Next Article in Journal
SO2 Detection over a Wide Range of Concentrations: An Exploration on MOX-Based Gas Sensors
Previous Article in Journal
Preparation, Characterization and Electrochemical Response of Nanostructured TiAlV with Potentiostatically Deposited IrOx as a pH Sensor for Rapid Detection of Inflammation
Previous Article in Special Issue
Highly Sensitive and Selective MEMS Gas Sensor Based on WO3/Al2O3/Graphite for 2-Chloroethyl Ethyl Sulfide (2-CEES) Detection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

CoNiTe2 Nanomaterials as an Efficient Non-Enzymatic Electrochemical Sensing Platform for Detecting Dopamine

1
Department of Chemical Engineering, National Chung Hsing University, Taichung 40227, Taiwan
2
Department of Chemical Engineering, National United University, Miaoli 360302, Taiwan
3
Pesticide Analysis Center, National United University, Miaoli 360302, Taiwan
*
Authors to whom correspondence should be addressed.
Chemosensors 2024, 12(6), 110; https://doi.org/10.3390/chemosensors12060110
Submission received: 17 May 2024 / Revised: 7 June 2024 / Accepted: 11 June 2024 / Published: 13 June 2024
(This article belongs to the Special Issue Nanomaterial-Based Chemosensors and Biosensors for Smart Sensing)

Abstract

:
Dopamine (DA) is an important catecholamine neurotransmitter in the mammalian central nervous system that affects many physiological functions. Hence, a highly sensitive and selective sensing platform is necessary for quantification of DA in the human body. In this study, ternary transition metal tellurides of CoNiTe2 were successfully synthesized using the hydrothermal method. The proposed CoNiTe2 nanomaterials were dispersed well in Nafion to form a well-dispersed suspension and, when dropped on a glassy carbon electrode (GCE) as the working electrode (CoNiTe2/Nafion/GCE) for electrochemical non-enzymatic DA sensing, displayed excellent electrocatalytic activity for dopamine electrooxidation. The morphology and physical/chemical properties of CoNiTe2 nanomaterials were characterized using field emission scanning electron microscopy (FESEM), transmission electron microscopy (TEM), X-ray diffraction (XRD), and X-ray photoelectron spectroscopy (XPS). In order to obtain the best electrochemical response to DA from the fabricated CoNiTe2/Nafion/GCE, the experimental conditions of electrochemical sensing, including the CoNiTe2 loading amounts and pH values of the phosphate buffer solution (PBS), were explored to achieve the best electrochemical sensing performance. Under optimal conditions (2 mg of CoNiTe2 and pH 6.0 of PBS), the fabricated CoNiTe2/Nafion/GCE showed excellent electrocatalytic activity of DA electrooxidation. The CoNiTe2/Nafion/GCE sensing platform demonstrated excellent electrochemical performance owing to the optimal structural and electronic characteristics originating from the synergistic interactions of bimetallic Co and Ni, the low electronegativity of Te atoms, and the unique morphology of the CoNiTe2 nanorod. It exhibited a wide linear range from 0.05 to 100 μM, a high sensitivity of 1.2880 µA µM−1 cm−2, and a low limit of detection of 0.0380 µM, as well as acceptable selectivity for DA sensing. Therefore, the proposed CoNiTe2/Nafion/GCE could be considered a promising electrode material for electrochemical non-enzymatic DA sensing.

1. Introduction

Carlsson et al. (1958) reported that dopamine (DA) is a key neurotransmitter in the brain and is used as an immediate precursor in the biosynthesis of noradrenaline [1]. Current data reveal that DA is the predominant neurotransmitter expressed in the mammalian central nervous system (accounting for 80% of the catecholamine content in the brain) [2]. The amount of DA released correlates exclusively to diverse brain functions and is expressed specifically in the ventral tegmental area (VTA) of the midbrain, the substantia nigra pars compacta, and the hypothalamic arcuate nucleus of the human brain [3]. VTA dopaminergic activity controls natural motivation, reward prediction, and contextual learning [4,5]. Parkinson’s disease (PD), one of the most common neurodegenerative disorders, is linked to abnormal dopamine release in the substantia nigra pars compacta. PD is diagnosed by characteristic motor tremors [6,7]. The hypothalamic arcuate nucleus’ dopaminergic neurons are essential in regulating prolactin release. Prolactin is a protein hormone synthesized and secreted by lactotropic cells in the anterior pituitary gland and is involved in the prolactin homeostasis of the body [8,9].
Unquestionably, finding a precise and efficient method to quickly detect DA level deviations, thereby mitigating the societal and economic burdens associated with neurological disorders, is urgent. So far, analytical methods such as fluorescence [10], colorimetry [11], chemiluminescence [12], high-performance liquid chromatography (HPLC) [13], capillary electrophoresis (CE) [14], and electrochemistry [15,16] have been used to detect DA. Electrochemical methods stand out among the conventional analytical techniques because they are affordable, have a short analysis time, give an immediate response, and are user-friendly [17,18]. Electrochemical analytical methods detect DA using two sensing platforms: enzymatic and non-enzymatic. Although enzymatic sensors demonstrate high sensitivity and specificity, they are expensive, are complex to produce, have limited reproducibility, and are susceptible to environmental factors such as pH and temperature [19]. An electrochemical non-enzymatic DA sensor needs to be developed to overcome these drawbacks.
Previous reports reveal that electrochemical DA sensors based on transition metal chalcogenides (including transition metal sulfides (TMSs) [20], transition metal selenides (TMSs) [21], and transition metal tellurides (TMTs) [22]) have been successfully used in the development of electrode materials. Transition metal chalcogenides are a promising choice for electrode materials because of their excellent conductivity, abundant active sites, simplified synthesis procedures, and low preparation costs [23]. Furthermore, the electronegativity of Te (2.1) is low compared to S (2.58) and Se (2.55), resulting in weaker bonds between transition metals and Te. This facilitates electrochemical redox reactions [24]. These characteristics make TMTs suitable for extensive applications in the electrochemical field such as in supercapacitors [25], water splitting [26], and CO2 reduction [27]. From the above-mentioned reports, TMTs have attracted significant attention in the electrochemical field due to the synergic effect caused by multiple redox-active transition metals, the low electronegativity of Te, and their unique structural features. In TMTs, the electronegativity of Te is lower than S and Se, which would tend to weaken covalent bonding between the transition metal and Te. Electronic structure tuning in transition metal chalcogenides should facilitate ion insertion/extraction and electron transport to give an excellent electrochemical performance. Nevertheless, the existing literature provides limited information on TMTs’ use as electrochemical biosensors, especially in DA sensing.
In this study, we prepared CoNiTe2 ternary transition metal telluride nanomaterials using the hydrothermal method. We fabricated a CoNiTe2/Nafion/glassy carbon electrode (GCE) sensing platform by drop-casting the prepared CoNiTe2/Nafion nanocomposite on GCE. The CoNiTe2/Nafion nanocomposite catalyzed DA electrochemical redox to deliver remarkable electrochemical performance through the synergistic effects of bimetallic Co and Ni, the relatively low electronegativity of Te, and the unique features of nanorod morphology.

2. Materials and Methods

2.1. Reagents

Cobalt (II) chloride hexahydrate (CoCl2·6H2O), nickel (II) chloride hexahydrate (NiCl2·6H2O), hydrazine monohydrate (N2H4·H2O), and ethylenediamine were obtained from Alfa Aesar (Ward Hill, MA, USA). Tellurium powder (~200 mesh), Nafion solution (5 wt % in mixture of lower aliphatic alcohols and water), sodium phosphate dibasic (Na2HPO4), sodium phosphate monobasic (NaH2PO4), dopamine hydrochloride (DA), uric acid (UA), and L-ascorbic acid (AA) were purchased from Sigma-Aldrich (St. Louis, MO, USA). Anhydrous ethanol (C2H5OH, 99.9%) was purchased from J.T. Baker (Phillipsburg, NJ, USA). The deionized water (DI water) was produced using a Milli-Q water purification system from Millipore Co. (Bedford, MA, USA). All chemicals were analytical grade and were used as received without further purification.

2.2. Synthesis of CoNiTe2 Nanomaterials

The CoNiTe2 transition metal tellurides were successfully synthesized via a facile hydrothermal method. As is typical, 0.409 g CoCl2·6H2O and 0.411 g NiCl2·6H2O were precisely weighed and fully dissolved in 30 mL DI water by stirring for 30 min to form solution (A). At the same time, 0.235 g Te powder was dispersed into 8 mL ethylenediamine and stirred for 30 min to form solution (B). In the next step, solution (B) was added dropwise to solution (A) and stirred for 30 min to make sure all the reactants were totally dissolved. After that, 8 mL N2H4·H2O was selected as the reducing agent and added to the mixture. The mixture was allowed to stir until a black precipitate appeared. Subsequently, the obtained black precipitate containing the Ni–Co telluride precursor was transferred into a 100 mL Teflon-lined stainless autoclave and heated to 180 °C for 12 h in an oven. Under mild hydrothermal conditions, the as-fabricated Ni–Co telluride precursor went through the amorphous–crystalline phase transformation induced by incorporated metal ions. Finally, the resulting CoNiTe2 nanomaterials were cleaned by with DI water and anhydrous ethanol in a centrifuge 3 times, and then dried in the oven overnight. Then, the dried CoNiTe2 nanomaterials were collected for subsequent characterization. A schematic illustration of the synthesis of the CoNiTe2 nanomaterials is shown in Figure 1.

2.3. Fabrication of CoNiTe2/Nafion/GCE Working Electrode

To prepare the glassy carbon electrode (GCE, diameter 3 mm, Tokai Carbon, Tokyo, Japan) for use, a thorough polishing procedure was performed by using 0.3 μm and 0.05 μm alumina slurries. Subsequently, the electrode was subjected to a rapid ultrasonic cleaning process using deionized (DI) water and anhydrous ethanol, followed by drying in a 70 °C oven. Next, 2 mg CoNiTe2 nanomaterials were weighed and dispersed in 1 mL 0.5 wt % Nafion solution through ultrasonication for 30 min to form a well-distributed suspension. To fabricate the working electrode, 6 μL of the suspension was dropped on the precleaned GCE (bare GCE) and dried in the oven at 60 °C for 20 min. In this way, CoNiTe2/Nafion/GCE was obtained for the following electrochemical measurements.

2.4. Characterizations

The morphology was characterized using field emission scanning electron microscopy (FESEM, JSM-7800F, JEOL, Akishima, Japan) and transmission electron microscopy (TEM, JEM-2100F, JEOL, Akishima, Japan). The crystal phase was analyzed by using X-ray diffraction (XRD) (D8 Discover X-ray diffractometer with Cu Kα radiation (Bruker, Karlsruhe, Germany). The chemical structure and composition were determined by using X-ray photoelectron spectroscopy (XPS, PHI-5000 Versaprobe, ULVAC-PHI, Chigasaki, Japan). Electrochemical measurements were performed by using an electrochemical analyzer (Autolab, model PGSTAT 30, Eco Chemie, Utrecht, The Netherlands) in a three-electrode system (the as-prepared samples modified with GCE were used as the working electrode, along with a platinum wire counter electrode and an Ag/AgCl (3 M KCl) reference electrode). All electrochemical measurements were conducted in 0.1 M phosphate-buffered saline (PBS) as the supporting electrolyte in the absence or presence of DA. Cyclic voltammetry (CV) and differential pulse voltammetry (DPV) measurements were taken in the working potential window of −0.2~1.0 V and 0~0.5 V. The DPV operational parameters (in 0.1 M PBS at pH 6.0) were optimized as follows: modulation time of 0.050 s, modulation amplitude of 0.050 V, pulse width of 0.050 s, internal time of 0.200 s, and step potential of 0.004 V.

3. Results and Discussion

The FESEM image (Figure 2a) revealed that the CoNiTe2 surface is composed of a rod-like structure (approximately 700 nm in length) with numerous short nanorods (approximately several nanometers in length) growing on it. The short nanorods provide abundant electrochemical active sites and this boosts electronic/ionic transport during the subsequent electrochemical measurements. The TEM image (Figure 2b) revealed that the CoNiTe2 nanomaterial’s inner structure concurred with the surface morphology observed by FESEM. Figure 3a–d show the scanning TEM (STEM) images and their corresponding EDS mapping for CoNiTe2. These affirm that the CoNiTe2 nanorod structure comprises well-distributed Co, Ni, and Te elements. The EDS spectrum (Figure 3e) of the CoNiTe2 nanomaterials further revealed that the atomic ratio of Co, Ni, and Te is approximately 0.2:0.2:0.5, which is close to the theoretical atomic ratio of CoNiTe2. The CoNiTe2 nanomaterials were then characterized by X-ray diffraction. The standard XRD pattern of CoNiTe2 (JCPDS No. 65-8961) [28] was obtained.
The XRD pattern identified the structure and phase composition of the CoNiTe2 nanomaterials based on the standard patterns in the crystalline phase of CoNiTe2 (JCPDS No. 65-8961) [28], as shown in Figure 4. In the XRD pattern of the CoNiTe2 nanomaterials, the diffraction peaks located at 2θ equaling 31.1°, 43.2°, 45.9°, 56.5°, and 58.9° are attributed to the (101), (102), (110), (201), and (103) planes, respectively (indexed to the crystalline phase of CoNiTe2). These characterizations confirm that the CoNiTe2 ternary transition metal tellurides were successfully synthesized via a facile hydrothermal method.
The surface elemental composition and valence states of the CoNiTe2 nanomaterials were characterized by XPS, as shown in Figure 5. Figure 5a shows the high-resolution Co 2p XPS spectra of the CoNiTe2 nanomaterials. Two spin-orbit doublets corresponding to Co 2p3/2 and Co 2p1/2 were found. Co2+ was observed at 780.8 eV and 797.2 eV (binding energies), while Co3+ was observed at 778.8 eV and 796 eV. In addition to the two spin-orbit doublets, shake-up satellites (marked as Sat.) were observed adjacent to each doublet, located at 785.4 eV and 802.6 eV, respectively. Figure 5b shows the high-resolution Ni 2p XPS spectra of the CoNiTe2 nanomaterials. Two spin-orbit doublets were ascribed to Ni 2p3/2 and Ni 2p1/2. Ni3+ was located at binding energies 855.4 eV and 873.4 eV, while Ni2+ was observed at binding energies 852.4 eV and 870.2 eV. Likewise, two shake-up satellites (marked as Sat.) were observed adjacent to the two spin-orbit doublets, which were situated at 861.4 eV and 879.0 eV, respectively. Figure 5c shows the high-resolution Te 3d XPS spectra of the CoNiTe2 nanomaterials. The spin-orbit doublets of Te 3d5/2 and Te 3d3/2 appeared at 573 eV and 583.4 eV, respectively, and are attributable to Te2−. The other two peaks at 576 eV and 586.4 eV are shake-up satellites (marked as Sat.) [29].
The electrochemical characteristics of CoNiTe2 nanomaterials for DA detection were performed by cyclic voltammetry (CV). Figure 6 shows the CV curves of bare GCE (black lines), Nafion/GCE (blue lines), and CoNiTe2/Nafion/GCE (red lines) in 0.1 M PBS (pH 7.0) in the absence (dashed lines) and presence (solid lines) of 1 mM DA at a 50 mV s−1 scan rate within the 0–1.0 V potential window. In the absence of DA, all electrodes show no redox peaks due to the absence of redox-active species. With the addition of 1 mM DA, all electrodes exhibited well-defined peaks (especially the CoNiTe2/Nafion/GCE electrode) related to the electrochemical DA redox reaction mechanism. The possible electrochemical DA redox reaction mechanism is described in the previous literature [30]. During the electrochemical oxidation process, DA was oxidized to dopaminoquinone (DAQ), and during the electrochemical reduction process, DAQ was reduced to DA. The electrochemical DA redox reaction is a reversible process involving two electron/two proton transfers. Notably, CoNiTe2/Nafion/GCE possesses the highest electrocatalytic activity in DA sensing. These results demonstrate that the synergistic interaction of bimetallic Co and Ni, the low electronegativity of Te, and the unique morphology of CoNiTe2 nanorods improved electrochemical performance. According to a previous report [31], bimetallic Ni–Co transition metal chalcogenides generally exhibited satisfactory electrochemical performance. In Ni–Co tellurides, the electronegativity of Te is lower than those of S and Se. This results in weakened covalent bonding between the transition metal and Te, further accelerating electron transfer and modulating electronic structure in the electrochemical process. This enhanced electrochemical performance.
To obtain CoNiTe2 nanomaterials with optimal electrochemical performance for DA sensing, the CoNiTe2 loading amounts and pH values of the phosphate buffer solution were varied to understand electrochemical parameter optimization. To do this, 1~4 mg of CoNiTe2 was weighed and dispersed into 1 mL 0.5 wt % Nafion to form a homogenous dispersion, then immobilized on the surface of GCE by a drop-casting method. Figure 7a shows the CV curves of CoNiTe2/Nafion/GCE at different CoNiTe2 loading amounts in 0.1 M PBS (pH 7.0) in the presence of 1 mM DA at a scan rate of 50 mV s−1. It was observed that the oxidation peak current dramatically increased as the CoNiTe2 loading amount varied from 1 to 2 mg. However, when the CoNiTe2 loading amount exceeded 2 mg, the oxidation peak current decreased, indicating that the presence of the excess CoNiTe2 hindered the mass transfer of dopamine [32]. Figure 7b displays the CoNiTe2/Nafion/GCE CV curves at different phosphate buffer solution pH values (from pH 5 to 8) in the presence of 1 mM DA. The oxidation peak potential (Epa) decreased with increasing pH value from 5 to 8, revealing that DA electrochemical behavior is associated with the pH dependence of interfacial electron–proton transfer. The corresponding linear relationship between oxidation peak potential and pH was calculated as follows: Epa (V) = 0.846–0.059 pH (R2 = 0.9771) (see the inset of Figure 7b). The results indicate a 0.059 V/pH linear slope. This is close to the theoretical value, implying that the electrochemical DA redox reaction mechanism involves two electron/two proton transfer steps, as governed by the Nernst equation [33]. Notably, the oxidation peak current increased slightly as the pH increased from 5 to 6, and then decreased rapidly from pH 6 to 8. The maximum DA oxidation peak current was at pH 6. Consequently, 2 mg of CoNiTe2 (loading amount) and a phosphate buffer solution pH of 6.0 (pH value) were selected as the optimal parameters for DA sensing.
To further examine the electrochemical behavior of the CoNiTe2 nanomaterials, CoNiTe2/Nafion/GCE was studied through CV performed in 0.1 M PBS (pH 6.0) in the presence of 1 mM DA at 50 to 300 mV s−1 scan rates (Figure 8a). It was observed that the redox peak current increased with increasing scan rate. Figure 8b reveals that both the oxidation peak current (Ipa) and reduction peak current (Ipc) were linear with the square root of the scan rate (v1/2). The linear regression equation is expressed as Ipa (μA) = 34.8800 + 3.2760 v1/2 ((mV s−1)1/2) (R2 = 0.9856) and Ipc (μA) = −22.7890–5.0160 v1/2 ((mV s−1)1/2) (R2 = 0.9897). These results establish that the electrochemical DA redox reaction is controlled by diffusion according to the Randles–Sevcik equation [34].
We measured the electrochemical performance of CoNiTe2/Nafion/GCE for DA electrooxidation using differential pulse voltammetry (DPV) in 0.1 M PBS (pH 6.0) with increasing DA concentrations (0~100 μM) under optimal experimental parameters to evaluate the feasibility of the proposed electrochemical non-enzymatic DA sensing. Figure 9a shows the CoNiTe2/Nafion/GCE DPV responses against various DA concentrations from 0 to 100 μM. It was observed that the DPV response increased with increasing DA concentration. The DPV responses corresponding to DA concentrations ranging from 0.05 to 100 μM were plotted to obtain the corresponding calibration curve (Figure 9b). The linear regression equation for the oxidation peak current (Ipa) against the DA concentration is expressed as Ipa (μA) = 1.0030 + 0.0910 C (μM). The DA sensor calibration curve was linear from 0.05 to 100 μM (R2 = 0.9928), and the slope and intercept were 0.0910 µA µM−1 and 1.0030 μA, respectively. The sensitivity (according to the geometric area of GCE (0.0707 cm2) and the slope of the calibration curve), limit of detection (LOD) based on 3 Sb/m, and limit of quantification (LOQ) based on 10 Sb/m (Sb is the blank signal standard deviation for n = 3, and m is the calibration plot slope) were estimated as 1.2880 µA µM−1 cm−2, 0.0380 µM, and 0.1270 mM, respectively. Table 1 summarizes previous reports regarding electrochemical non-enzymatic DA sensors based on the different transition metal chalcogenides [21,35,36,37]. It is noted that the CoNiTe2 electrode material is comparable to the other electrochemical non-enzymatic DA sensors based on the different transition metal chalcogenides.
In a physiological environment, the electrochemical similarities between dopamine (DA), uric acid (UA), and ascorbic acid (AA) biomolecules can cause overlapping detection signals, potentially impacting DA sensing accuracy and reliability. Therefore, it is necessary to perform an interference study to evaluate the sensing ability and distinguish the interfering species. This ensures the accurate determination of the target molecules. To evaluate the selectivity of the CoNiTe2 nanomaterials in the presence of the interfering molecules for DA sensing, Figure 10 displays the interference effect results for CoNiTe2/Nafion/GCE in 0.1 M PBS (pH 6.0) in the presence of 200 μM AA and 10 μM UA (both are reasonable concentrations in the human body) and various DA concentrations (0.05, 0.5, 1, 3, 5, 7.5, and 10 μM). As observed, the DPV responses of the interfering molecules AA and UA were found at 0.15 V and 0.35 V, respectively. However, the DPV response of DA was observed at about 0.22 V, and its response increased with increasing DA concentration. In the experimental results, no significant interference effect was observed for DA determination in the presence of AA and UA. The linear regression equation for the oxidation peak current (Ipa) against the DA concentration in the presence of AA and UA (interference effect) is expressed as Ipa (μA) = 2.3262 + 0.1401 C (μM). The calibration curve was linear from 0.05 to 10 μM (R2 = 0.9946), and the slope and intercept were 0.1401 µA µM−1 and 2.3262 μA, respectively. For interference effect evaluation, a comparison of slope differences in the solvent and interference addition calibration curves showed that there was a small increase in slope in the presence of interfering molecules, indicating a slight interference effect. To minimize interference effects, a quantitative analysis was further performed using matrix-matched calibration curves to reduce interference-induced effects. As discussed above, it can be deduced that CoNiTe2/Nafion/GCE has acceptable selectivity at DA concentrations ranging from 0.05 to 10 μM when AA and UA were also present, demonstrating the excellent anti-interference ability of the CoNiTe2 nanomaterials.

4. Conclusions

In this study, a hydrothermally synthesized CoNiTe2 nanomaterial-fabricated sensing platform (CoNiTe2/Nafion/GCE) demonstrated prominent electrochemical performance for DA sensing owing to its excellent properties caused by bimetallic Co and Ni, the low electronegativity of Te, and its unique structural features (nanorods). In CoNiTe2 nanomaterials, the lower electronegativity of Te leads to a weakened covalent bonding between transition metals and Te, resulting in the modulation of the electronic structure, which could propose an effective way to accelerate electron/ion transportation between the interface of the electroactive materials and electrolyte solution, exhibiting remarkable electrochemical performance for DA sensing. The electrochemical non-enzymatic DA sensor based on a CoNiTe2 nanomaterials sensing platform exhibits excellent performance for the linear range from 0.05 to 100 μM, a sensitivity of 1.2880 µA µM−1 cm−2, and a limit of detection (LOD) of 0.0380 µM. Furthermore, the acceptable selectivity of the presented CoNiTe2/Nafion/GCE in the presence of interfering species makes it a promising DA sensing platform for it practical use.

Author Contributions

Conceptualization, H.-W.C. and Y.-C.T.; methodology, H.-W.C., Z.-Y.W., C.-H.S. and S.-H.Y.; software, Z.-Y.W., C.-H.S. and S.-H.Y.; formal analysis, H.-W.C., Z.-Y.W., C.-H.S. and S.-H.Y.; investigation, H.-W.C., Z.-Y.W., C.-H.S., S.-H.Y. and Y.-C.T.; data curation, H.-W.C., Z.-Y.W., C.-H.S. and S.-H.Y.; writing—original draft preparation, H.-W.C., Z.-Y.W. and Y.-C.T.; writing—review and editing, H.-W.C. and Y.-C.T.; visualization, H.-W.C.; supervision, H.-W.C. and Y.-C.T.; project administration, H.-W.C. and Y.-C.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Science and Technology Council (NSTC), Taichung Veterans General Hospital (TCVGH), and National United University (NUU), Taiwan (NSTC 112-2221-E-239-001-MY3, NSTC 112-2221-E-005-007-MY3, TCVGH-NUU1128901, TCVGH-NUU1138902, and SE113002).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author. The data are not publicly available due to privacy.

Acknowledgments

The authors are grateful to the NSTC and NUU for the financial assistance granted in support of this work. For instrumentation support, we thank the NSTC and the Instrument Center of National Chung Hsing University, Taiwan, for help with FESEM, HRTEM, and XPS measurements (NSTC 112-2740-M-005-001).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Carlsson, A. Evidence for a role of dopamine in extrapyramidal functions. Acta Neuroveg. 1964, 26, 484–493. [Google Scholar] [CrossRef] [PubMed]
  2. Thomas, R.; Pooventhiran, T.; Bakht, M.A.; Alzahrani, A.Y.; Salem, M.A. Study of interaction between different solvents and neurotransmitters dopamine, l-adrenaline, and l-noradrenaline using LED, QTAIM and AIMD. J. Mol. Liq. 2022, 368, 120708. [Google Scholar] [CrossRef]
  3. Gaskill, P.J.; Khoshbouei, H. Dopamine and norepinephrine are embracing their immune side and so should we. Curr. Opin. Neurobiol. 2022, 77, 102626. [Google Scholar] [CrossRef] [PubMed]
  4. Jo, Y.S.; Lee, J.; Mizumori, S.J. Effects of prefrontal cortical inactivation on neural activity in the ventral tegmental area. J. Neurosci. 2013, 33, 8159–8171. [Google Scholar] [CrossRef] [PubMed]
  5. Douma, E.H.; de Kloet, E.R. Stress-induced plasticity and functioning of ventral tegmental dopamine neurons. Neurosci. Biobehav. Rev. 2020, 108, 48–77. [Google Scholar] [CrossRef] [PubMed]
  6. Surmeier, D.J.; Guzman, J.N.; Sanchez-Padilla, J.; Schumacker, P.T. The role of calcium and mitochondrial oxidant stress in the loss of substantia nigra pars compacta dopaminergic neurons in Parkinson’s disease. Neuroscience 2011, 198, 221–231. [Google Scholar] [CrossRef] [PubMed]
  7. Wegrzynowicz, M.; Bar-On, D.; Calo’, L.; Anichtchik, O.; Iovino, M.; Xia, J.; Ryazanov, S.; Leonov, A.; Giese, A.; Dalley, J.W. Depopulation of dense α-synuclein aggregates is associated with rescue of dopamine neuron dysfunction and death in a new Parkinson’s disease model. Acta Neuroveg. 2019, 138, 575–595. [Google Scholar] [CrossRef] [PubMed]
  8. Ayano, G. Dopamine: Receptors, functions, synthesis, pathways, locations and mental disorders: Review of literatures. J. Ment. Disord. Treat. 2016, 2, 2. [Google Scholar] [CrossRef]
  9. Blum, T.; Moreno-Pérez, A.; Pyrski, M.; Bufe, B.; Arifovic, A.; Weissgerber, P.; Freichel, M.; Zufall, F.; Leinders-Zufall, T. Trpc5 deficiency causes hypoprolactinemia and altered function of oscillatory dopamine neurons in the arcuate nucleus. Proc. Natl. Acad. Sci. USA 2019, 116, 15236–15243. [Google Scholar] [CrossRef]
  10. Zhuo, S.; Guan, Y.; Li, H.; Fang, J.; Zhang, P.; Du, J.; Zhu, C. Facile fabrication of fluorescent Fe-doped carbon quantum dots for dopamine sensing and bioimaging application. Analyst 2019, 144, 656–662. [Google Scholar] [CrossRef]
  11. Liu, X.; He, F.; Zhang, F.; Zhang, Z.; Huang, Z.; Liu, J. Dopamine and melamine binding to gold nanoparticles dominates their aptamer-based label-free colorimetric sensing. Anal. Chem. 2020, 92, 9370–9378. [Google Scholar] [CrossRef]
  12. Zhang, L.; Tang, Z.; Dong, Y. Silicon quantum dot involved luminol chemiluminescence and its sensitive detection of dopamine. Anal. Methods 2018, 10, 4129–4135. [Google Scholar] [CrossRef]
  13. De Benedetto, G.E.; Fico, D.; Pennetta, A.; Malitesta, C.; Nicolardi, G.; Lofrumento, D.D.; De Nuccio, F.; La Pesa, V. A rapid and simple method for the determination of 3, 4-dihydroxyphenylacetic acid, norepinephrine, dopamine, and serotonin in mouse brain homogenate by HPLC with fluorimetric detection. J. Pharm. Biomed. Anal. 2014, 98, 266–270. [Google Scholar] [CrossRef] [PubMed]
  14. Zhao, Y.; Zhao, S.; Huang, J.; Ye, F. Quantum dot-enhanced chemiluminescence detection for simultaneous determination of dopamine and epinephrine by capillary electrophoresis. Talanta 2011, 85, 2650–2654. [Google Scholar] [CrossRef] [PubMed]
  15. Shahzad, F.; Iqbal, A.; Zaidi, S.A.; Hwang, S.-W.; Koo, C.M. Nafion-stabilized two-dimensional transition metal carbide (Ti3C2Tx MXene) as a high-performance electrochemical sensor for neurotransmitter. J. Ind. Eng. Chem. 2019, 79, 338–344. [Google Scholar] [CrossRef]
  16. Dong, Y.; Liu, J.; Zheng, J. A sensitive dopamine electrochemical sensor based on hollow zeolitic imidazolate framework. Colloids Surf. A Physicochem. Eng. Asp. 2021, 608, 125617. [Google Scholar] [CrossRef]
  17. Xie, F.; Yang, M.; Jiang, M.; Huang, X.-J.; Liu, W.-Q.; Xie, P.-H. Carbon-based nanomaterials–a promising electrochemical sensor toward persistent toxic substance. TrAC Trends Anal. Chem. 2019, 119, 115624. [Google Scholar] [CrossRef]
  18. Ahammad, A.S.; Akter, T.; Al Mamun, A.; Islam, T.; Hasan, M.M.; Mamun, M.; Faraezi, S.; Monira, F.; Saha, J.K. Cost-effective electrochemical sensor based on carbon nanotube modified-pencil electrode for the simultaneous determination of hydroquinone and catechol. J. Electrochem. Soc. 2018, 165, B390. [Google Scholar] [CrossRef]
  19. Nair, J.A.; Saisree, S.; Aswathi, R.; Sandhya, K. Ultra-selective and real-time detection of dopamine using molybdenum disulphide decorated graphene-based electrochemical biosensor. Sens. Actuators B Chem. 2022, 354, 131254. [Google Scholar] [CrossRef]
  20. Sakthivel, R.; Kubendhiran, S.; Chen, S.-M.; Kumar, J.V. Rational design and facile synthesis of binary metal sulfides VS2-SnS2 hybrid with functionalized multiwalled carbon nanotube for the selective detection of neurotransmitter dopamine. Anal. Chim. Acta 2019, 1071, 98–108. [Google Scholar] [CrossRef]
  21. Baye, A.F.; Appiah-Ntiamoah, R.; Amalraj, J.; Reddy, K.K.; Kim, H. Graphene oxide interlayered Ga-doped FeSe2 nanorod: A robust nanocomposite with ideal electronic structure for electrochemical dopamine detection. Electrochim. Acta 2020, 363, 137245. [Google Scholar] [CrossRef]
  22. Padmanaban, A.; Padmanathan, N.; Dhanasekaran, T.; Manigandan, R.; Srinandhini, S.; Sivaprakash, P.; Arumugam, S.; Narayanan, V. Hexagonal phase Pt-doped cobalt telluride magnetic semiconductor nanoflakes for electrochemical sensing of dopamine. J. Electroanal. Chem. 2020, 877, 114658. [Google Scholar] [CrossRef]
  23. Yadav, S.; Yashas, S.R.; Shivaraju, H.P. Transitional metal chalcogenide nanostructures for remediation and energy: A review. Environ. Chem. Lett. 2021, 19, 3683–3700. [Google Scholar] [CrossRef]
  24. Zhou, J.; Liu, F.; Lin, J.; Huang, X.; Xia, J.; Zhang, B.; Zeng, Q.; Wang, H.; Zhu, C.; Niu, L. Large-area and high-quality 2D transition metal telluride. Adv. Mater. 2017, 29, 1603471. [Google Scholar] [CrossRef] [PubMed]
  25. Zaw, N.Y.W.; Jo, S.; Park, J.; Kitchamsetti, N.; Jayababu, N.; Kim, D. Clay-assisted hierarchical growth of metal-telluride nanostructures as an anode material for hybrid supercapacitors. Appl. Clay Sci. 2022, 225, 106539. [Google Scholar] [CrossRef]
  26. De Silva, U.; Masud, J.; Zhang, N.; Hong, Y.; Liyanage, W.P.; Zaeem, M.A.; Nath, M. Nickel telluride as a bifunctional electrocatalyst for efficient water splitting in alkaline medium. J. Mater. Chem. A 2018, 6, 7608–7622. [Google Scholar] [CrossRef]
  27. Brea, C.; Hu, G. Shifting and breaking scaling relations at transition metal telluride edges for selective electrochemical CO2 reduction. J. Mater. Chem. A 2022, 10, 10162–10170. [Google Scholar] [CrossRef]
  28. Qian, G.; Mo, Y.; Yu, C.; Zhang, H.; Yu, T.; Luo, L.; Yin, S. Free-standing bimetallic CoNiTe2 nanosheets as efficient catalysts with high stability at large current density for oxygen evolution reaction. Renew. Energy 2020, 162, 2190–2196. [Google Scholar] [CrossRef]
  29. Yang, F.; Yang, M.; Zhang, J.; Guo, H.; Zhang, T.; Yang, W. High-efficiency hybrid supercapacitor based on three-dimensional interconnected nitrogen-rich caCTF-1@ pC-C3N4 network and NiCoTe2. J. Alloys Compd. 2022, 911, 164782. [Google Scholar] [CrossRef]
  30. Wang, L.; Wang, J.; Yan, L.; Ding, Y.; Wang, X.; Liu, X.; Li, L.; Ju, J.; Zhan, T. Prussian Blue Analogue-Derived Iron Sulfide–Cobalt Sulfide Nanoparticle-Decorated Hollow Nitrogen-Doped Carbon Nanocubes for the Selective Electrochemical Detection of Dopamine. ACS Sustain. Chem. Eng. 2022, 10, 17230–17240. [Google Scholar] [CrossRef]
  31. Zhang, S.; Yang, D.; Zhang, M.; Liu, Y.; Xu, T.; Yang, J.; Yu, Z.-Z. Synthesis of novel bimetallic nickel cobalt telluride nanotubes on nickel foam for high-performance hybrid supercapacitors. Inorg. Chem. Front. 2020, 7, 477–486. [Google Scholar] [CrossRef]
  32. Deng, Z.-P.; Sun, Y.; Wang, Y.-C.; Gao, J.-D. A NiFe alloy reduced on graphene oxide for electrochemical nonenzymatic glucose sensing. Sensors 2018, 18, 3972. [Google Scholar] [CrossRef] [PubMed]
  33. Li, F.; Ni, B.; Zheng, Y.; Huang, Y.; Li, G. A simple and efficient voltammetric sensor for dopamine determination based on ZnO nanorods/electro-reduced graphene oxide composite. Surf. Interfaces 2021, 26, 101375. [Google Scholar] [CrossRef]
  34. Sahoo, R.C.; Moolayadukkam, S.; Thomas, S.; Zaeem, M.A.; Matte, H.R. Solution processed Ni2Co layered double hydroxides for high performance electrochemical sensors. Appl. Surf. Sci. 2021, 541, 148270. [Google Scholar] [CrossRef]
  35. Dai, H.; Chen, D.; Li, Y.; Cao, P.; Wang, N.; Lin, M. Voltammetric sensing of dopamine based on a nanoneedle array consisting of NiCo2S4 hollow core-shells on a nickel foam. Mikrochim. Acta 2018, 185, 157. [Google Scholar] [CrossRef] [PubMed]
  36. Gao, L.; Ma, J.; Zheng, J. Solvothermal synthesis of Sb2S3-graphene oxide nanocomposite for electrochemical detection of dopamine. J. Electrochem. Soc. 2020, 167, 107503. [Google Scholar] [CrossRef]
  37. Shen, H.; Jang, B.; Park, J.; Mun, H.-J.; Cho, H.-B.; Choa, Y.-H. In situ synthesis of a Bi2Te3-nanosheet/reduced-graphene-oxide nanocomposite for non-enzymatic electrochemical dopamine sensing. Nanomaterials 2022, 12, 2009. [Google Scholar] [CrossRef]
Figure 1. The schematic diagram for the synthesis of CoNiTe2.
Figure 1. The schematic diagram for the synthesis of CoNiTe2.
Chemosensors 12 00110 g001
Figure 2. (a) FESEM image and (b) TEM image of CoNiTe2 nanomaterials.
Figure 2. (a) FESEM image and (b) TEM image of CoNiTe2 nanomaterials.
Chemosensors 12 00110 g002
Figure 3. (a) STEM image of CoNiTe2 nanomaterials and their corresponding element mapping images for (b) Co, (c) Ni, and (d) Te; (e) EDS spectrum of CoNiTe2 nanomaterials.
Figure 3. (a) STEM image of CoNiTe2 nanomaterials and their corresponding element mapping images for (b) Co, (c) Ni, and (d) Te; (e) EDS spectrum of CoNiTe2 nanomaterials.
Chemosensors 12 00110 g003
Figure 4. XRD patterns of CoNiTe2 nanomaterials.
Figure 4. XRD patterns of CoNiTe2 nanomaterials.
Chemosensors 12 00110 g004
Figure 5. (a) Co 2p, (b) Ni 2p, and (c) Te 3d XPS spectra of CoNiTe2 nanomaterials.
Figure 5. (a) Co 2p, (b) Ni 2p, and (c) Te 3d XPS spectra of CoNiTe2 nanomaterials.
Chemosensors 12 00110 g005
Figure 6. CV curves of bare GCE (blue line), Nafion/GCE (pink line), and CoNiTe2/Nafion/GCE (green line) in 0.1 M PBS (pH 7.0) in the absence (dashed line) and presence (solid line) of 1 mM DA at a scan rate of 50 mV s−1.
Figure 6. CV curves of bare GCE (blue line), Nafion/GCE (pink line), and CoNiTe2/Nafion/GCE (green line) in 0.1 M PBS (pH 7.0) in the absence (dashed line) and presence (solid line) of 1 mM DA at a scan rate of 50 mV s−1.
Chemosensors 12 00110 g006
Figure 7. CV curves of CoNiTe2/Nafion/GCE with different (a) CoNiTe2 loading amounts (1~4 mg) and (b) pH values of phosphate buffer solution (pH 5.0~8.0) in 0.1 M PBS in the presence of 1 mM DA at a scan rate of 50 mV s−1. Inset of (a): plot of the oxidation peak current (Ipa) vs. different CoNiTe2 loading amounts. Bottom right inset of (b): plot of the oxidation peak current (Ipa) vs. pH. Bottom left inset of (b): plot of the oxidation peak potential (Epa) vs. pH. (The error bars stand for the standard deviation of 3 repeat measurements).
Figure 7. CV curves of CoNiTe2/Nafion/GCE with different (a) CoNiTe2 loading amounts (1~4 mg) and (b) pH values of phosphate buffer solution (pH 5.0~8.0) in 0.1 M PBS in the presence of 1 mM DA at a scan rate of 50 mV s−1. Inset of (a): plot of the oxidation peak current (Ipa) vs. different CoNiTe2 loading amounts. Bottom right inset of (b): plot of the oxidation peak current (Ipa) vs. pH. Bottom left inset of (b): plot of the oxidation peak potential (Epa) vs. pH. (The error bars stand for the standard deviation of 3 repeat measurements).
Chemosensors 12 00110 g007
Figure 8. (a) CV curves of CoNiTe2/Nafion/GCE in 0.1 M PBS (pH 6.0) in the presence of 1 mM DA at different scan rates from 50 mV s−1 to 300 mV s−1. (b) The plot of the oxidation peak current (Ipa) and reduction peak current (Ipc) vs. square root of scan rate (v1/2) (the error bars stand for the standard deviation of 3 repeat measurements).
Figure 8. (a) CV curves of CoNiTe2/Nafion/GCE in 0.1 M PBS (pH 6.0) in the presence of 1 mM DA at different scan rates from 50 mV s−1 to 300 mV s−1. (b) The plot of the oxidation peak current (Ipa) and reduction peak current (Ipc) vs. square root of scan rate (v1/2) (the error bars stand for the standard deviation of 3 repeat measurements).
Chemosensors 12 00110 g008
Figure 9. (a) DPV response of CoNiTe2/Nafion/GCE in 0.1 M PBS (pH 6.0) with the successive addition of various DA concentrations from 0 to 100 μM. (b) The plot of DPV response vs. DA concentration.
Figure 9. (a) DPV response of CoNiTe2/Nafion/GCE in 0.1 M PBS (pH 6.0) with the successive addition of various DA concentrations from 0 to 100 μM. (b) The plot of DPV response vs. DA concentration.
Chemosensors 12 00110 g009
Figure 10. Interference study of CoNiTe2/Nafion/GCE in 0.1 M PBS (pH 6.0) with the various DA concentration (0.05, 0.5, 1, 3, 5, 7.5, and 10 μM), as well as 200 μM AA and 10 μM UA.
Figure 10. Interference study of CoNiTe2/Nafion/GCE in 0.1 M PBS (pH 6.0) with the various DA concentration (0.05, 0.5, 1, 3, 5, 7.5, and 10 μM), as well as 200 μM AA and 10 μM UA.
Chemosensors 12 00110 g010
Table 1. Comparison of electrochemical DA sensor performance based on the different transition metal chalcogenides.
Table 1. Comparison of electrochemical DA sensor performance based on the different transition metal chalcogenides.
Electrode MaterialsLinear Range
(mM)
Sensitivity
(μA μM−1 cm−2)
LOD
(μM)
Reference
NiCo2S4/NF0.50~10014.99200.200[35]
Sb2S3/GO/GCE1.55~15.550.30770.800[36]
fGO-Ga0.7Fe0.3Se2/GCE2~1700.21880.110[21]
Bi2Te3/rGO/GCE10~10000.22290.060[37]
CoNiTe2/GCE0.05~1001.28800.038This Work
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, Z.-Y.; Shen, C.-H.; Yang, S.-H.; Chang, H.-W.; Tsai, Y.-C. CoNiTe2 Nanomaterials as an Efficient Non-Enzymatic Electrochemical Sensing Platform for Detecting Dopamine. Chemosensors 2024, 12, 110. https://doi.org/10.3390/chemosensors12060110

AMA Style

Wang Z-Y, Shen C-H, Yang S-H, Chang H-W, Tsai Y-C. CoNiTe2 Nanomaterials as an Efficient Non-Enzymatic Electrochemical Sensing Platform for Detecting Dopamine. Chemosensors. 2024; 12(6):110. https://doi.org/10.3390/chemosensors12060110

Chicago/Turabian Style

Wang, Zhi-Yuan, Chi-Hung Shen, Shih-Hao Yang, Han-Wei Chang, and Yu-Chen Tsai. 2024. "CoNiTe2 Nanomaterials as an Efficient Non-Enzymatic Electrochemical Sensing Platform for Detecting Dopamine" Chemosensors 12, no. 6: 110. https://doi.org/10.3390/chemosensors12060110

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