions; a 85-fold excess of Co(II); a 30-fold excess of Ni(II) and Cu(II); a 22-fold excess of BrO3; and a 10-fold excess of ascorbic and citric acids do not interfere with nitrite response. Thus, the developed sensor can tolerate a high concentration of interfering ions and, therefore, can be applied in the analysis of real samples.
Table 2 compares the analytical characteristics of the proposed carbon veil-based electrode and existing screen-printed sensors.
It can be seen from Table 2 that the detection limit, which is an important characteristic of the developed sensor, is not only comparable, but is even better than that of other modified electrodes. The other distinctive feature of the proposed carbon veil sensor is its inexpensiveness, as it is made of cheap materials and with simple productional technology (575 electrodes for 10 min). All these features allow the use of this sensor as disposable.

3.7. Determination of Nitrite in Real Samples

3.7.1. Analysis of Sausage Products

The electrodes were immersed into an electrochemical cell containing 9.0–9.5 mL of BRB (pH = 3), and a background voltammogram was recorded. An aliquot of the sausage extract (0.5–1.0 mL), obtained as described in Section 2.3.4, was added, thoroughly mixed, and a linear voltammogram was recorded from 0.3 to 1.5 V at a scan rate of 0.05 Vs−1. The measurements were taken three times. Then additives of a standard nitrite solution were added to the cell and the corresponding voltammograms were registered. The content of NO2 was found by applying the additional method. The impact of the matrix on nitrite determination was evaluated by the added-found method.
Table 3 presents the results of the NO2 concentration analysis of the sausage extracts on the proposed sensor. High recovery (93–95%) demonstrates that the proposed sensor could be well applied for NO2 sensing in sausage extracts.
Validation of the techniques for nitrite determination in sausage extracts with the TrX100/FCVE was performed by comparing the obtained results with the findings of the standard UV-spectroscopic method [55]. The latter is based on red-coloring of a solution containing nitrite by adding 4-aminobenzenesulfamide and N-1(1-naphthyl)-1,2-diaminoethandihydrochloride and the consequent photometric measurement at a wavelength of 538 nm. This comparison is illustrated in Table 4. The RSD values confirm high precision of the data obtained. F- and t-tests data are less than critical values at p = 0.95 confirming the absence of systematic errors in determinations.

3.7.2. Analysis of Water Samples

The determination of initial nitrite concentrations in water was performed in the same way as described in Section 3.7.1. A water sample (2.5 mL) was added with nitrite as nitrite-ions were not detected in natural water.
The applicability of the developed TrX100/FCVE was tested on the spiked natural water (Table 5). The accuracy of nitrite sensing was tested using the added-found method. The relative standard deviation in the range of 1.1–2.3 and recovery of 98–103% confirm the absence of the random errors and high accuracy of nitrite determination.
Validation of the techniques for nitrite determination in water samples with the TrX100/FCVE was performed by comparing the obtained results with the findings of the standard UV-spectroscopic method [56]. The latter is based on the interaction of nitrite with 4-aminobenzenesulfamide in the presence of ortho-phosphoric acid (pH 1.9) and the consequent formation of a colored compound detected at a wavelength of 540 nm. This comparison is illustrated in Table 6. As can be observed in Table 6, the RSD value for the TrX100/FCVE does not exceed 2.3%. The values obtained by both techniques concur quite well. F- and t-tests data are less than critical values at P = 0.95 confirming the absence of systematic errors in determinations.

4. Conclusions

A new sensor based on a carbon veil, modified with TrX100, for determination of nitrite is described. A simple and producible technique of hot lamination (heat sealing) ensures fast and mass production of disposable carbon veil electrodes at low cost. Miniature size of the sensor makes it highly promising for application in portable equipment for field conditions. The developed sensor is characterized by a wide linear range, a low detection limit, and a good reproducibility of measurements. High selectivity of the sensor ensures its successful application for the analysis of sausage products and natural water.

Author Contributions

Discussion and interpretation of the results, preparation of the final text N.Y.S.; carrying out experimental work and discussion of the results E.I.K.; writing—original manuscript M.A.B.; editing of the text A.V.T.; manufacturing sensor’s transducer S.V.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Genualdi, S.; Matos, M.; Mangrum, J.; de Jager, L.S. Investigation into the concentration and sources of nitrates and nitrites in milk and plant-based powders. J. Agric. Food Chem. 2020, 68, 1725–1730. [Google Scholar] [CrossRef]
  2. Song, P.; Wu, L.; Guan, W. Dietary nitrates, nitrites, and nitrosamines intake and the risk of 400 gastric cancer: A meta-analysis. Nutrients 2015, 7, 9872–9895. [Google Scholar] [CrossRef]
  3. Barry, K.H.; Jones, R.R.; Cantor, K.P. Ingested nitrate and nitrite and bladder cancer in northern new England. Epidemiology 2020, 31, 136–144. [Google Scholar] [CrossRef]
  4. WHO. Nitrate and nitrite in drinking-water. In Background Document for Preparation of WHO Guidelines for Drinking-Water Quality; World Health Organization: Geneva, Switzerland, 2011. [Google Scholar]
  5. State Standard 33673-2015. Cooked sausages. In General Specifications; Izdatelstvo Standartinform: Moscow, Russian, 2016. [Google Scholar]
  6. Murray, E.; Roche, P.; Briet, M.; Moore, B.; Morrin, A.; Diamond, D.; Paull, B. Fully automated, low-cost ion chromatography system for in-situ analysis of nitrite and nitrate in natural waters. Talanta 2020, 216, 120955. [Google Scholar] [CrossRef]
  7. Lenghartova, K.; Lauko, L.; Cacho, F.; Beinrohr, E. Determination of nitrites in water by in-electrode coulometric titration in reticulated vitreous carbon electrode. Acta Chim. Slov. 2015, 62, 152–158. [Google Scholar] [CrossRef] [Green Version]
  8. He, Z.K.; Fuhrmann, B.; Spohn, U. Precise and sensitive determination of nitrite by coulometric back titration under flow conditions. Fresenius J. Anal. Chem. 2000, 367, 264–269. [Google Scholar] [CrossRef]
  9. Li, Y.-S.; Zhao, C.-L.; Li, B.-L.; Gao, X.-F. Evaluating nitrite content changes in some Chinese home cooking a newely-developed CDs diazotization spectrophotometry. Food Chem. 2020, 330, 127151. [Google Scholar] [CrossRef]
  10. Abou-Melha, K.S. Analytical chemistry optical chemosensor for spectrophotometric determination of nitrite in wastewater. Chem. Select. 2020, 5, 6216–6223. [Google Scholar] [CrossRef]
  11. Nithyayini, K.N.; Harish, M.N.K.; Nagashree, K.L. Electrochemical detection of nitrite at NiFe2O4 nanoparticles synthesized by solvent deficient method. Electrochim. Acta 2019, 317, 701–710. [Google Scholar] [CrossRef]
  12. Peng, Z.W.; Yuan, D.; Jiang, Z.W.; Li, Y.F. Novel metal-organic gels of bis(benzimidazole)-based ligands with copper(II) for electrochemical selectively sensing of nitrite. Electrochim. Acta 2017, 238, 1–8. [Google Scholar] [CrossRef]
  13. Yenil, N.; Yemiş, F. Nitrite in nature: Determination with polymeric materials. Pak. J. Anal. Environ. Chem. 2018, 19, 104–114. [Google Scholar] [CrossRef]
  14. Stozhko, N.Y.; Malakhova, N.A.; Byzov, I.V.; Brainina, K.Z. Electrodes in stripping voltammetry: From a macro- to a micro- and nano-structured surface. J. Anal. Chem. 2009, 64, 1148–1157. [Google Scholar] [CrossRef]
  15. Brainina, K.; Stozhko, N.; Bukharinova, M.; Vikulova, E. Nanomaterials: Electrochemical properties and application in sensors. Phys. Sci. Rev. 2019, 3, 8050. [Google Scholar]
  16. Su, C.-H.; Sun, C.-L.; Liao, Y.-C. Printed combinatorial sensors for simultaneous detection of ascorbic acid, uric acid, dopamine, and nitrite. ACS Omega 2017, 2, 4245–4252. [Google Scholar] [CrossRef] [Green Version]
  17. Jian, J.-M.; Fu, L.; Ji, J.; Lin, L.; Guo, X.; Ren, T.-L. Electrochemically reduced graphene oxide/gold nanoparticles composite modified screen-printed carbon electrode for effective electrocatalytic analysis of nitrite in foods. Sens. Actuators B 2018, 262, 125–136. [Google Scholar] [CrossRef]
  18. Talbi, M.; Al-Hamry, A.; Teixeira, P.R.; Bouhamed, A.; Azzouzi, S.; Paterno, L.G.; Kanoun, O. Graphite screen printed electrodes functionalized with AuNPs-PEI for nitrite detection. In Proceedings of the 16th International Multi-Conference on Systems, Signals & Devices (SSD), Istanbul, Turkey, 21–24 March 2019; pp. 607–610. [Google Scholar]
  19. Lo, N.-C.; Sun, I.-W.; Chen, P.-Y. CuAg nanoparticles formed in situ on electrochemically pre-anodized screen-printed carbon electrodes for the detection of nitrate and nitrite anions. J. Chin. Chem. Soc. 2018, 65, 928–988. [Google Scholar] [CrossRef]
  20. Jaiswal, N.; Tiwari, I.; Foster, C.W.; Banks, C.E. Highly sensitive amperometric sensing of nitrite utilizing bulk-modified MnO2 decorated graphene oxide nanocomposite screen-printed electrodes. Electrochim. Acta 2017, 227, 255–266. [Google Scholar] [CrossRef]
  21. Palanisamy, S.; Thirumalraj, B.; Chen, S.-M. A novel amperometric nitrite sensor based on screen printed carbon electrode modified with graphite/β-cyclodextrin composite. J. Electroanal. Chem. 2016, 760, 97–104. [Google Scholar] [CrossRef]
  22. Promsuwan, K.; Thavarungkul, P.; Kanatharana, P.; Limbut, W. Flow injection amperometric nitrite sensor based on silver microcubics-poly (acrylic acid)/poly (vinyl alcohol) modified screen printed carbon electrode. Electrochim. Acta 2017, 232, 357–369. [Google Scholar] [CrossRef]
  23. Zhe, T.; Sun, X.; Wang, Q.; Liu, Y.; Li, R.; Li, F.; Wang, L. A screen printed carbon electrode modified with a lamellar nanocomposite containing dendritic silver nanostructures, reduced graphene oxide, and β-cyclodextrin for voltammetric sensing of nitrite. Microchim. Acta 2019, 186, 319. [Google Scholar] [CrossRef]
  24. Kuntolaksono, S.; Matsuura, H. Coulometric analysis of nitrite using electrochemically activated carbon felt electrode. Sens. Mater. 2019, 31, 1215–1224. [Google Scholar] [CrossRef]
  25. Wei, W.; Wu, S.-G. Study of electrooxidation behavior of nitrite on gold nanoparticles/graphitizing carbon felt electrode and its analytical application. Chin. J. Anal. Chem. 2019, 47, e19014–e19020. [Google Scholar] [CrossRef]
  26. Zhu, W.; Zhang, Y.; Gong, J.; Ma, Y.; Sun, J.; Li, T.; Wang, J. Surface engineering of carbon fiber paper toward exceptionally high-performance and stable electrochemical nitrite sensing. ACS Sens. 2019, 4, 2980–2987. [Google Scholar] [CrossRef] [PubMed]
  27. Muthumariappan, A.; Govindasamy, M.; Chen, S.-M.; Sakthivel, K.; Mani, V. Screen-printed electrode modified with a composite prepared from graphene oxide nanosheets and Mn3O4microcubes for ultrasensitive determination of nitrite. Microchim. Acta 2017, 184, 3625–3634. [Google Scholar] [CrossRef]
  28. Wang, L.; Tricard, S.; Cao, L.; Liang, Y.; Zhao, J.; Fang, J.; Shen, W. Prussian blue/1-butyl-3-methylimidazolium tetrafluoroborate–graphite felt electrodes for efficient electrocatalytic determination of nitrite. Sens. Actuators B 2015, 214, 70–75. [Google Scholar] [CrossRef]
  29. Kunpatee, K.; Traipop, S.; Chailapakul, O.; Chuanuwatanakul, S. Simultaneous determination of ascorbic acid, dopamine and uric acid using grapheme quantum dots/ionic liquid modified screen-printed carbon electrode. Sens. Actuators B 2020, 314, 128059. [Google Scholar] [CrossRef]
  30. Pérez-Fernández, B.; Costa-García, A.; De La Escosura-Muñiz, A. Electrochemical (bio)sensors for pesticides detection using screen-printed electrodes. Biosensors 2020, 10, 32. [Google Scholar]
  31. Brainina, K.; Bukharinova, M.A.; Stozhko, N.Y.; Sokolkov, S.V.; Tarasov, A.V.; Vidrevich, M.B. Electrochemical sensor based on a carbon veil modified by phytosynthesized gold nanoparticles for determination of ascorbic acid. Sensors 2020, 20, 1800. [Google Scholar] [CrossRef] [Green Version]
  32. Zhang, Y.; Zhu, W.; Wang, Y.; Ma, Y.; Sun, J.; Li, T.; Ji, Y. High-performance electrochemical nitrite sensing enabled by the commercial carbon fiber cloth. Inorg. Chem. Front. 2019, 6, 1501–1506. [Google Scholar] [CrossRef]
  33. Yatsunami, T.; Takase, S.; Shimizu, Y. Amperometric nitrite-ion sensor based on electrodeposited Sm-based perovskite-type oxide thick-film electrode. Sens. Mater. 2016, 28, 777–784. [Google Scholar]
  34. Brainina, K.; Stozhko, N.; Bukharinova, M.; Khamzina, E.; Vidrevich, M. Potentiometric method of plant microsuspensions antioxidant activity determination. Food. Chem. 2019, 278, 653–658. [Google Scholar] [CrossRef] [PubMed]
  35. Stozhko, N.Y.; Bukharinova, M.A.; Khamzina, E.I.; Tarasov, A.V.; Vidrevich, M.B.; Brainina, K.Z. The effect of the antioxidant activity of plant extracts on the properties of gold nanoparticles. Nanomaterials 2019, 9, 1655. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Balasubramanian, P.; Settu, R.; Chen, S.-M.; Chen, T.-W.; Sharmila, G. A new electrochemical sensor for highly sensitive and selective detection of nitrite in food samples based on sonochemical synthesized Calcium Ferrite (CaFe2O4) clusters modified screen printed carbon electrode. J. Colloid Interface Sci. 2018, 524, 417–426. [Google Scholar] [CrossRef] [PubMed]
  37. Yildiz, G.; Oztekin, N.; Orbay, A.; Senkal, F. Voltammetric determination of nitrite in meat products using polyvinylimidazole modified carbone paste electrode. Food Chem. 2014, 152, 245–250. [Google Scholar] [CrossRef] [PubMed]
  38. Burns, D.T.; Danzer, K.; Townshend, A. Use of the terms “recovery” and “apparent recovery” in analytical procedures (IUPAC Recommendations 2002). Pure Appl. Chem. 2003, 74, 2201–2205. [Google Scholar] [CrossRef]
  39. Chen, H.; Yang, T.; Liu, F.; Li, W. Electrodeposition of gold nanoparticles on Cu-based metal-organic framework for the electrochemical detection of nitrite. Sens. Actuators B 2018, 286, 401–407. [Google Scholar] [CrossRef]
  40. Pan, F.; Chen, D.; Zhuang, X.; Wu, X.; Luan, F.; Zhang, S.; Li, X. Fabrication of gold nanoparticles/l -cysteine functionalized graphene oxide nanocomposites and application for nitrite detection. J. Alloys Compd. 2018, 744, 51–56. [Google Scholar] [CrossRef]
  41. Brainina, K.Z.; Galperin, L.G.; Bukharinova, M.A.; Stozhko, N.Y. Mathematical modeling and experimental study of electrode processes. J. Solid State Electrochem. 2014, 19, 599–606. [Google Scholar] [CrossRef]
  42. Fotouhi, L.; Fatollahzadeh, M.; Heravi, M.M. Electrochemical behavior and voltammetric determination of sulfaguanidine at a glassy carbon electrode modified with a multi-walled carbon nanotube. Int. J. Electrochem. Sci. 2012, 7, 3919–3928. [Google Scholar]
  43. Bard, A.J.; Faulkner, L.R. Electrochemical Methods: Fundamentals and Applications, 2nd ed.; John Wiley & Sons Inc: Hoboken, NJ, USA, 2001. [Google Scholar]
  44. Afkhami, A.; Soltani-Felehgari, F.; Madrakian, T.; Ghaedi, H. Surface decoration of multi-walled carbone nanotubes modified carbone paste electrode with gold nanoparticles for electrooxidation and sensitive determination of nitrite. Biosens. Bioelectron. 2014, 51, 379–385. [Google Scholar] [CrossRef]
  45. Huang, X.; Li, Y.; Chen, Y.; Wang, L. Electrochemical determination of nitrite and iodate by use of gold nanoparticles/poly(3-methylthiophene) composites coated glassy carbon electrode. Sens. Actuators B 2008, 134, 780–786. [Google Scholar] [CrossRef]
  46. Pham, X.-H.; Li, C.A.; Han, K.N.; Huynh-Nguyen, B.-C.; Le, T.-H.; Ko, E.; Kim, J.H.; Seong, G.H. Electrochemical detection of nitrite using urchin-like palladium nanostructures on carbon nanotube thin film electrodes. Sens. Actuators B 2014, 193, 815–822. [Google Scholar] [CrossRef]
  47. Fu, L.; Yu, S.; Thompson, L.; Yu, A. Development of a novel nitrite electrochemical sensor by stepwise in situ formation of palladium and reduced graphene oxide nanocomposites. RSC Adv. 2015, 5, 40111–40116. [Google Scholar] [CrossRef]
  48. Gowda, J.I.; Nandibewoor, S.T. Electrochemical behavior of paclitaxel and its determination at glassy carbon electrode. Asian J. Pharm. Sci. 2014, 9, 42–49. [Google Scholar] [CrossRef] [Green Version]
  49. Heli, H.; Eskandari, I.; Sattarahmady, N.; Moosavi-Movahedi, A.A. Cobalt nanoflowers: Synthesis, characterization and derivatization to cobalt hexacyanoferrate—Electrocatalytic oxidation and determination of sulfite and nitrite. Electrochim. Acta 2012, 77, 294–301. [Google Scholar] [CrossRef]
  50. Zhang, S.; Li, B.Q.; Zheng, J.B. An electrochemical sensor for the sensitive determination of nitrites based on Pt-PANI-graphenenanocomposites. Anal. Methods 2015, 7, 8366–8372. [Google Scholar] [CrossRef]
  51. Balasubramanian, P.; Velmurugan, M.; Chen, S.-M.; Chen, T.-W.; Ye, Y.-T. A single-step electrochemical preparation of cadmium sulfide anchored ERGO/β-CD modified screen-printed carbon electrode for sensitive and selective detection of nitrite. J. Electrochem. Soc. 2019, 166, B690–B696. [Google Scholar] [CrossRef]
  52. Caetano, L.P.; Lima, A.P.; Tormin, T.F.; Richter, E.M.; Espindola, F.S.; Botelho, F.V.; Munoz, R.A.A. Carbon-nanotube modified screen-printed electrode for the simultaneous determination of nitrite and uric acid in biological fluids using batch-injection amperometric detection. Electroanalysis 2018, 30, 1862–1871. [Google Scholar] [CrossRef]
  53. Koyun, O.; Sahin, Y. Voltammetric determination of nitrite with gold nanoparticles/poly(methylene blue)-modified pencil graphite electrode: Application in food and water samples. Ionics 2018, 24, 3187–3197. [Google Scholar] [CrossRef]
  54. Monteiro, T.; Rodrigues, P.R.; Gonçalves, A.L.; Moura, J.J.G.; Jubete, E.; Añorga, L.; Almeida, M.G. Construction of effective disposable biosensors for point of care testing of nitrite. Talanta 2015, 142, 246–251. [Google Scholar] [CrossRef] [Green Version]
  55. State Standard 29299-92. Meat and meat products. In Determination of Nitrite Content; Izdatelstvo Standartov: Moscow, Russia, 2003. [Google Scholar]
  56. State Standard 33045-2014. Water. In Methods for Determination of Nitrogen-Containing Matters; Izdatelstvo Standartinform: Moscow, Russia, 2019. [Google Scholar]
Figure 1. Manufacturing of the film carbon veil-based electrode (FCVE).
Figure 1. Manufacturing of the film carbon veil-based electrode (FCVE).
Chemosensors 08 00078 g001
Figure 2. Cyclic voltammograms on the FCVE (1, 2) and TrX100/FCVE (3, 4), in Britton–Robinson buffer (BRB) (pH 3) in the absence (1, 3) and in the presence of 0.02 mM NO2 (2, 4). Potential scan rate: 0.05 Vs−1.
Figure 2. Cyclic voltammograms on the FCVE (1, 2) and TrX100/FCVE (3, 4), in Britton–Robinson buffer (BRB) (pH 3) in the absence (1, 3) and in the presence of 0.02 mM NO2 (2, 4). Potential scan rate: 0.05 Vs−1.
Chemosensors 08 00078 g002
Figure 3. Cyclic voltammograms on the phytosynthesized gold nanoparticles phyto-Au/FCVE (1, 2) and phyto-Au/TrX100/FCVE (3, 4) in BRB (pH 3) in the absence (1, 3) and in the presence of 0.02 mM NO2 (2, 4). Potential scan rate: 0.05 Vs−1.
Figure 3. Cyclic voltammograms on the phytosynthesized gold nanoparticles phyto-Au/FCVE (1, 2) and phyto-Au/TrX100/FCVE (3, 4) in BRB (pH 3) in the absence (1, 3) and in the presence of 0.02 mM NO2 (2, 4). Potential scan rate: 0.05 Vs−1.
Chemosensors 08 00078 g003
Figure 4. Dependence of the nitrite oxidation peak current on the logarithm of surfactant concentration (mM) used for the FCVE modification: cetyltrimethylammonium bromide (CTAB) (1), sodiumdodecyl sulfate (SDS) (2) and TrX100 (3). Solution: BRB (pH 3) + 0.02 mM NO2.
Figure 4. Dependence of the nitrite oxidation peak current on the logarithm of surfactant concentration (mM) used for the FCVE modification: cetyltrimethylammonium bromide (CTAB) (1), sodiumdodecyl sulfate (SDS) (2) and TrX100 (3). Solution: BRB (pH 3) + 0.02 mM NO2.
Chemosensors 08 00078 g004
Figure 5. SEM-images of the FCVE (a) and the TrX100/FCVE (b) surfaces.
Figure 5. SEM-images of the FCVE (a) and the TrX100/FCVE (b) surfaces.
Chemosensors 08 00078 g005
Figure 6. Cyclic voltammograms of 1 mM K4[Fe(CN)6] on (3) FCVE and (4) TrX100/FCVE and (1, 2) in a supporting 0.1 M KCl solution. Potential scan rate: 0.05 Vs−1.
Figure 6. Cyclic voltammograms of 1 mM K4[Fe(CN)6] on (3) FCVE and (4) TrX100/FCVE and (1, 2) in a supporting 0.1 M KCl solution. Potential scan rate: 0.05 Vs−1.
Chemosensors 08 00078 g006
Figure 7. Chronoamperograms obtained at 0.9 V potential on the FCVE (1) and the TrX100/FCVE (2) in 0.1 M KCl containing 1.0 mM K4[Fe(CN)6] (a). Dependencies I = f (t−1/2) are obtained from chronoamperograms on the corresponding electrodes (b).
Figure 7. Chronoamperograms obtained at 0.9 V potential on the FCVE (1) and the TrX100/FCVE (2) in 0.1 M KCl containing 1.0 mM K4[Fe(CN)6] (a). Dependencies I = f (t−1/2) are obtained from chronoamperograms on the corresponding electrodes (b).
Chemosensors 08 00078 g007
Figure 8. Effect of background electrolyte pH on the nitrite oxidation peak current (Ip) and peak potential (Ep) on the TrX100/FCVE. Solution: BRB + 0.02 mM NO2.
Figure 8. Effect of background electrolyte pH on the nitrite oxidation peak current (Ip) and peak potential (Ep) on the TrX100/FCVE. Solution: BRB + 0.02 mM NO2.
Chemosensors 08 00078 g008
Figure 9. Linear sweep voltammograms of the TrX100/FCVE in BRB (pH 3) containing 0.02 mM NO2 at different potential scan rates (0.02, 0.025, 0.04, 0.05, 0.075, 0.1, 0.15, 0.25 Vs−1) (a) and dependence of the oxidation peak potential on the natural logarithm of the potential scan rate (b).
Figure 9. Linear sweep voltammograms of the TrX100/FCVE in BRB (pH 3) containing 0.02 mM NO2 at different potential scan rates (0.02, 0.025, 0.04, 0.05, 0.075, 0.1, 0.15, 0.25 Vs−1) (a) and dependence of the oxidation peak potential on the natural logarithm of the potential scan rate (b).
Chemosensors 08 00078 g009
Figure 10. Chronoamperograms obtained at potential 1.1 V on the TrX100/FCVE in BRB (pH 3) containing 0.02 mM NO2 (a) and corresponding dependence I = f (t−1/2) (b). Insert: (a) chronoamperograms obtained at potential 1.1 V on the FCVE in BRB (pH 3) containing 0.02 mM NO2 and (b) corresponding dependence I = f (t−1/2).
Figure 10. Chronoamperograms obtained at potential 1.1 V on the TrX100/FCVE in BRB (pH 3) containing 0.02 mM NO2 (a) and corresponding dependence I = f (t−1/2) (b). Insert: (a) chronoamperograms obtained at potential 1.1 V on the FCVE in BRB (pH 3) containing 0.02 mM NO2 and (b) corresponding dependence I = f (t−1/2).
Chemosensors 08 00078 g010
Figure 11. Linear sweep voltammograms of the TrX100/FCVE at different nitrite concentrations (a) and corresponding dependences Ip vs. CNO2−on the TrX100/FCVE (b,c).
Figure 11. Linear sweep voltammograms of the TrX100/FCVE at different nitrite concentrations (a) and corresponding dependences Ip vs. CNO2−on the TrX100/FCVE (b,c).
Chemosensors 08 00078 g011
Table 1. Interfering effect of various foreign species on nitrite response (nitrite response at CNO2− = 5 μM taken as 100%).
Table 1. Interfering effect of various foreign species on nitrite response (nitrite response at CNO2− = 5 μM taken as 100%).
Interfering
Substance
Concentration of Interfering Substance, μMResponse Change, %
Glucose10000.0
Cl10000.0
SO42−10000.0
Fe(III)10000.0
CH3COO5000.0
Co(II)425−3.1
Ni(II)150−4.5
Cu(II)150−3.3
BrO3110−2.3
Citric acid50−2.9
Ascorbic acid50−0.5
Table 2. Analytical characteristics of nitrite determination in real samples for different sensors.
Table 2. Analytical characteristics of nitrite determination in real samples for different sensors.
Sensor *Limit of Detection, μMLinear Range, μMTechnique **SampleRef.
GO/PEDOT:PSS/SPCE0.0180.05–16.55Ambackground solution[16]
rGO/AuNPs/SPCE0.131–6000DPV, CVpurified water, packaged mineral water, dried shrimps, cured/salted fish, sausage[17]
AuNPs-PEI/GSPE1.01–10DPVbackground solution[18]
CuAgNPs/SPCEanodized11.1/15.620–370CVtap water, river water
deionized water
[19]
MnO2/GO-SPCE0.090.1–1
1–1000
Amtap water, packaged water[20]
GR+ β-CD/SPCE0.260.7–2150Amdrinking water
tap water
[21]
AgMCs-PAA/PVA/SPCE4.452–800FI-Amham, bacon, fermented pork, and sausage[22]
Ag/rGO/β-CD/SPCE0.241–2000LSVspiked pickles[23]
EACFE1.01.0–1000Cmbackground solution[24]
AuNPs/GCFE0.951.0–3350CVmustard[25]
ACFPE0.070.1–3838.5Ammineral water, sausage[26]
PB/[Bmim][BF4]-GFE0.0131.0–8.0CVtap water[28]
CFE0.030.25–3838.5Ammineral water, sausage[32]
SmFeO3/CFE5050–1000Ambackground solution[33]
rGO/β-CD/CdS/SPCE0.0210.05–447Amtap water and river water[51]
MWCNT/SPE0.021–500DPV, Amsaliva, urine, and blood samples[52]
AuNPs/PMB/PGE0.3145–5000DPVsausage, mineral water[53]
ccNiR + carbon ink/SPE1.20.7–370CVdrinking water, tap water, milk, urine, plasma[54]
TrX100/FCVE0.010.1–0.9
0.9–100
LSVsausage products, waterthis paper
* GO—graphene oxide; PEDOT:PSS—poly(3,4 ethylenedioxythiophene): polystyrene sulfonat; SPCE—screen-printed carbon electrode; rGO—reduced graphene oxide; β-CD—β-cyclodextrin; AuNPs—gold nanoparticles; MWCNT—multi-walled carbon nanotube; SPE—screen-printed electrode; PEI—polyethyleneimine; SPE—graphite screen printed electrodes; CuAgNPs—copper-silver nanoparticles; PMB—poly(methylene blue); PGE—pencil graphite electrode; GR—graphite; ccNiR—cytochrome cnitrite reductase; AgMCs—silver microcubics; PAA/PVA—polyacrylic acid/poly vinyl alcohol; ACFPE—annealed carbon fiber paper electrode; CFE—carbon felt electrode; EACFE—electrochemical activated carbon felt electrode; PB—Prussian blue; [Bmim][BF4]–1-butyl-3-methylimidazolium tetrafluoroborate; GCFE—graphitizing carbon felt electrode. ** Am—amperometry; Cm—coulometry; CV—cyclic voltammetry; DPV—differential pulse voltammetry; FI-Am—flow injection amperometry; LSV—linear sweep voltammetry.
Table 3. Results of nitrite determination in sausage extracts by applying the TrX100/FCVE (n = 3, p = 0.95) and the added-found method.
Table 3. Results of nitrite determination in sausage extracts by applying the TrX100/FCVE (n = 3, p = 0.95) and the added-found method.
SampleFound in Extract, μMAdded, μMFound in Extract with Additive, μMFound Additive, μMR, %
Sausages24 ± 22043 ± 119 ± 193
“Doctorskaya” sausage88 ± 180163 ± 176 ± 595
“Molochnaya” sausage228 ± 6300512 ± 19284 ± 1795
Table 4. Results of NO2 determination in sausage extracts applying the TrX100/FCVE and the UV-spectroscopic method (n = 3, p = 0.95).
Table 4. Results of NO2 determination in sausage extracts applying the TrX100/FCVE and the UV-spectroscopic method (n = 3, p = 0.95).
SampleVoltammetry with TrX100/FCVE, μMRSD, %Spectrophotometry, μMRSD, %F-Testt-Test
Sausages24 ± 22.623 ± 11.53.212.29
“Doctorskaya” sausage88 ± 10.587 ± 20.83.450.08
“Molochnaya” sausage228 ± 61.2239 ± 111.93.022.43
Fcrit = 19.00 (P = 0.95; df1 = 2, df2 = 2), tcrit = 2.78, P = 0.95; df = 4.
Table 5. Determination of NO2 in water samples, using the TrX100/FCVE (n = 3, p = 0.95).
Table 5. Determination of NO2 in water samples, using the TrX100/FCVE (n = 3, p = 0.95).
Water SampleAdded, μMFound, μMRSD,%R,%
Sample 10.991.02 ± 0.062.3103
Sample 24.954.87 ± 0.311.998
Sample 310.09.8 ± 0.21.199
Table 6. Results of NO2 determination in water samples applying the TrX100/FCVE and the UV-spectroscopic method (n = 3, p = 0.95).
Table 6. Results of NO2 determination in water samples applying the TrX100/FCVE and the UV-spectroscopic method (n = 3, p = 0.95).
Water SampleVoltammetry with TrX100/FCVE, μMRSD,%Spectrophotometry, μMRSD,%F-testt-test
Sample 11.02 ± 0.062.30.96 ± 0.124.94.191.32
Sample 24.87 ± 0.311.95.15 ± 0.342.72.181.90
Sample 39.8 ± 0.21.19.3 ± 0.52.24.332.67
Fcrit = 19.00 (P = 0.95; df1 = 2; df2 = 2), tcrit = 2.78; P = 0.95; df = 4.

Share and Cite

MDPI and ACS Style

Stozhko, N.Y.; Bukharinova, M.A.; Khamzina, E.I.; Tarasov, A.V.; Sokolkov, S.V. Film Carbon Veil-Based Electrode Modified with Triton X-100 for Nitrite Determination. Chemosensors 2020, 8, 78. https://0-doi-org.brum.beds.ac.uk/10.3390/chemosensors8030078

AMA Style

Stozhko NY, Bukharinova MA, Khamzina EI, Tarasov AV, Sokolkov SV. Film Carbon Veil-Based Electrode Modified with Triton X-100 for Nitrite Determination. Chemosensors. 2020; 8(3):78. https://0-doi-org.brum.beds.ac.uk/10.3390/chemosensors8030078

Chicago/Turabian Style

Stozhko, Natalia Yu., Maria A. Bukharinova, Ekaterina I. Khamzina, Aleksey V. Tarasov, and Sergey V. Sokolkov. 2020. "Film Carbon Veil-Based Electrode Modified with Triton X-100 for Nitrite Determination" Chemosensors 8, no. 3: 78. https://0-doi-org.brum.beds.ac.uk/10.3390/chemosensors8030078

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