Next Article in Journal
Short-Term Effect of Cigarette Smoke on Exhaled Volatile Organic Compounds Profile Analyzed by an Electronic Nose
Next Article in Special Issue
Affinity Assays for Cannabinoids Detection: Are They Amenable to On-Site Screening?
Previous Article in Journal
Paper-Based Fluidic Sensing Platforms for β-Adrenergic Agonist Residue Point-of-Care Testing
Previous Article in Special Issue
The Integration of Gold Nanoparticles with Polymerase Chain Reaction for Constructing Colorimetric Sensing Platforms for Detection of Health-Related DNA and Proteins
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Novel Amperometric Biosensor Based on Tyrosinase/Chitosan Nanoparticles for Sensitive and Interference-Free Detection of Total Catecholamine

1
Department of Experimental Medicine, Sapienza University of Rome, Viale Regina Elena 324, 00166 Rome, Italy
2
Department of Biological and Ecological Sciences, University of Tuscia, 01100 Viterbo, Italy
3
Department of Chemistry and Drug Technologies, Sapienza University of Rome, Piazzale Aldo Moro 5, 00185 Rome, Italy
*
Author to whom correspondence should be addressed.
Biosensors 2022, 12(7), 519; https://doi.org/10.3390/bios12070519
Submission received: 27 June 2022 / Revised: 7 July 2022 / Accepted: 9 July 2022 / Published: 12 July 2022

Abstract

:
The regulation of nervous and cardiovascular systems and some brain-related behaviors, such as stress, panic, anxiety, and depression, are strictly dependent on the levels of the main catecholamines of clinical interest, dopamine (DA), epinephrine (EP), and norepinephrine (NEP). Therefore, there is an urgent need for a reliable sensing device able to accurately monitor them in biological fluids for early diagnosis of the diseases related to their abnormal levels. In this paper, we present the first tyrosinase (Tyr)-based biosensor based on chitosan nanoparticles (ChitNPs) for total catecholamine (CA) detection in human urine samples. ChitNPs were synthetized according to an ionic gelation process and successively characterized by SEM and EDX techniques. The screen-printed graphene electrode was prepared by a two-step drop-casting method of: (i) ChitNPS; and (ii) Tyr enzyme. Optimization of the electrochemical platform was performed in terms of the loading method of Tyr on ChitNPs (nanoprecipitation and layer-by-layer), enzyme concentration, and enzyme immobilization with and without 1-ethyl-3-(3-dimethylaminopropyl)-carbodiimide (EDC) and N-hydroxysuccinimide (NHS) as cross-linking agents. The Tyr/EDC-NHS/ChitNPs nanocomposite showed good conductivity and biocompatibility with Tyr enzyme, as evidenced by its high biocatalytic activity toward the oxidation of DA, EP, and NEP to the relative o-quinone derivatives electrochemically reduced at the modified electrode. The resulting Tyr/EDC-NHS/ChitNPs-based biosensor performs interference-free total catecholamine detection, expressed as a DA concentration, with a very low LOD of 0.17 μM, an excellent sensitivity of 0.583 μA μM−1 cm−2, good stability, and a fast response time (3 s). The performance of the biosensor was successively assessed in human urine samples, showing satisfactory results and, thus, demonstrating the feasibility of the proposed biosensor for analyzing total CA in physiological samples.

1. Introduction

Catecholamines (Cas) are both neurotransmitters and hormones that play fundamental roles in central nervous and cardiovascular systems [1,2,3]. High CA levels, in particular dopamine (DA), epinephrine (EP), and norepinephrine (NEP), indicate cardiotoxicity leading to tachycardia and heart failure [4]. In contrast, low levels of CA are responsible for several neurological diseases, such as Parkinson’s disease, Alzheimer’s disease, epilepsy, stress, and depression [5,6]. Therefore, accurate monitoring of these important biomarkers in serum, blood, and urine is very important in early diagnostics [7].
To address this issue, several analytical methods have been developed for in vitro CA monitoring, such as high-performance liquid chromatography (HPLC) [8], HPLC with tandem mass spectrometric detection [9,10,11], fluorescence [12], and field-effect transistor [13] methods. These methods show high sensitivity and selectivity but require sophisticated and expensive equipment, time-consuming procedures, and skilled operators. Electrochemical methods represent an interesting alternative thanks to their unique advantages, such as their low cost, rapid response, simple operation, and capability for in situ detection [14,15,16,17].
Unlike glucose, CAs are redox-active compounds that are easily oxidized in a potential window between 0 and 0.4 V vs. Ag/AgCl [18], depending on the electrode modification. Therefore, total CA concentration can effectively be measured without the use of redox mediators and/or enzymes. The direct oxidation of DA, EP, and NEP on different electrode platforms allowed for the development of several sensors for CA detection [19,20,21,22,23,24,25]. However, these sensors suffer from relatively high oxidation potential and electrode surface passivation values due to the electrogenerated phenoxy radicals [26,27]. To this end, attempts at the bio-catalytic oxidation of CA, in particular DA and NEP, have been realized by integrating the Tyr enzyme with conductive nanomaterials and/or polymers on the electrode surface [28,29], leading to significant increases in DA and NEP oxidation peaks with concomitant cathodic shifts in their peak potential. Nevertheless, signal interference from other biological molecules, such as ascorbic acid (AA) and uric acid (UA), which show similar oxidation potentials with overlapping voltametric signals, represents another critical issue for CA detection based on oxidation signals that may limit its applicability. To address this issue, Tyr-based biosensors for CA detection are mainly used by monitoring the reduction signal of the bio-catalytically produced quinone compounds at a relatively moderate potential (about 0.1 V vs. Ag/AgCl on a carbon electrode) [30,31,32], where no electrochemical interference signal is present. Nanoparticles and/or nanostructures have been integrated into different sensing platforms to enhance both biosensor sensitivity and biosensor selectivity toward CA detection [32,33,34,35,36,37,38,39,40].
Tyrosinase (Tyr, polyphenol oxidase, EC 1.14.18.1) is a binuclear blue copper protein that acts as a polyphenol oxidase [41]. It catalyzes two consecutive oxidation reactions in presence of molecular oxygen, showing both monophenolase and diphenolase activity: (i) the o-hydroxylation of phenols to guaiacol; and (ii) the oxidation of guaiacol to o-quinones. The biosensor principle is the measurement of the increase in the cathodic signal due to the electrochemical reduction of the o-catecholamine quinone involving two electrons and two protons (Reaction 2), enzymatically generated on the electrode surface by Tyr (Reaction 1), according to the following reactions:
Tyr catecholamine + ½ O2 → o-catecholamine-quinone + H2O
o-catecholamine-quinone + 2H+ + 2e → catecholamine + H2O
The signal amplification is assured by the recycling of the o-catecholamine quinone by Tyr enzyme, leading to an enhancement of the selectivity and sensitivity of the biosensor. Therefore, the key factor in biosensor construction is adequate and effective Tyr immobilization in order to enhance the catalytic activity. Common approaches utilized for Tyr immobilization include carbon paste immobilization [42], sol–gel immobilization [43], entrapment within electropolymerized conducting polymers [37], and physical adsorption [44]. A considerable amount of attention has been paid to the use of chitosan (Chit), a polysaccharidic biopolymer composed of glucosamine with proper surface functional groups for biological/chemical binding and/or rapid adsorption [45,46]. It has found a great deal of applicability in sensing applications because of its non-toxicity, biocompatibility, biodegradability, and low cost [45]. Thanks to its polycationic character, chitosan can lock negatively charged proteins by charge attractions and hydrogen-bond interactions. Moreover, the presence of reactive amino and hydroxyl groups in its linear structure allows for covalent protein immobilization through the coupling chemistry via glutaraldehyde or EDC/NHS cross-linking agents [47,48]. However, its poor electrical conductivity may limit its applicability. To overcome this issue, chitosan has generally been combined with nanomaterials, such as graphene [49], multi-walled carbon nanotubes [50], and metallic nanoparticles [20,34,36], as well as conducting polymers, such as polyaniline and polypyrrole [51,52]. Chitosan nanoparticles (ChitNPs) with a size smaller than 100 nm showed superior chemical and biological performances than pristine chitosan, making them an excellent source for biomedical and biotechnological applications [53,54]. Thanks to their high surface–volume ratio, they can be utilized in biosensing applications as enzyme immobilization supports to achieve ultrasensitive detection of biomolecules. To date, only a few reports have been published with this aim. Anusha et al. [55] immobilized glucose oxidase (GOx) over ChitNPs on a gold electrode and, subsequently, Kim et al. [56] developed a platform based on a poly-γ-glutamic acid/ChitNPs hydrogel, which embedded GOx and magnetic nanoparticles.
In this paper, we report a new biosensing platform for the effective detection of total catecholamine based on the immobilization of Tyr in ChitNPs by using the zero-length EDC-NHS cross-linker on a graphene screen-printed electrode. To the best of our knowledge, no reports are present in the literature on electrode platforms based on a ChitNPs/EDC-NHS/Tyr nanocomposite for CA detection.

2. Experimental

2.1. Chemicals and Reagents

Tyrosinase (EC 1.14.18.1, 8503 U mg−1 from mushroom, Tyr), dopamine hydrochloride (DA), epinephrine (EP), norepinephrine (NEP), chitosan (low molecular weight) (Chit), sodium tripolyphosphate (TPP), 1-ethyl-3-(3-dimethylaminopropyl) carbodiimide (EDC), N-hydroxysuccinimide (NHS), ascorbic acid (AA), and uric acid (UA) were purchased from Sigma-Aldrich (St. Louis, MO, USA). Phosphate buffer was prepared with Na2HPO4 and NaH2PO4. NaOH and HCl were used to adjust pH values. Graphene screen-printed carbon electrodes (GPH/SPEs), used for biosensor construction, were purchased from Dropsens (Metrohm Dropsense, Asturias, Spain, model DRP-110GPH, screen-printed carbon electrodes modified with graphene).

2.2. Apparatus

Electrochemical measurements were performed in a 10 mL conventional three-electrode thermostated glass cell (model 6.1415.150, Metrohm, Herisau, Switzerland) using a graphene screen-printed carbon electrode (GPH/SPE) as a working electrode, an external Ag/AgCl/KClsat electrode (198 mV vs. NHE) as a reference electrode (cat. 6.0726.100, Metrohm, Herisau, Switzerland), and a glassy carbon rod as a counter electrode (cat. 6.1248.040, Metrohm, Herisau, Switzerland). An Autolab Potentiostat/Galvanostat (Eco Chemie, The Netherlands) was utilized in the electrochemical measurements. Scanning electron microscopy (SEM) and energy-dispersive X-ray spectroscopy (EDX) measurements were performed with a High-Resolution Field Emission Scanning Electron Microscope (HR FESEM, Zeiss Auriga Microscopy, Jena, Germany). The chronoamperometry experiments were carried out using a Sensit/SMART portable potentiostat (PalmSens, Houten, The Netherlands), a smartphone-based sensing device directly connected to a smartphone for POC signal reading.

2.3. Synthesis of Chitosan Nanoparticles

Chitosan nanoparticles (ChitNPs) were synthesized according to an ionic gelation process by adding 9 mL of chitosan dispersion (1 mg mL−1) to 3 mL of TPP solution (1 mg mL−1) and stirring the solution for 2 h at 25 °C with gentle magnetic stirring until an opalescent solution formed. The solution was successively centrifuged for 15 min at 6000 rpm, and the precipitated pellets were rinsed with DI water, dried, and used for further characterization [57].

2.4. Fabrication of TYR/ChitNPs-Modified Electrodes

Two different approaches were employed to immobilize Tyr and ChitNPs on a graphene SPE (the layer-by-layer (LbL) method and the nanoprecipitation (Np) method).

2.4.1. Layer-by-Layer Method

The layer-by-layer method (LbL) consists of the direct deposition “layer-by-layer” of each component of the electrochemical platform (ChitNPs and Tyr) onto the electrode. LbL-Tyr/ChitNPs/GPH/SPE was prepared as follows:
(1)
A total of 6 μL of ChitNPs solution (1 mg mL−1 in a water solution) was drop-cast onto GPH/SPE and left to dry at room temperature in air for approximately 1 h;
(2)
A total of 6 μL of Tyr solution (0.5 mg mL−1 in PBS 0.1 M, pH 7.2, KCl 0.1 M) was drop-cast onto the GPH/SPE previously modified with ChitNPs and left to dry at room temperature in air for approximately 1 h.
In the case of LbL-Tyr/EDC-NHS/ChitNPs/GPH/SPE, an additional electrode modification step was introduced between Step 1 and Step 2 by drop-casting onto the electrode surface a mixture of 3 μL of EDC solution (0.5 mM in PBS 0.1 M, pH 7.2, KCl 0.1 M) and 3 μL of NHS solution (0.1 mM in PBS 0.1 M, pH 7.2, KCl 0.1 M).
Three electrodes of each type were prepared in order to evaluate the reproducibility of the biosensor.

2.4.2. Nanoprecipitation Method

In the nanoprecipitation method (Np), Tyr and ChitNPs are co-precipitated in the same solution. Briefly, a solution of ChitNPs (1 mg mL−1) was added to a solution of Tyr (0.5 mg mL−1) under gentle magnetic stirring for 2 h at room temperature and subsequently deposited onto the electrode surface. The Np-Tyr/ChitNPs/GPH/SPE was prepared by drop-casting 12 μL of the Tyr–ChitNPs solution onto the GPH/SPE, which was then dried at room temperature in air for approximately 1 h.
Three Np-Tyr/ChitNPs/GPH/SPEs were prepared in order to evaluate the reproducibility of the biosensor.

2.5. Electrochemical Measurements

Cyclic voltammetry (CV) experiments were performed in PBS at a scan rate of 10 mV s−1. Magnetically stirred PBS (0.1 M, pH 7.2, KCl 0.1 M) was used to perform amperometric measurements. Electrochemical impedance spectroscopy (EIS) was performed at the open circuit potential (OCP) without bias voltage in the 0.1–104 Hz frequency range using an ac signal with an amplitude of 10 mV. Differential pulse voltammetry (DPV) experiments were recorded from −0.6 to 0.6 V with an amplitude 20 mV and a step potential of −5 mV. Baseline corrections were done for all DPV data using the NOVA software.

2.6. Preparation of Human Urine Samples

Human urine samples were diluted 100-fold with PBS (0.1 M, pH 7.2; KCl 0.1 M) before all measurements.

3. Results and Discussion

3.1. Characterization of ChitNPs

The morphological analysis of ChitNPs was realized by field emission scanning electron microscopy (FE-SEM) and energy-dispersive X-ray spectroscopy (EDX). Figure 1 shows the FE-SEM images of a bare GPH/SPE before (Panel A) and after (Panel C) the modification with ChitNPs. The bare GPH/SPE showed a densely packed rough surface, consistent with previous literature [58]. The EDX spectrum confirmed the presence of only carbon and oxygen, with an atomic percentage, calculated from the quantification of the peaks, of about 98% and 2%, respectively (Figure 1, Panel B). The FE-SEM image of ChitNPs/GPH/SPE shows aggregates of spherical nanoparticles with different diameters ranging from 70 to 300 nm (Figure 1, Panel C). The presence of the ChitNP aggregates may be ascribed to the inter- and intra-molecular cross-linkages mediated by the interaction of the positively charged amino groups of Chit and the negatively charged TPP ions [59]. The EDX spectrum of ChitNPs/GPH/SPE (Figure 1, Panel D) confirmed the presence of C (96%), O (3%), and a lower amount of Na (0.5%), probably due to the use of the TPP cross-linker in the procedure for the synthesis of ChitNPs.
All GPH/SPEs and ChitNPs/GPH/SPEs were subsequently characterized by cyclic voltammetry (CV) experiments in a solution of Fe(CN)64− (data not shown) in order to calculate the electroactive area (AEA), the heterogeneous electron transfer rate constant (K0, cm s−1), and the roughness factor (electroactive–geometrical area ratio, ρ), which are reported in Table S1. The AEA was evaluated using the Randles–Sevick equation by the slope of the peak current vs. the square root of the scan rate (v1/2) [60]. The K0 was calculated using the extended method, which merges the Klingler–Kochi and Nicholson–Shain methods for totally irreversible and reversible systems, respectively [61,62].

3.2. Electrochemical Characterization of the Tyr/ChitNPs Biosensor Platform

The electrochemical characterization of the Tyr/ChitNPs platform was performed using cyclic voltammetry (CV) and electrochemical impedance spectroscopy (EIS).

3.2.1. Optimization of Methods for Loading Tyr on ChitNPs

Two different methods for loading Tyr on the ChitNPs/GHP/SPE platform were investigated by CV experiments: (i) the nanoprecipitation method (Np); and (ii) the layer-by-layer (LbL) method. Figure 2 shows the cyclic voltammograms of DA in the absence (ChitNPs/GPH/SPE, black line) and the presence of Tyr enzyme immobilized by the Np method (Np-Tyr/ChitNPs/GPH/SPE, red line), and the LbL method (LbL-Tyr/ChitNPs/GPH/SPE, blue line) in comparison with Tyr enzyme immobilized on a simple layer of unstructured chitosan (Tyr/Chit/GPH/SPE, green line) deposited by a drop-casting method. Two well-defined redox peaks, due to the oxidation/reduction of DA, are clearly visible on the bare electrode (black curve) without Tyr, while no redox peaks are present with the Tyr/Chit/GPH/SPE electrode (green curve), attesting to the lower conductivity of the unstructured chitosan layer [45] compared with nanostructured chitosan. A similar result was obtained by recording the CVs of DA on a chitosan-modified electrode in the absence of Tyr enzyme (Chit/GPH/SPE). The cyclic voltammogram of DA on ChitNPs/GPH/SPE shows two clear redox peaks (Figure S1, black curve), while smaller oxidation and reduction peaks in the reverse scan can be observed with unstructured chitosan (Figure S1, red curve), thus confirming that the nanostructuration strongly improves the chitosan’s conductivity. Similar behavior was described by Marroquin et al. [63], who reported that chitosan nanocomposite films realized with Fe3O4 and MWCNTs could simultaneously have enhanced the mechanical properties, thermal stability, and electrical conductivity.
An increase in the reduction peaks was observed for both Np-Tyr/ChitNPs/GPH/SPE (Figure 2, red curve) and LbL-Tyr/ChitNPs/GPH/SPE (Figure 2, blue curve) due to the reduction in the dopamine-o-quinone produced by the enzymatic reaction on the electrode surface to DA [39], demonstrating that the immobilization of Tyr within the nano-biopolymer greatly enhanced the catalytic activity of the enzyme. The reduction peak current of DA on LbL-Tyr/ChitNPs/GPH/SPE was about three times higher than that on Np-Tyr/ChitNPs/GPH/SPE, indicating the superior electrochemical performances of the LbL loading method in terms of the better electron transfer between the dopamine-o-quinone and the electrode surface [64]. The results reveal that the LbL immobilization method promotes the Tyr’s catalytic property better than the Np method and probably plays a significant role in: (i) the improvement of enzyme immobilization by favoring physical adsorption and the formation of electrostatic interactions between positively charged ChitNPs [57] and negatively charged Tyr enzyme [41]; and (ii) facilitating the direct electron transfer between the substrate molecules and the modified electrode surface.

3.2.2. Optimization of Tyr Concentration

The amount of immobilized enzyme is one of the most relevant parameters in electrode modification. Figure 3 shows the effect of the concentration of Tyr, immobilized by the LbL method on the Tyr/ChitNPs/GPH/SPE, on the cathodic peak current registered in the presence of 50 μM DA. The intensity current differences (∆I), calculated as the current differences obtained in the presence and the absence of Tyr, increase with decreasing concentrations of Tyr immobilized on the electrode surface. The optimal Tyr concentration was found to be 0.5 mg mL−1, attesting to the fact that electron transfer might be hindered at higher enzyme concentrations. Lower concentrations were also tested, showing no significant increase in the ∆I values.

3.2.3. Tyrosinase Covalent Functionalization via EDC-NHS Coupling Chemistry

In order to enhance the efficacy of the Tyr-based biosensor, the Tyr covalent immobilization strategy using EDC-NHS as a cross-linking agent was investigated. The EDC-NHS coupling chemistry takes advantage of the properties of the EDC-NHS cross-linker, which is able to form amide bonds between carboxyl and amine groups, as shown in Scheme 1 [48]. In particular, the carboxyl groups of ChitNPs immobilized on the electrode surface are activated by the EDC/NHS, forming a typical NHS-ester, and, subsequently, the amino group from the Tyr enzyme acts on this NHS-ester, forming a very strong covalent amidic linkage.
Figure 4 shows the cyclic voltammograms of DA in pH 7.2 PBS on ChitNPs/GPH/SPE (black line) and LbL-Tyr/ChitNPs/GPH/SPE without (blue line) and with (red line) the EDC-NHS cross-linking agent. The reduction peak current of DA observed on LbL-Tyr/EDC-NHS/ChitNPs/GPH/SPE was 1.5 times higher than that on LbL-Tyr/ChitNPs/GPH/SPE, attesting to the enhanced electrochemical performances of the method based on covalent functionalization compared with the LbL method utilizing simple electrostatic interactions, probably thanks to the higher stability of the covalently immobilized Tyr for enhanced biocatalysis.

3.2.4. Electrochemical Impedance Spectroscopy Behavior of the Modified Electrodes

The EIS technique was used to investigate the electrochemical properties of the modified electrodes after each surface modification step. Figure 5 shows the Nyquist plots of the ChitNPs/GPH/SPE (black), the Tyr/ChitNPs/GPH/SPE (blue), and the Tyr/EDC-NHS/ChitNPs/GPH/SPE (red) in 5 mM [Fe(CN)6]3−/4− solution. The semicircle diameter of the Nyquist plot represents the charge transfer resistance (Rct), which is related to small changes at the electrode–electrolyte interface in the solution. The immobilization of a protein biomolecule over the electrode surface causes an increase in the Rct value that is related to the hindering of the electron transfer between the redox probe and the electrode surface. The impedance spectra were fitted by using the simple Randles circuit (R(Q(RW))) reported in Figure 5. Experimental Rct values are in the following order: ChitNPs/GPH/SPE (357 Ω) < Tyr/ChitNPs/GPH/SPE (403 Ω) < Tyr/EDC-NHS/ChitNPs/GPH/SPE (436 Ω) due to the hampering effect of Tyr on the charge transfer. The Rct values confirm the successful immobilization of Tyr enzyme on the electrode surface, resulting in strong optimization with the use of the EDC-NHS cross-linker (red curve).

3.3. Amperometric Response of Dopamine

Figure 6 (Panel A and B) illustrates a typical steady-state current–time response upon successive additions of DA on Tyr/ChitNPs/GPH/SPE (Panel A) and Tyr/EDC-NHS/ChitNPs/GPH/SPE (Panel B), showing a fast response time (3 s). The measurements were carried out by immersing the modified electrode in blank PBS (0.1 M). Once a steady baseline was achieved, DA was added to the solution, causing an increase in the cathodic current due to the reduction in o-quinone generated by the enzymatic oxidation of DA [64]. Thus, the target DA molecule was regenerated during the reduction reaction at the electrode, allowing for the amplification of the electrochemical signal, which leads to enhanced sensitivity [65]. Figure 6 (Panel C) exhibits the typical calibration curves of the DA biosensors corresponding to the electrode platforms with (black curve) and without (red curve) EDC-NHS, both following Michelis–Menten kinetics. As can be seen in the inset of Panel C, the linear range spanned from 4 μM to 12 μM with a regression equation of y = 0.03325x + 0.0404 (R2 = 0.994) and from 0.5 μM to 5 μM with a regression equation of y = 0.80066x + 0.0039121 (R2 = 0.997) in the case of the Tyr/ChitNPs/GPH/SPE biosensor and the Tyr/EDC-NHS/ChitNPs/GPH/SPE biosensor, respectively. The biosensor based on the EDC-NHS cross-linker showed a lower LOD of 0.17 μM and a two-fold higher sensitivity of 0.583 μA mM−1 cm−2 compared with the corresponding biosensor without EDC-NHS. According to the Lineweaver–Burk equation [66], the apparent Michaelis–Menten constants ( K M app ) were 25.59 and 9.51 μM for the Tyr/ChitNPs/GPH/SPE biosensor and the Tyr/EDC-NHS/ChitNPs/GPH/SPE biosensor, respectively. It is interesting to note that the experimental K M app value is lower than the KM value of the free enzyme towards dopamine (KM = 2.2 mM) [67], attesting to the optimal accessibility of the target molecules to the active sites of the enzyme. The electrochemical characteristics of the two biosensors in terms of the K M app , maximum current density, linearity range, LOD, and sensitivity are summarized in Table 1.

3.4. Amperometric Response of Epinephrine and Norepinephrine

The Tyr/EDC-NHS/ChitNPs/GPH/SPE-based biosensor, which showed the best performance, was utilized for EP and NEP detection. Chronoamperometric experiments were performed at an applied potential of 0 V vs. Ag/AgCl as EP and NEP gave cathodic peaks at the same redox potential of DA (curves not shown). The corresponding calibration plots are reported in Figure 7, together with the calibration plot of DA (black curve) as a comparison. Table 2 summarizes the analytical and kinetic performances of the proposed biosensor for the three catecholamines studied, including the linear range, LOD, sensitivity, maximum current density, and K M app .   The sensitivity of the DA biosensor was 5 times higher than that of EP and NEP. The K M app values were in the following order: DA < NEP < EP, attesting to the stronger affinity and binding of the Tyr enzyme to DA, leading to a higher catalytic activity [27].
The total catecholamine concentration is therefore expressed as a DA concentration, thanks to the best analytical and kinetic performance of this biosensor.
Table 3 shows a comparison between the results obtained with our biosensor and other Tyr-based biosensors for catechol and catecholamine recently reported in the literature. It is possible to observe that the proposed ChitNPs/graphene platform shows superior analytical performances compared with the other platforms for CA detection, whereas inferior performances can be observed when compared with Tyr-based catechol biosensors. These results can be ascribed to the stronger affinity of the Tyr enzyme to catechol.

3.5. Selectivity

The selectivity of the Tyr/EDC-NHS/ChitNPs/GPH/SPE-based biosensor for DA was investigated by recording the current response in the presence of potential interferents, which can coexist in biological matrices. In particular, differential pulse voltammograms were recorded in a solution containing a 100-fold concentration of ascorbic acid (AA) and uric acid (UA) in the presence of DA. AA and UA usually show overlapping voltametric curves on both unmodified and chemically modified electrodes [23]. As shown in Figure 8, three well-resolved cathodic peaks were recorded at about −450, +100, and +400 mV for AA, DA, and UA, respectively. The results clearly demonstrate that AA and UA did not significantly interfere with DA detection, showing the high selectivity of the proposed biosensor.

3.6. Stability

The stability and lifetime of the biosensor for DA, EP, and NEP detection were investigated by measuring the DPV responses of the biosensor every 5 days (10 measurements) over a period of 30 days using a 10 µM solution of each CA (Figure S2). The biosensor was stored in a dry refrigerator at 4 °C in between the measurements. The stability was good as the biosensor retained about 90% of its initial current response after 30 days. These results may be ascribed to the synergistic effect of the bio-nanocomposite including ChitNPs and graphene, which reduces the enzyme denaturation and the subsequent loss of catalytic properties.

3.7. Real Sample Detection

Finally, the Tyr/EDC-NHS/ChitNPs/GPH/SPE-based biosensor was tested in human urine samples spiked with known concentrations of DA in order to evaluate the effectiveness of the biosensor in physiological samples. A fixed amount of DA (5 μM) was added to three urine samples without any sample pretreatment. All measurements were done in triplicate using chronoamperometry and the results are reported in Table 4. The recovery rates ranged from 94% to 98% and the relative standard deviation (RSD) values did not exceed 6.3%, demonstrating the efficacy of the proposed biosensor in real samples.
The results were subsequently compared to those obtained with a standard spectrophotometric method used as reference (Table 3). An unpaired t-test was performed, and the results show that there was no statistically significant difference in the DA quantification between the proposed biosensor and the standard spectrophotometric method (t = −2429; degrees of freedom = 4; p = 0.072). This result indicates the biosensor’s excellent capacity to provide mean values close to those obtained when the same samples are measured by using the standard spectrophotometric method.

4. Conclusions

In this work, we demonstrated the feasibility of developing an amperometric biosensor for total catecholamine detection by using a novel electrochemical nano-platform based on the immobilization of Tyr/ChitNPs on a graphene SPE. The ChitNPs were synthesized by the ionic gelation process and subsequently immobilized with Tyr enzyme on the graphene electrode through a layer-by-layer approach with and without EDC-NHS as a cross-linking agent. The ChitNPs provided a friendly environment for Tyr immobilization, thus enhancing the catalytic activity of the enzyme and, at the same time, showing an improved platform conduction pathway thanks to the higher conductivity compared with pristine chitosan. The ChitNPs/EDC-NHS/graphene nanocomposite matrix showed superior electrochemical performances, which can be attributed to the synergistic effect of graphene and ChitNPs. The excellent conductivity and large surface area of the graphene electrode in combination with the properties of the ChitNPs produced the good analytical performances of the developed catecholamine biosensor in terms of a short response time (3 s), high sensitivity, a low detection limit, good stability, and satisfactory recovery when tested in spiked human urine samples. At the current state of research, the proposed ChitNPs/graphene platform does not provide superior performance compared with other nanostructured platforms for CA detection reported in the literature, but the ease of preparation and low cost of the proposed ChitNPs make them a promising component and/or additive for electrochemical biosensor platforms thanks to the described beneficial properties.

Supplementary Materials

The following supporting information can be downloaded at: www.mdpi.com/article/10.3390/bios12070519/s1, Table S1: Electrochemical parameters related to GPH/SPE and ChitNPs/GPH/SPE modiefied electrode: electroactive area (AEA), roughness factor (ρ) and heterogenous electron transfer rate costant (K0). Experimental condition: 1.1 mM Fe(CN)64− in 0.1 M KCl; Figure S1: CVs of 50 μM Dopamine in PBS 0.1 M (pH 7.2; KCl 0.1 M), on a ChitNPs/GPH/SPE (black line) and Chit/GPH/SPE (red line); Figure S2: Stability measurements over a period of 30 days in presence of 10 μM of DA (Panel A), EP (Panel B) and NEP (Panel C) in PBS 0.1 M (pH 7.2; KCl 0.1 M); Figure S3: Optimization of applied potential with Tyr/EDC-NHS/ChitNPs/GPH/SPE in 50 mM dopamine in 0.1 M PBS (pH 7.2; KCl 0.1 M).

Author Contributions

Investigation: V.G., C.T.; Methodology: E.C.; Resources: A.A., A.L.; Supervision: R.A.; Writing original draft: V.G., R.A. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Italian Ministry of Education, Universities, and Research (Progetto di Ateneo 2022, No. RP12117A8B0CA73F).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to thank Andrea Isidori and Mary Anna Venneri of the Department of Experimental Medicine, Sapienza University of Rome, for the support provided.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Emran, M.Y.; Shenashen, M.A.; Abdelwahab, A.A.; Khalifa, H.; Mekawy, M.; Akhtar, N.; Abdelmottaleb, M.; EI-Safty, S.A.J. Design of hierarchical electrocatalytic mediator for one step, selective screening of biomolecules in biological fluid samples. Appl. Electrochem. 2018, 48, 529–542. [Google Scholar] [CrossRef]
  2. Li, L.; Lu, Y.Z.; Qian, Z.; Yang, Z.; Yang, K.; Zong, S.; Wang, Z.; Cui, Y. Ultra-sensitive surface enhanced Raman spectroscopy sensor for in-situ monitoring of dopamine release using zipper-like ortho-nanodimers. Biosens. Bioelectron. 2021, 180, 113100. [Google Scholar] [CrossRef] [PubMed]
  3. Liss, B.; Roeper, J. Individual dopamine midbrain neurons: Functional diversity and flexibility in health and disease. Brain Res. Rev. 2008, 58, 314–321. [Google Scholar] [CrossRef] [PubMed]
  4. Bucolo, C.; Leggio, G.M.; Drago, F.; Salomone, S. Dopamine outside the brain: The eye, cardiovascular system and endocrine pancreas. Pharmacol. Ther. 2019, 203, 107392. [Google Scholar] [CrossRef] [PubMed]
  5. Klein, M.O.; Battagello, D.S.; Cardoso, A.R.; Hauser, D.N.; Bittencourt, J.C.; Correa, R.G. Dopamine: Functions, Signaling, and Association with Neurological Diseases. Cell. Mol. Neurobiol. 2019, 39, 31–59. [Google Scholar] [CrossRef]
  6. Medina, M.Á.; Urdiales, J.L.; Rodríguez-Caso, C.; Ramírez, F.J.; Sánchez-Jiménez, F. Biogenic amines and polyamines: Similar biochemistry for different physiological missions and biomedical applications. Crit. Rev. Biochem. Mol. Biol. 2003, 38, 23–59. [Google Scholar] [CrossRef]
  7. Marc, D.T.; Ailts, J.W.; Ailts Campeau, D.C.; Bull, M.J.; Olson, K.L. Neurotransmitters excreted in the urine as biomarkers of nervous system activity: Validity and clinical applicability. Neurosci. Biobehav. Rev. 2011, 35, 635–644. [Google Scholar] [CrossRef]
  8. Carrera, V.; Sabater, E.; Vilanova, E.; Sogorb, M.A. A simple and rapid HPLC–MS method for the simultaneous determination of epinephrine, norepinephrine, dopamine and 5-hydroxy- tryptamine: Application to the secretion of bovine chromaffin cell cultures. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2007, 847, 88–94. [Google Scholar] [CrossRef]
  9. Mizuguchi, H.; Nishimori, D.; Kuwabara, T.; Takeuchi, M.; Iiyama, M.; Takayanagi, T. Track-etched membrane-based dual-electrode coulometric detector for microbore/capillary high-performance liquid chromatography. Anal. Chim. Acta 2020, 1102, 46–52. [Google Scholar] [CrossRef]
  10. Matuszewski, B.K.; Constanzer, M.L.; Chavez-Eng, C.M. Strategies for the Assessment of Matrix Effect in Quantitative Bioanalytical Methods Based on HPLC-MS/MS. Anal. Chem. 2003, 75, 3019–3030. [Google Scholar] [CrossRef]
  11. Moghzi, F.; Soleimannejad, J.; Sañudo, E.C.; Janczak, J. Dopamine Sensing Based on Ultrathin Fluorescent Metal–Organic Nanosheets. ACS Appl. Mater. Interfaces 2020, 12, 44499–44507. [Google Scholar] [CrossRef] [PubMed]
  12. Liu, N.; Xiang, X.; Fu, L.; Cao, Q.; Huang, R.; Liu, H.; Han, G.; Wu, L. Regenerative field effect transistor biosensor for in vivo monitoring of dopamine in fish brains. Biosens. Bioelectron. 2021, 188, 113340. [Google Scholar] [CrossRef] [PubMed]
  13. Ribeiro, J.A.; Fernandes, P.M.V.; Pereira, C.M.; Silva, F. Electrochemical sensors and biosensors for determination of catechola- mine neurotransmitters: A review. Talanta 2016, 160, 653–679. [Google Scholar] [CrossRef] [PubMed]
  14. Xu, C.; Gu, C.; Xiao, Q.; Chen, J.; Yin, Z.Z.; Liu, H.; Fan, K.; Li, L. A highly selective and sensitive biosensor for dopamine based on a surface molecularly imprinted layer to coordinate nano-interface functionalized acupuncture needle. Chem. Eng. J. 2022, 436, 135203. [Google Scholar] [CrossRef]
  15. Rahman, S.F.; Min, K.; Park, S.H.; Park, J.H.; Yoo, J.C.; Park, D.H. Selective determination of dopamine with an amperometric biosensor using electrochemically pretreated and activated carbon/tyrosinase/Nafion®-modified glassy carbon electrode. Biotechnol. Bioprocess Eng. 2016, 21, 627–633. [Google Scholar] [CrossRef]
  16. Rand, E.; Periyakaruppan, A.; Tanaka, Z.; Zhang, D.A.; Marsh, M.P.; Andrews, R.J.; Lee, K.H.; Chen, B.; Meyyappan, M.; Koehne, J.E. A carbon nanofiber-based biosensor for simultaneous detection of dopamine and serotonin in the presence of ascorbic acid. Biosens. Bioelectron. 2013, 42, 434–438. [Google Scholar] [CrossRef] [Green Version]
  17. De-Qian, H.; Cheng, C.; Yi-Ming, W.; Hong, Z.; Liang-Quan, S.; Hua-Jie, X.; Zhao-Di, L. The Determination of Dopamine Using Glassy Carbon Electrode Pretreated by a Simple Electrochemical Method. J. Electrochem. Sci. 2012, 7, 5510–5520. [Google Scholar]
  18. Gao, F.; Cai, X.; Wang, X.; Gao, C.; Liu, S.; Gao, F.; Wang, Q. Highly sensitive and selective detection of dopamine in the presence of ascorbic acid at graphene oxide modified electrode. Sens. Actuators B Chem. 2013, 186, 380–387. [Google Scholar] [CrossRef]
  19. Qiu, H.J.; Zhou, G.P.; Ji, G.L.; Zhang, Y.; Huang, X.R.; Ding, Y. A novel nanoporous gold modified electrode for the selective determination of dopamine in the presence of ascorbic acid. Coll. Surf. B Biointerfaces 2008, 69, 105–108. [Google Scholar] [CrossRef]
  20. Üge, A.; Zeybek, D.K.; Zeybek, B. An electrochemical sensor for sensitive detection of dopamine based on MWCNTs/CeO2-PEDOT composite. J. Electroanal. Chem. 2018, 15, 134–142. [Google Scholar] [CrossRef]
  21. Yang, C.; Jacobs, C.B.; Nguyen, M.D.; Ganesana, M.; Zestos, A.G.; Ivanov, I.N.; Puretzky, A.A.; Rouleau, C.M.; Geohegan, D.B.; Venton, B.J. Carbon nanotubes grown on metal microelectrodes for the detection of dopamine. Anal. Chem. 2016, 88, 645–652. [Google Scholar] [CrossRef] [PubMed]
  22. Deepika, J.; Sha, R.; Badhulika, S. A ruthenium (IV) disulfide based non-enzymatic sensor for selective and sensitive amperometric determination of dopamine. Microchim. Acta 2019, 186, 480. [Google Scholar] [CrossRef] [PubMed]
  23. Tortolini, C.; Cass, A.E.G.; Pofi, R.; Lenzi, A.; Antiochia, R. Microneedle-based nanoporous gold electrochemical sensor for real-time catecholamine detection. Microchim. Acta 2022, 189, 1–14. [Google Scholar] [CrossRef] [PubMed]
  24. Mphuthi, N.G.; Adekunle, A.S.; Ebenso, E.E. Electrocatalytic oxidation of epinephrine and norepinephrine at metal oxide doped phthalocyanine/MWCNT composite sensor. Sci. Rep. 2016, 6, 26938. [Google Scholar] [CrossRef] [Green Version]
  25. Wang, J.; Martinez, T.; Yaniv, D.R.; McCormick, L.D. Scanning tunneling microscopic investigation of surface fouling of glassy carbon surfaces due to phenol oxidation. J. Electroanal. Chem. Interfacial Electrochem. 1991, 313, 129–140. [Google Scholar] [CrossRef]
  26. Lane, R.F.; Blaha, C.D. Detection of catecholamines in brain tissue: Surface-modified electrodes enabling in vivo investigations of dopamine function. Langmuir 1990, 6, 56–65. [Google Scholar] [CrossRef]
  27. Florescu, M.; David, M. Tyrosinase-based biosensors for selective dopamine detection. Sensors 2017, 17, 1314. [Google Scholar] [CrossRef] [Green Version]
  28. Apetrei, I.M.; Apetrei, C. Development of a novel biosensor based on tyrosinase/platinum nanoparticles/Chitosan/graphene nanostructured layer with applicability in bioanalysis. Materials 2019, 12, 1009. [Google Scholar] [CrossRef] [Green Version]
  29. Baluta, S.; Lesiak, A.; Cabaj, J. Simple and cost-effective electrochemical method for norepinephrine determination based on carbon dots and tyrosinase. Sensors 2020, 20, 4567. [Google Scholar] [CrossRef]
  30. Apetrei, I.M.; Apetrei, C. Biosensor based on tyrosinase immobilized on a single-walled carbon nanotube-modified glassy carbon electrode for detection of epinephrine. Int. J. Nanomed. 2013, 8, 4391–4398. [Google Scholar] [CrossRef] [Green Version]
  31. Wang, Y.; Zhai, F.; Hasebe, Y.; Jia, H.; Zhang, Z. A highly sensitive electrochemical biosensor for phenol derivatives using a graphene oxide-modified tyrosinase electrode. Bioelectrochem 2018, 122, 174–182. [Google Scholar] [CrossRef] [PubMed]
  32. Janegitz, B.C.; Medeiros, R.A.; Rocha-Fillo, R.C.; Fatibello-Filho, O. Direct electrochemistry of tyrosinase and biosensing for phenol based on gold nanoparticles electrodeposited on a boron-doped diamond electrode. Diam. Relat. Mater. 2012, 25, 128–133. [Google Scholar] [CrossRef]
  33. Yang, L.; Xiong, H.; Zhang, X.; Wang, S. A novel tyrosinase biosensor based on chitosan-carbon-coated nickel nanocomposite film. Bioelectrochem 2012, 84, 44–48. [Google Scholar] [CrossRef] [PubMed]
  34. Vicentini, F.C.; Janegitz, B.C.; Brett, C.M.A.; Fatibello-Filho, O. Tyrosinase biosensor based on a glassy carbon electrode modified with multi-walled carbon nanotubes and 1-butyl-3-methylimidazolium chloride within a dihexadecylphosphate film. Sens. Actuators B Chem. 2013, 188, 1101–1108. [Google Scholar] [CrossRef]
  35. Fartas, F.M.; Abdullah, J.; Yusof, N.A.; Sulamain, Y.; Saiman, M.I. Biosensor based on tyrosinase immobilized on graphene-decorated gold nanoparticle chitosan for phenolic detection in aqueous. Sensors 2017, 17, 1132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Apetrei, C.; Rodriguez-Mendez, M.L.; De Saja, J.A. Amperometric tyrosinase based biosensor using an electropolymerized phosphate-doped polypyrrole film as an immobilization support. Application for detection of phenolic compounds. Electrochim. Acta 2011, 56, 8919–8925. [Google Scholar] [CrossRef]
  37. Wee, Y.; Park, S.; Kwon, Y.H.; Ju, Y.; Yeon, K.M.; Kim, J. Tyrosinase-immobilized CNT based biosensor for highly-sensitive detection of phenolic compounds. Biosens. Bioelectron. 2019, 132, 279–285. [Google Scholar] [CrossRef] [PubMed]
  38. Da Silva, W.; Ghica, M.E.; Ajayi, R.F.; Iwuoha, E.I.; Brett, C.M.A. Tyrosinase based amperometric biosensor for determination of tyramine in fermented food and beverages with gold nanoparticles doped poly(8-anilino-1-naphtalene sulphonic acid) modified electrode. Food Chem. 2019, 282, 18–26. [Google Scholar] [CrossRef]
  39. Kim, D.S.; Kang, E.S.; Seungho, B.; Choo, S.S.; Chung, Y.H.; Lee, D.; Min, J.; Kim, T.H. Electrochemical detection of dopamine using periodic cylindrical gold nanoelectrode arrays. Sci. Rep. 2018, 8, 14049. [Google Scholar] [CrossRef] [Green Version]
  40. Sainio, S.; Palomaki, T.; Tujunen, N.; Prototopova, V.; Koene, J.; Kordas, K.; Koskinen, J.; Meyyappan, M.; Laurila, T. Integrated Carbon Nanostructures for Detection of Neurotransmitters. Mol. Neurobiol. 2015, 52, 859–866. [Google Scholar] [CrossRef]
  41. Sanchez-Ferrer, A.; Rodrìguez-Lòpez, J.N.; Garcìa-Cánovas, F.; Garcìa-Camona, F. Tyrosinase: A comprehensive review of its mechanism. Biochim. Biophys. Acta 1995, 1247, 1–11. [Google Scholar] [CrossRef]
  42. Granero, A.M.; Fernandez, H.; Agostini, E.; Zon, M.L. An amperometric biosensor based on peroxidases from Brassica napus for the determination of the total polyphenolic content in wine and tea samples. Talanta 2010, 83, 249–255. [Google Scholar] [CrossRef] [PubMed]
  43. Zejli, H.; de Cisneros, H.H.; Naranjo-Rodriguez, I.; Liu, B.; Temsamani, K.R.; Marty, J.L. Phenol biosensor based on Sonogel-Carbon transducer with tyrosinase alumina sol–gel immobilization. Anal. Chim. Acta 2008, 612, 198–203. [Google Scholar] [CrossRef]
  44. Shiddiky, M.J.A.; Torriero, A.A.J. Application of ionic liquids in electrochemical sensing systems. Bios. Bioelectron. 2011, 26, 1775–1787. [Google Scholar] [CrossRef] [PubMed]
  45. Petrucci, R.; Pasquali, M.; Scaramuzzo, F.A.; Curulli, A. Recent Advances in Electrochemical Chitosan-Based Chemosensors and Biosensors: Applications in Food Safety. Chemosensors 2021, 9, 254. [Google Scholar] [CrossRef]
  46. Karrat, A.; Amine, A. Recent advances in chitosan-based electrochemical sensors and biosensors. Arab. J. Chem. Environ. Res. 2020, 7, 66–93. [Google Scholar]
  47. Fisher, M.J.E. Amine coupling through EDC/NHS: A practical approach. Methods Mol. Biol. 2010, 627, 55–73. [Google Scholar]
  48. Ducker, R.E.; Montague, M.T.; Leggett, G.J. A comparative investigation of methods for protein immobilization on self-assembled monolayers using glutaraldehyde, carbodiimide, and anhydride reagents. Biointerphases 2008, 3, 59. [Google Scholar] [CrossRef]
  49. Liu, C.Y.; Chou, Y.C.; Tsai, J.H.; Huang, T.M.; Chen, J.Z.; Yeh, Y.C. Tyrosinase/Chitosan/Reduced Graphene Oxide Modified Screen-Printed Carbon Electrode for Sensitive and Interference-Free Detection of Dopamine. Appl. Sci. 2019, 9, 622. [Google Scholar] [CrossRef] [Green Version]
  50. Cheng, Y.; Liu, Y.; Huang, J.; Xian, Y.; Zhang, Z.; Jin, L. Fabrication of Tyrosinase Biosensor Based on Multiwalled Carbon Nanotubes-Chitosan Composite and Its Application to Rapid Determination of Coliforms. Electroanalysis 2008, 20, 1463–1469. [Google Scholar] [CrossRef]
  51. Wan, D.; Yuan, S.; Li, G.L.; Neoh, K.G.; Kang, E.T. Glucose biosensor from covalent Immobilization of Chitosan-Coupled Carbon Nanotubes on Polyaniline-Modified Gold Electrode. ACS Appl. Mater. Interfaces 2010, 2, 3083–3091. [Google Scholar] [CrossRef] [PubMed]
  52. Lupu, S.; Lete, C.; Cătălin Balaure, P.; Ion Caval, D.; Mihailciuc, C.; Lakard, B.; Hihn, J.Y.; del Campo, F.J. Development of Amperometric Biosensors Based on Nanostructured Tyrosinase-Conducting Polymer Composite Electrodes. Sensors 2013, 13, 6759–6774. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Gao, Y.; Wu, Y. Recent advances of chitosan-based nanoparticles for biomedical and biotechnological applications. Int. J. Biol. Macromol. 2022, 203, 379–388. [Google Scholar] [CrossRef] [PubMed]
  54. Wang, J.; Zhuang, S. Chitosan-based materials: Preparation, modification and application. J. Clean. Prod. 2022, 355, 131825. [Google Scholar] [CrossRef]
  55. Anusha, J.R.; Raj, C.J.; Cho, B.-B.; Fleming, A.T.; Yu, K.-H.; Kim, B.C. Amperometric glucose biosensor based on glucose oxidase immobilized over chitosan nanopartoicles from glaudius of Uroteuthis duvauceli. Sens. Actuators B Chem. 2015, 215, 536–543. [Google Scholar] [CrossRef]
  56. Kim, H.S.; Lee, J.S.; Kim, M.I. Poly-γ-glutamic Acid/Chitosan hydrogel nanoparticles entrapping glucose oxidase and magnetic nanoparticles for glucose biosensing. J. Nanosci. Nanotechnol. 2020, 20, 5333–5337. [Google Scholar] [CrossRef]
  57. Kalam, M.A. Development of chitosan nanoparticles coated with hyaluronic acid for topical ocular delivery of dexamethasone. Int. J. Biol. Macromol. 2016, 89, 127–136. [Google Scholar] [CrossRef]
  58. Obaje, E.A.; Cummins, G.; Schulze, H.; Mahmood, S.; Desmulliez, M.P.; Bachmann, T.T. Carbon screen-printed electrodes on ceramic substrates for label-free molecular detection of antibiotic resistance. J. Interdiscip. Nanomed. 2016, 1, 93–109. [Google Scholar] [CrossRef]
  59. Rampino, A.; Borgogna, M.; Blasi, P.; Bellich, B.; Cesaro, A. Chitosan nanoparticles: Preparation, size evolution, and stability. Int. J. Pharm. 2013, 455, 219–228. [Google Scholar] [CrossRef]
  60. Oldham, K.B. Analytical expression for the reversible Randles-Sevcik function. J. Electroanal. Chem. Interfacial Electrochem. 1979, 105, 373–375. [Google Scholar] [CrossRef]
  61. Lavagnini, I.; Antiochia, R.; Magno, F. An extended method for the practical evaluation of the standard rate constant from cyclic voltammetric data. Electroanalysis 2004, 16, 505–506. [Google Scholar] [CrossRef]
  62. Lavagnini, I.; Antiochia, R.; Magno, F. A calibration-base method for the evaluation of the detection limit of an electrochemical biosensor. Electroanalysis 2007, 19, 1127–1230. [Google Scholar] [CrossRef]
  63. Marroquin, J.B.; Rhee, K.Y.; Park, S.J. Chitosan nanocomposite films: Enhanced electrical conductivity, thernmal stability, and mechanical properties. Carbohydr. Polym. 2013, 92, 1783–1791. [Google Scholar] [CrossRef] [PubMed]
  64. Zhang, Z.; Wang, B.Q.; Xu, B.; Cheng, G.J.; Dong, S.J. Amperometric quantification of polar organic solvents based on a tyrosinase biosensor. Anal. Chem. 2000, 72, 3455–3460. [Google Scholar] [CrossRef] [PubMed]
  65. Cosnier, S.; Popescu, I.C. Poly(amphilic pyrrole)-tyrosinase-peroxidase electrode for amplified flow injection-amperometric detection of phenol. Anal. Chim. Acta 1996, 319, 145–151. [Google Scholar] [CrossRef]
  66. Kamin, R.A.; Wilson, G.S. Rotating ring-disk enzyme electrode for biocatalysis kinetic studies and characterization of the immobilized enzyme layer. Anal. Chem. 1980, 52, 1198–1205. [Google Scholar] [CrossRef]
  67. Espin, J.C.; Varon, R.; Fenoll, L.G.; Gilabert, M.A.; Garcia-Ruiz, P.A.; Tudela, J.; Garcia-Canovas, F. Kinetic characterization of the substrate specificity and mechanism of mushroom tyrosinase. Eur. J. Biochem. 2000, 267, 1270–1279. [Google Scholar] [CrossRef] [Green Version]
  68. Zou, Y.S.; Lou, D.; Dou, K.; He, L.L.; Dong, Y.H.; Wang, S.L. Amperometric tyrosinase biosensor based on boron-doped nanocrystalline diamond film electrode for the detection of phenolic compounds. J. Solid State Electrochem. 2016, 20, 47–54. [Google Scholar] [CrossRef]
  69. Rahman, S.F.; Min, K.; Park, S.H.; Park, J.H.; Yoo, J.C.; Park, D.H. Highly sensitive and selective dopamine detection by an amperometric biosensor based on tyrosinase/MWNT/GCE. Korean J. Chem. Eng. 2016, 33, 3442–3447. [Google Scholar] [CrossRef]
  70. Li, H.Q.; Hu, X.; Zhu, H.M.; Zang, Y.; Xue, H.G. Amperometric phenol biosensor based on a new immobilization matrix: Polypyrrole nanotubes derived from methyl orange as dopant. Int. J. Electrochem. Sci. 2017, 12, 6714–6728. [Google Scholar] [CrossRef]
  71. Sethuraman, V.; Muthuraja, P.; Anandha Raj, J.; Manisankar, P. A highly sensitive electrochemical biosensor for catechol using conducting polymer reduced graphene oxide–metal oxide enzyme modified electrode. Biosens. Bioelectron. 2016, 84, 112–119. [Google Scholar] [CrossRef] [PubMed]
  72. Vicentini, F.C.; Garcia, L.L.C.; Figueiredo-Filho, L.C.S.; Janegitz, B.C.; Fatibello-Filho, O. A biosensor based on gold nanoparticles, dihexadecylphosphate, and tyrosinase for the determination of catechol in natural water. Enzym. Microb. Technol. 2016, 84, 17–23. [Google Scholar] [CrossRef] [PubMed]
  73. Araque, E.; Villalonga, R.; Gamella, M.; Martínez-Ruiz, P.; Reviejo, J.; Pingarrón, J.M. Crumpled reduced graphene oxide–polyamidoamine dendrimer hybrid nanoparticles for the preparation of an electrochemical biosensor. J. Mater. Chem. B 2013, 1, 2289–2296. [Google Scholar] [CrossRef] [PubMed]
  74. Borisova, B.; Ramos, J.; Díez, P.; Sánchez, A.; Parrado, C.; Araque, E.; Villalonga, R.; Pingarrón, J.M. A Layer-by-Layer biosensing architecture based on polyamidoamine dendrimer and carboxymethylcellulose-modified graphene oxide. Electroanalysis 2015, 27, 2131–2138. [Google Scholar] [CrossRef]
  75. Lete, C.; Lupu, S.; Lakard, B.; Hihn, J.Y.; del Campo, F.J. Multi-analyte determination of dopamine and catechol at single-walled carbon nanotubes—Conducting polymer—Tyrosinase based electrochemical biosensors. J. Electroanal. Chem. 2015, 744, 53–61. [Google Scholar] [CrossRef]
  76. Min, K.; Yoo, Y.J. Amperometric detection of dopamine based on tyrosinase-SWNTs-Ppy composite electrode. Talanta 2009, 80, 1007–1011. [Google Scholar] [CrossRef]
  77. Meloni, F.; Spychalska, K.; Zając, D.; Pilo, M.I.; Zucca, A.; Cabaj, J. Application of a Thiadiazole-derivative in a Tyrosinase-based Amperometric Biosensor for Epinephrine Detection. Electroanalysis 2021, 33, 1639–1645. [Google Scholar] [CrossRef]
Figure 1. FE-SEM images and EDX spectra for GPH/SPE (Panel (A,B)) and ChitNPs/GPH/SPE (Panel (C,D)). Experimental conditions: magnification, 20,000×; voltage, 1.50 kV.
Figure 1. FE-SEM images and EDX spectra for GPH/SPE (Panel (A,B)) and ChitNPs/GPH/SPE (Panel (C,D)). Experimental conditions: magnification, 20,000×; voltage, 1.50 kV.
Biosensors 12 00519 g001
Figure 2. CVs of 50 μM dopamine in PBS (0.1 M, pH 7.2, KCl 0.1 M) on a GPH/SPE (pink line), a ChitNPs/GPH/SPE (black line), a Tyr/Chit/GPH/SPE (green line), a Np-Tyr/ChitNPs/GPH/SPE (red line), and a LbL-Tyr/ChitNPs/GPH/SPE (blue line). Scan rate: 10 mV s−1.
Figure 2. CVs of 50 μM dopamine in PBS (0.1 M, pH 7.2, KCl 0.1 M) on a GPH/SPE (pink line), a ChitNPs/GPH/SPE (black line), a Tyr/Chit/GPH/SPE (green line), a Np-Tyr/ChitNPs/GPH/SPE (red line), and a LbL-Tyr/ChitNPs/GPH/SPE (blue line). Scan rate: 10 mV s−1.
Biosensors 12 00519 g002
Figure 3. Effect of Tyr concentration on the cathodic peak current of the Tyr/ChitNPs/GPH/SPE in PBS (0.1 M, pH 7.2, KCl 0.1 M) with 50 μM DA.
Figure 3. Effect of Tyr concentration on the cathodic peak current of the Tyr/ChitNPs/GPH/SPE in PBS (0.1 M, pH 7.2, KCl 0.1 M) with 50 μM DA.
Biosensors 12 00519 g003
Scheme 1. Schematic representation of the Tyr/EDC-NHS/ChitNPs/GPH-SPE electrochemical platform and the CA biosensor mechanism.
Scheme 1. Schematic representation of the Tyr/EDC-NHS/ChitNPs/GPH-SPE electrochemical platform and the CA biosensor mechanism.
Biosensors 12 00519 sch001
Figure 4. CVs of 50 μM DA in PBS (0.1 M, pH 7.2, KCl 0.1 M) on a ChitNPs/GPH/SPE (black line), a LbL-Tyr (0.5 mg/mL)/ChitNPs/GPH/SPE (blue line), and a LbL-Tyr (0.5 mg/mL)/EDC-NHS/ChitNPs/GPH/SPE (red line). Scan rate: 10 mV s−1.
Figure 4. CVs of 50 μM DA in PBS (0.1 M, pH 7.2, KCl 0.1 M) on a ChitNPs/GPH/SPE (black line), a LbL-Tyr (0.5 mg/mL)/ChitNPs/GPH/SPE (blue line), and a LbL-Tyr (0.5 mg/mL)/EDC-NHS/ChitNPs/GPH/SPE (red line). Scan rate: 10 mV s−1.
Biosensors 12 00519 g004
Figure 5. Nyquist plots of GPH/SPE (green), ChitNPs/GPH/SPE (black), Tyr/ChitNPs/GPH/SPE (blue), and Tyr/EDC-NHS/ChitNPs/GPH/SPE (red) measured in the presence of 5 mM [Fe(CN)6]3−/4− containing 0.1 M KCl solution. The inset shows the equivalent circuit used for fitting the experimental data.
Figure 5. Nyquist plots of GPH/SPE (green), ChitNPs/GPH/SPE (black), Tyr/ChitNPs/GPH/SPE (blue), and Tyr/EDC-NHS/ChitNPs/GPH/SPE (red) measured in the presence of 5 mM [Fe(CN)6]3−/4− containing 0.1 M KCl solution. The inset shows the equivalent circuit used for fitting the experimental data.
Biosensors 12 00519 g005
Figure 6. Typical chronoamperometric curves of Tyr/ChitNPs/GPH/SPE (Panel A) and Tyr/EDC-NHS/ChitNPs/GPH/SPE (Panel B) upon successive additions of different concentrations of DA to the stirred PBS (0.1 M, pH 7.2, KCl 0.1 M) (applied potential: 0 V vs. Ag/AgCl) and the relative calibration curves of Tyr/ChitNPs/GPH/SPE (Panel C, red curve) and Tyr/EDC-NHS/ChitNPs/GPH/SPE (Panel C, black curve) as a function of DA concentration. Inset: the magnification of the linear range.
Figure 6. Typical chronoamperometric curves of Tyr/ChitNPs/GPH/SPE (Panel A) and Tyr/EDC-NHS/ChitNPs/GPH/SPE (Panel B) upon successive additions of different concentrations of DA to the stirred PBS (0.1 M, pH 7.2, KCl 0.1 M) (applied potential: 0 V vs. Ag/AgCl) and the relative calibration curves of Tyr/ChitNPs/GPH/SPE (Panel C, red curve) and Tyr/EDC-NHS/ChitNPs/GPH/SPE (Panel C, black curve) as a function of DA concentration. Inset: the magnification of the linear range.
Biosensors 12 00519 g006
Figure 7. Calibration curves of Tyr/EDC-NHS/ChitNPs/GPH/SPE as a function of the concentration of catecholamine: DA (black curve), EP (blue curve), and NEP (green curve). Inset: the magnification of the linear range.
Figure 7. Calibration curves of Tyr/EDC-NHS/ChitNPs/GPH/SPE as a function of the concentration of catecholamine: DA (black curve), EP (blue curve), and NEP (green curve). Inset: the magnification of the linear range.
Biosensors 12 00519 g007
Figure 8. DPVs of 20 μM DA in the presence of 2 mM AA and 2 mM UA with the Tyr/EDC-NHS/ChitNPs/GPH/SPE biosensor.
Figure 8. DPVs of 20 μM DA in the presence of 2 mM AA and 2 mM UA with the Tyr/EDC-NHS/ChitNPs/GPH/SPE biosensor.
Biosensors 12 00519 g008
Table 1. Comparison of the kinetic and analytical characteristics of DA biosensors with different electrochemical platforms.
Table 1. Comparison of the kinetic and analytical characteristics of DA biosensors with different electrochemical platforms.
Modified
SPE
K M app
(μM)
Jmax
(μA cm−2)
Linear Range
(μM)
LOD
(μM)
Sensitivity
(μA μM−1 cm−2)
R2
Tyr/ChitNPs/GPH25.59 ± 3.01311.1 ± 0.844–121.360.2750.994
Tyr/EDC-NHS/ChitNPs/GPH9.51 ± 0.5449.16 ± 0.0090.5–50.170.5830.996
Table 2. Comparison of the kinetic and analytical characteristics of Tyr/EDC-NHS/ChitNPs/GPH/SPE biosensors with different catecholamines.
Table 2. Comparison of the kinetic and analytical characteristics of Tyr/EDC-NHS/ChitNPs/GPH/SPE biosensors with different catecholamines.
Catecholamine K M app
(μM)
Jmax
(μA cm−2)
Linear Range
(μM)
LOD
(μM)
Sensitivity
(μA μM−1 cm−2)
R2
DA9.51 ± 0.5449.16 ± 0.0090.5–50.170.5830.996
EP25.46 ± 1.0954.95 ± 0.0992–110.710.1250.998
NEP20.83 ± 2.28074.74 ± 0.3231–110.400.1320.995
Table 3. Comparison of Tyr-based biosensors reported in the literature.
Table 3. Comparison of Tyr-based biosensors reported in the literature.
Catecholamine/CatecholBiosensor PlatformLinear Range (μM)LOD (μM)Ref.
catecholTyr/BOND/film electrode5–1202.35[68]
DATyr/MWNT/GCE50–100050[69]
catecholPPy-NTs/PPO0.1–350.0012[70]
catecholPEDOT-rGO-Fe2O3-PPO0.04–620.007[71]
catecholTyr-AuNPs-DHP/GCE2.5–950.17[72]
catecholTyr-PAMAM-Sil-rGO/GCE0.01–220.006[73]
catecholTyr/PAMAM/GO-CMC/GCE0.002–0.40.0009[74]
catecholTyr/GO/GA/GCE0.05–500.03[31]
NEPTyr-CDs-CA/Au-E1–2000.196[29]
DATyr-CoP-Au2–300.43[27]
DAIDE/PEDOT-CNT-Tyr-GAD-HFR100–50011[75]
DATyr-SWNTs-Ppy5–505[76]
EP
DA
Tyr/TBT-film/Au
Tyr/EDC-NHS/ChitNPs/GPH
0.1–50
0.5–5
0.06
0.17
[77]
this work
Table 4. Determination of DA in urine samples (n = 3) using the calibration curve from Figure 6.
Table 4. Determination of DA in urine samples (n = 3) using the calibration curve from Figure 6.
SampleAdded (μM)Found (μM)Recovery
(%)
RSD
(n = 3) (%)
Spectrophotometric Method (μM)
Sample 154.8964.25.10
Sample 254.9982.15.25
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gigli, V.; Tortolini, C.; Capecchi, E.; Angeloni, A.; Lenzi, A.; Antiochia, R. Novel Amperometric Biosensor Based on Tyrosinase/Chitosan Nanoparticles for Sensitive and Interference-Free Detection of Total Catecholamine. Biosensors 2022, 12, 519. https://doi.org/10.3390/bios12070519

AMA Style

Gigli V, Tortolini C, Capecchi E, Angeloni A, Lenzi A, Antiochia R. Novel Amperometric Biosensor Based on Tyrosinase/Chitosan Nanoparticles for Sensitive and Interference-Free Detection of Total Catecholamine. Biosensors. 2022; 12(7):519. https://doi.org/10.3390/bios12070519

Chicago/Turabian Style

Gigli, Valeria, Cristina Tortolini, Eliana Capecchi, Antonio Angeloni, Andrea Lenzi, and Riccarda Antiochia. 2022. "Novel Amperometric Biosensor Based on Tyrosinase/Chitosan Nanoparticles for Sensitive and Interference-Free Detection of Total Catecholamine" Biosensors 12, no. 7: 519. https://doi.org/10.3390/bios12070519

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