Next Article in Journal
In-Ear SpO2: A Tool for Wearable, Unobtrusive Monitoring of Core Blood Oxygen Saturation
Previous Article in Journal
SNS-CF: Siamese Network with Spatially Semantic Correlation Features for Object Tracking
Previous Article in Special Issue
N-Doped Reduced Graphene Oxide/Gold Nanoparticles Composite as an Improved Sensing Platform for Simultaneous Detection of Dopamine, Ascorbic Acid, and Uric Acid
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

CeO2-Based Two-Dimensional Layered Nanocomposites Derived from a Metal–Organic Framework for Selective Electrochemical Dopamine Sensors

1
School of Microelectronics, Southern University of Science and Technology, Shenzhen 518055, China
2
SUSTech Academy for Advanced Interdisciplinary Studies, Southern University of Science and Technology, Shenzhen 518055, China
3
Engineering Research Center of Integrated Circuits for Next-Generation Communications, Ministry of Education, Shenzhen 518055, China
*
Author to whom correspondence should be addressed.
Submission received: 12 July 2020 / Revised: 14 August 2020 / Accepted: 24 August 2020 / Published: 28 August 2020
(This article belongs to the Special Issue Nanomaterials Based Sensors and the Application)

Abstract

:
In this work, we demonstrate the incorporation of two-dimensional (2D) layered materials into a metal–organic framework (MOF) derived from one-dimensional (1D) cerium oxide (CeO2) for the electrochemical detection of dopamine. Ce-MOF was employed as a sacrificial template for preparing CeO2 with 2D materials by the pyrolysis process. The influence of the pyrolysis temperature was studied to achieve a better crystal structure of CeO2. Siloxene improved the dopamine sensing performance of CeO2 compared with graphitic carbon nitride (g-C3N4) due to the basal plane surface oxygen and hydroxyl groups of 2D siloxene. Under optimal conditions, the fabricated CeO2/siloxene electrode exhibited a detection limit of 0.292 μM, with a linear range from 0.292 μM to 7.8 μM. This work provides a novel scheme for designing the CeO2 material with siloxene for excellent dopamine sensors, which could be extended towards other biosensing applications.

1. Introduction

Dopamine (DA) is one of the most abundant catecholamine neurotransmitters, playing a crucial role in the human brain [1,2]. Abnormal dopamine content causes a few diseases, like brain aging, Parkinson’s syndrome, and schizophrenia. Thus, the detection of DA molecules in the central nervous system is becoming a necessary and significant task in the biosensor field [3,4]. Till now, many methods have been used to detect dopamine molecules, such as spectrophotometry, titrimetry, the chemiluminescence method, and the electrochemical method. Among these, the electrochemical detection method offers high sensitivity, low cost, and excellent selectivity. Hence, the electrochemical detection of DA has been considered to be effective for practical usage [5,6,7,8,9,10].
On the other hand, metal–organic frameworks (MOFs) and their derived composites have drawn much interest in various applications, such as energy storage, conversion, catalysis, separation, and gas sensors, thanks to their superior material properties like a large surface area, a tunable structure, and high thermal and chemical stability [11,12,13,14,15]. Additionally, an MOF has been employed as a flexible template to prepare porous metal oxides/sulfides and carbon materials using the direct pyrolysis process [16,17,18,19]. Up to now, various metal oxides such as Co3O4, CuO, cerium oxide (CeO2), and NiO have been synthesized using MOFs as their sacrificial template and applied in various electrochemical applications, such as supercapacitors, batteries, and biosensors [20,21,22,23].
Among them, CeO2 is a rare earth metal oxide that has attracted much interest in biomolecules detection due to its facile oxygen vacancy formation properties, excellent biocompatibility, and natural abundance. Furthermore, the stable redox behavior between Ce3+ and Ce4+ enhances the electrocatalytic properties of CeO2. Nevertheless, attaining highly electrocatalytic and selective CeO2 for DA detection is a challenging task, as a result of its poor electrical conductivity [24,25,26]. Therefore, for high sensitivity and selectivity it could be an effective strategy to combine two-dimensional (2D) conductive materials with CeO2.
Emerging two-dimensional materials such as siloxene and graphitic carbon nitride (g-C3N4) have recently been used as excellent substrates for the electrochemical sensor field because of their high surface area, excellent electrical conductivity, and better redox abilities [27,28,29,30,31]. A large number of experiments have been performed on g-C3N4 and other two-dimensional materials combined with MOFs and metal oxides for electrochemical sensors. The 2D conductive materials can increase the rate of electron transport, resulting in high electrocatalytic activity [32,33,34]. Siloxene, a type of low-buckled structure, has been explored as an excellent candidate for the dopamine sensor, and the reported detection limit was 0.327 μM [27].
Taking into consideration the above factors, we successfully synthesized the 2D siloxene and g-C3N4 with CeO2 and studied their electrocatalytic properties for DA detection. Although a few reports are available on MOF-derived CeO2 for electrochemical applications, there has been no reported work on the combination of CeO2 with g-C3N4 and siloxene, derived from a Ce-MOF for electrochemical dopamine sensors. This paper demonstrates the simple wet chemical and pyrolysis routes for the preparation of CeO2/siloxene and CeO2/g-C3N4 composites from Ce-MOF materials and investigates their electrochemical sensing ability. Combining 2D siloxene and g-C3N4 can significantly enhance the electroactive sites of CeO2 for DA detection. As a result of the synergistic effect of CeO2 and g-C3N4/siloxene, the composites exhibited a better performance towards DA detection. Interestingly, CeO2/siloxene showed higher detection performance than CeO2/g-C3N4, with a detection limit of 0.292 μM. The strong coordination modes between the Ce atom and siloxene oxygen functional groups in the CeO2/siloxene composite are beneficial for improving the DA oxidation performance with good selectivity and stability, which can be applied in future electrochemical sensor applications.

2. Materials and Methods

2.1. Materials

Cerium (II) nitrate hexahydrate (Ce(NO3)2·6H2O), 1,3,5-tricarboxylic acid (H3BTC), urea (CH4N2O), and hydrochloric acid (HCl, 37% assay) were purchased from Sigma–Aldrich. Calcium disilicide (CaSi2) was purchased from Alfa Asear, China. All of the solutions were prepared using Milli-Q water (pH 7.2).

2.2. Preparation of Ce-MOF, Ce-MOF/Siloxene, and Ce-MOF/g-C3N4 Composites

Siloxene and g-C3N4 samples were synthesized based on our previous reports [27,35]. In line with our previous report, the Ce-MOF-based composites were synthesized by a simple wet chemical route [36]. Briefly, 0.454 g of Ce(NO3)2·6H2O was added into a 40 mL of water and ethanol solution (3:1 volume ratio) and stirred for 10 min (mixture I). Meanwhile, 0.494 g of H3BTC was added into another 40 mL 25% ethanol solution and stirred for 15 min. (mixture II). Afterwards, mixture I was added drop by drop into mixture II with continuous stirring. Finally, the solid white powder was washed and centrifuged several times with ethanol and water, then dried in the vacuum oven at 80 °C for 12 h. The product was named Ce-MOF. Ce-MOF/siloxene and Ce-MOF/g-C3N4 were also synthesized by a similar procedure as the abovementioned, with the addition of 20 mg of siloxene and g-C3N4 into the H3BTC solution.

2.3. Preparation of CeO2, CeO2/Siloxene and CeO2/g-C3N4

The bare CeO2 and composites with siloxene and g-C3N4 were prepared by the pyrolysis of Ce-MOF, Ce-MOF/siloxene, and Ce-MOF/g-C3N4, respectively. The pyrolysis temperature was set to 350 and 450 °C for the three samples. The process of CeO2/siloxene composite preparation is illustrated in Scheme 1.

2.4. Materials Characterization

The Rigaku Smartlab diffractometer collected the powder X-ray diffraction (XRD) pattern of the prepared samples. Cu-Kα radiation (λ = 1.540 Å) was utilized to characterize the samples with the 2θ interval from 6 to 70°. A scanning electron microscope (SEM, Zeiss Merlin, Jena, Germany) and the high-resolution transmission electron microscope (TEM, Tecnai F30, FEI Ltd., Hillsboro, OR, USA) were used to identify the samples’ structural morphology. The porosity and surface area of the samples was characterized by the Brunauer–Emmett–Teller (BET, ASAP2020, Duluth, GA, USA) instrument.

2.5. Preparation of Modified Electrodes for Dopamine Detections and Electrochemical Analysis of Detections

Initially, aluminum oxide powder of 0.05 μM was used to polish a bare glassy carbon electrode (GCE). After that, the electrode was sonicated for 5 min in the deionized water and dried at room temperature. Meanwhile, 5 mg of active material (CeO2, CeO2/siloxene, CeO2/g-C3N4) was inserted into 5 mL of ethanol solution and sonicated for 30 min. Later, 6 μL of the active material solution was coated on the polished GCE by the drop-casting method and allowed to dry at room temperature. Scheme 2 shows the fabrication of the CeO2/siloxene- and g-C3N4-modified GCE electrodes and the DA sensing mechanism. Cyclic Voltammetry (CV) was used to optimize the electrodes for enhanced redox performance of DA detection. After that, the differential pulse voltammetry (DPV) technique was employed to investigate the DA detection of the modified electrode (CeO2/siloxene/GCE) in the 0.1 M phosphate-buffered saline (PBS, pH = 7.0) solution under different concentrations of DA. The interference study in the presence of interfering compounds was carried out after washing the electrodes by the consecutive addition of various analytes to avoid the formation of contaminates. The repeatability test was carried out with a CeO2/siloxene/GCE electrode for five consecutive runs, and the stability of the CeO2/siloxene/GCE electrode was examined for 15 days in the presence of 1000 µL of 0.2 mM of DA-containing 0.1 M PBS solution by the DPV measurement.

3. Results and Discussion

3.1. Structural and Morphological Characterization

The X-ray diffraction patterns of the Ce-MOFs composites and CeO2 composites are shown in Figure 1. The Ce-MOF’s sharp diffraction peaks are in close agreement with the XRD pattern of previously reported work [36], which confirms the better crystalline nature and phase purity of the Ce-MOF (Figure S1). After annealing, the diffraction peaks of the Ce-MOF were relocated at 28.7°, 33.1°, 47.4°, and 56.6°, corresponding to the (111), (200), (220), and (311) planes of CeO2 cubic phase, which matched with JCPDS card no. 34-0394 (Figure 1) [37]. The disappearance of the MOF peaks and the rise of new peaks after pyrolysis confirms the effective conversion of MOF to the crystalline phase of CeO2. The crystalline nature of CeO2 was investigated at two pyrolysis temperatures. As shown in Figure 1, the diffraction peak intensity of CeO2 prepared at 450 °C is higher than that of samples prepared at 350 °C, indicating better crystallinity increases in grain size. Interestingly, there is no distinct different peak among the CeO2, CeO2/siloxene, and CeO2/g-C3N4, probably due to the excellent crystallinity of CeO2 and the relatively low weight percentage of the siloxene and g-C3N4.
The morphologies of the samples as prepared were characterized by SEM and the images of the samples are shown in Figure 2. Figure 2a,b shows the nanorod structure of the Ce-MOF with a diameter of ~150 nm and lengths of a few micrometers. After being pyrolyzed, the obtained CeO2 retained a structure identical to that of the Ce-MOF without affecting the rod shape. Interestingly, a more extended nanorod structure was observed in the CeO2 sample prepared at 450 °C (Figure 2d), whereas some broken structures in the rod were observed for the CeO2 sample prepared at 350 °C (Figure 2c). The well-oriented nanorod CeO2 structure and its composites would enhance the redox abilities of DA sensors. Figure 2e–h shows the SEM images of CeO2/siloxene and CeO2/g-C3N4 samples, which were pyrolyzed at 350 °C and 450 °C, respectively. From these figures, the CeO2 nanorods were uniformly distributed on the surface of the siloxene and g-C3N4 with a slight agglomeration of CeO2 nanorods. Since the sensing properties of the active materials are strongly dependent on their crystalline nature and well-defined structure (as shown in Figure 1 and Figure 2), the samples pyrolyzed at 450 °C have been used for other characterization and electrochemical measurements, as a result of their higher crystallinity behavior.
The surface area and porosity of the as-prepared CeO2-based composites (pyrolyzed at a temperature of 450 °C) were evaluated by BET analysis, and the N2 adsorption–desorption isotherms are shown in Figure 3a,b. As shown in Figure 3a, the isotherms of bare CeO2 and CeO2/siloxene and of CeO2/g-C3N4 showed the type IV isotherm having a sharp hysteresis loop in the relative pressure between 0.8 and 1.0, which suggests that the samples have a mesoporous structure. The BET surface area is calculated as 95.91, 75.73, and 78.23 m2 g−1 for CeO2, CeO2/siloxene, and CeO2/g-C3N4, respectively. The multilayer structure of siloxene and g-C3N4 might be responsible for dropping the surface area of the composites. Though the surface area of the CeO2/siloxene composite was lower than that of CeO2/g-C3N4, the planner structure of the siloxene is beneficial for π–π interaction with the dopamine structure, resulting in the enhanced sensing properties. Additionally, the mesoporous feature of the CeO2 composites was validated with a Barrett–Joyner–Halenda (BJH) analysis, as shown in Figure 3b. On the BJH curve, the samples exhibited a peak centered at around 3.7 nm, which could be the optimal pore size for electrolyte ion transportation during the electrochemical DA detection.
The morphology of CeO2/siloxene was further confirmed by high-resolution transmission electron microscopy (HRTEM) analysis. As seen in Figure 4a–d, CeO2 nanorods with uniform size (~50 nm) and lengths of a few micrometers could be observed in the CeO2/siloxene. Additionally, the siloxene sheets were firmly anchored with the CeO2 nanorods, which can enable faster electron transportation and active sites for DA oxidation. The HRTEM image validated the crystalline phase of CeO2 and the amorphous nature of the siloxene sheets (Figure 4d). Energy-dispersive X-ray spectroscopy (EDS) mapping was used to confirm the presence of the Ce, Si, and O elements in the CeO2/siloxene, and the presence of C, N, Ce, and O in the CeO2/g-C3N4. The obtained results are shown in Figure 5. As shown in Figure 5, the homogenous distribution of the elements can be found throughout the samples. Additionally, the CeO2/siloxene sample exhibited a high density of O and Si atoms distributed on the siloxene surface, which further confirms the strong coordination between the oxygen and silicon atoms in the siloxene.

3.2. Electrochemical Detection of Dopamine

Cyclic Voltammetry (CV) was employed to record the bare GCE and CeO2-, CeO2/g-C3N4-, and CeO2/siloxene-modified GCE electrodes’ electrochemical sensing performance. A pair of redox peaks were obtained for CeO2 and its composite-modified GCE electrodes in the presence of 0.4 µM of DA, which confirms the electrocatalytic activity of the modified electrodes for DA detection (Figure 6a). The redox peaks of the modified electrodes can be assigned to the oxidation of DA to dopaminequinone (DAQ) and the reduction of DAQ to DA [38]. The high CV current response was attributed to the larger surface area of the CeO2/siloxene/GCE electrode. It can be seen that the current response for DA oxidation is significantly enhanced for the CeO2/siloxene-modified electrode, which proves the outstanding catalytic behavior over other electrodes. The strong coordination modes between the Ce atom and siloxene oxygen functional group in CeO2/siloxene composite could improve the active sites for the DA oxidation process. From the CV results, the CeO2/siloxene-modified electrode was considered an optimal electrode for further electrochemical measurements towards DA detection.
The kinetics of the electrochemical process on the CeO2/siloxene-modified electrode surface was investigated by CV at different scan rates in the presence of 0.4 µM of DA containing 0.1 M of PBS electrolyte. As shown in Figure 6b, the oxidation and reduction peaks were shifted towards positive and negative directions upon increasing the scan rate from 20 mV s−1 to 200 mV s−1. Additionally, the anodic and cathodic (Ipa and Ipc) peak currents showed an excellent linear relationship with the square root of the scan rate. They exhibited a correlation coefficient of 0.9915 and 0.9812, respectively (Figure 6c). This result confirms that DA’s oxidation behavior with the CeO2/siloxene/GCE electrode is an adsorption-controlled process [39].
Since the loading amount of the active materials on the electrode can influence the current response, the CV measurements for various amounts of CeO2/siloxene loaded on the GCE electrode were carried out in 0.4 µM of DA containing 0.1 M of PBS solution. As shown in Figure S2, the intensity of the redox peak and the current response of the 6 µL coated GCE electrode reached the maximum value, which confirms the higher DA oxidation activity on the electrode/electrolyte surface. After increasing the loading amount (8 µL and 10 µL), the redox peak current decreased gradually. As per Figure 6d, the GCE electrode coated with 6 µL of CeO2/siloxene material exhibited the highest DA oxidation current response, which indicates the significant electron transportation at the electrode/electrolyte interfaces. On the other hand, excessive loading of CeO2/siloxene may lead to the formation of a thicker film with aggregated CeO2/siloxene on the GCE surface, and could hinder the DA’s interaction with the electrode. Thus, electron transportation was blocked to the electrode surface due to the increase of the active material film’s thickness on the GCE, resulting in a lower DA oxidation current [40,41]. From this, 6 µL of CeO2/siloxene is considered to be the optimal loading for further electrochemical analysis.

3.3. DA Detection in the CeO2/Siloxene-Modified GCE Electrode

The differential pulse voltammetry technique (DPV) has several advantages: operational simplicity, high sensitivity, high accuracy, and a broad linear range for DA detection [42]. Therefore, the DPV technique was used to recognize the dopamine sensing ability of the CeO2/siloxene-modified electrode under the optimized conditions. Figure 7a shows the DPV response of the CeO2/siloxene-modified electrode at various concentrations of DA containing 0.1 M of PBS solution. From Figure 7a, it can be seen that the DA oxidation current response of the electrode gradually increased with increasing concentrations of DA (0 to 9.8 µM), proving the excellent electrocatalytic behavior of the electrode for dopamine. The fabricated CeO2/siloxene sensor possesses a good linear range between DA concentration and the oxidation current with a correlation coefficient R2 = 0.9808 (Figure 7b). Moreover, the sensor exhibited a detection limit of 0.292 μM.
Table 1 shows the electrocatalytic DA sensing performance of our sensor compared with previously reported electrodes. As the table shows, the CeO2/siloxene sensor exhibited a superior performance of DA detection, which shows a promising application in the biosensor field. The repeatability of the sensor is an essential characteristic of the electrochemical detection of DA. Figure 7c shows the DPV current response of a CeO2/siloxene-modified electrode in 0.4 µM DA containing 0.1 M PBS solution for ten replicant measurements. The DA oxidation current response of all of the runs is almost at the same level, demonstrating the excellent repeatability of the sensing performance.
Selectivity is another crucial factor in discriminating between some common species in similar analytes. Figure 8 represents an interference study of the CeO2/siloxene-modified electrode in the presence of other interfering compounds such as ascorbic acid (AA), uric acid (UA), Adenosine triphosphate (ATP), glucose, and rutin. As Figure 8a shows, the initial current response was observed after the addition of 1000 µL of 0.2 mM DA solution into 0.1 M PBS electrolyte. Afterwards, 1000 µL of 0.2 mM of rutin and ATP were added, and negligible changes in the oxidation current could be noticed. A remarkable change in the oxidation current was observed after the second addition of 1000 µL DA, indicating the sensor’s excellent selectivity. Furthermore, the other interfering compounds (AA, UA, and glucose, 1000 µL of 0.2 mM) were also injected into the PBS solution, which revealed no fluctuations in the current and maintained the same DA oxidation current response (Figure 8b). The results prove that the CeO2/siloxene-modified electrode has better selectivity towards detecting DA molecules under optimal conditions. Stability is a vital characteristic for sensors; the DPV analysis was performed to investigate the stability of the CeO2/siloxene-modified GCE electrode. The electrode was kept in the refrigerator when not in use. As per Figure 8c, there was only a tiny change in the DA oxidation current even after 15 days compared with the initial measurement (first day). The DA oxidation current of 82.3% was maintained after two weeks of measurements, which indicates the good stability of the modified electrode for the DA sensor. Additionally, the sensor’s recovery was investigated by the DPV technique before and after washing the electrodes in the PBS solution and DA containing the PBS solution. As per Figure 9a,b, the CeO2/siloxene-modified electrode possessed an excellent recovery characteristic in the PBS solution after washing the electrodes. However, a tiny change in the current response was observed after washing the electrode (washed 2 and washed 3 in Figure 9b) compared with the initial measurement (blank) in the PBS solution, suggesting the formation of a small quantity of contaminants on the electrode surface. Although a small difference was noticed after washing the electrodes, it was negligible compared to the DA oxidation current. These results demonstrate that one-dimensional (1D) CeO2 nanorods with 2D siloxene sheets could be employed as a selective dopamine sensor.

4. Conclusions

In this work, we successfully combined Ce-MOF with siloxene and g-C3N4 and prepared CeO2/siloxene and CeO2/g-C3N4 composites using the pyrolysis method. The as-prepared samples were characterized by different tools and optimized the pyrolysis condition for better crystallization of CeO2. Compared with pure CeO2 and CeO2/g-C3N4-modified electrodes, the CeO2/siloxene-modified GCE exhibited superior electrochemical performance towards DA detection thanks to its strong redox mode between Ce and siloxene functional groups. The proposed sensor showed a detection limit of 0.292 μM with a linear range between 0.292 μM and 7.8 μM. Moreover, the CeO2/siloxene-modified sensor possessed good repeatability and selectivity for DA detection. This work provides a new insight into preparing CeO2 with siloxene from an MOF for highly electrocatalytic DA sensors and could be extended to the detection of other suitable biomolecules in the near future.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/1424-8220/20/17/4880/s1, Figure S1: XRD pattern of the Ce-MOF, Figure S2: Cyclic voltammetry response of different amount of CeO2/siloxene loading in 0.4 µM concentration DA containing 0.1 M PBS solution.

Author Contributions

C.G.: Methodology and Writing—original draft; R.R.: Conceptualization, Supervision and Writing—review & editing. F.W.: Supervision, Writing—review & editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported in part by the National Natural Science Foundation of China (Project No. 51950410598), in part by the Shenzhen Science and Technology Innovation Committee (Projects No. JCYJ20170412154426330), and in part by the Guangdong Natural Science Funds (Project No.: 2016A030306042 and 2018A050506001).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wightman, R.M.; May, L.J.; Michael, A.C. Detection of dopamine dynamics in the brain. Anal. Chem. 1988, 60, 769A–793A. [Google Scholar] [CrossRef] [PubMed]
  2. Cheng, M.; Zhang, X.; Wang, M.; Huang, H.; Ma, J. A facile electrochemical sensor based on well-dispersed graphene-molybdenum disulfide modified electrode for highly sensitive detection of dopamine. J. Electroanal. Chem. 2017, 786, 1–7. [Google Scholar] [CrossRef]
  3. Liao, C.; Zhang, M.; Niu, L.; Zheng, Z.; Yan, F. Organic electrodes for highly sensitive and selective dopamine sensors. J. Mater. Chem. B 2014, 2, 191–200. [Google Scholar] [CrossRef] [PubMed]
  4. Hui, X.; Xuan, X.; Kim, J.; Park, J. A highly flexible and selective dopamine sensor based on Pt-Au nanoparticle-modified laser-induced graphene. Electrochim. Acta 2019, 328, 135066. [Google Scholar] [CrossRef]
  5. Jiao, J.; Zuo, J.; Pang, H.; Tan, L.; Chen, T.; Ma, H. A dopamine electrochemical sensor based on Pd-Pt alloy nanoparticles decorated polyoxometalate and multiwalled carbon nanotubes. J. Electroanal. Chem. 2018, 827, 103–111. [Google Scholar] [CrossRef]
  6. Lin, M.; Song, P.; Zhou, G.; Zuo, X.; Aldalbahi, A.; Lou, X.; Shi, J.; Fan, C. Electrochemical detection of nucleic acids, proteins, small molecules and cells using a DNA-nanostructure-based universal biosensing platform. Nat. Protoc. 2016, 11, 1244–1263. [Google Scholar] [CrossRef]
  7. Koo, K.M.; Carrascosa, L.G.; Shiddiky, M.J.; Analy, M.T. Poly(A) Extensions of miRNAs for Amplification-Free Electrochemical Detection on Screen-Printed Gold Electrodes. Anal. Chem. 2016, 88, 2000–2005. [Google Scholar] [CrossRef] [Green Version]
  8. Haque, M.H.; Gopalan, V.; Yadav, S.; Islam, M.N.; Eftekhari, E.; Lin, Q.; Carrascosa, L.G.; Nguyen, N.T.; Lam, A.K.; Shiddiky, M.J. Detection of regional DNA methylation using DNA-graphene affinity interactions. Biosens. Bioelectron. 2017, 87, 615–621. [Google Scholar] [CrossRef] [Green Version]
  9. Joutsa, J.; Voon, V.; Johansson, J.; Niemela, S.; Bergman, J.; Kaasinen, V. Dopaminergic function and intertemporal choice. Transl. Psychiatry 2015, 5, 491. [Google Scholar] [CrossRef]
  10. Zhuang, X.; Mazzoni, P.; Kang, U.J. The role of neuroplasticity in dopaminergic therapy for Parkinson disease. Nat. Rev. Neurol. 2013, 9, 248–256. [Google Scholar] [CrossRef]
  11. Ramachandran, R.; Zhao, C.; Rajkumar, M.; Rajavel, K.; Zhu, P.; Xuan, W.; Xu, Z.X.; Wang, F. Porous nickel oxide microsphere and Ti3C2Tx hybrid derived from metal-organic framework for battery-type supercapacitor electrode and non-enzymatic H2O2 sensor. Electrochim. Acta 2019, 322, 134771. [Google Scholar] [CrossRef]
  12. Kong, B.; Tang, J.; Wu, Z.; Wei, J.; Wu, H.; Wang, Y.; Zheng, G.; Zhao, D. Ultralight Mesoporous Magnetic Frameworks by Interfacial Assembly of Prussian Blue Nanocubes. Angew. Chem. Int. Ed. 2014, 53, 2888–2892. [Google Scholar] [CrossRef] [PubMed]
  13. Xu, X.; Zhang, Z.; Wang, X. Well-Defined Metal-Organic-Framework Hollow Nanostructures for Catalytic Reactions in Volving Gases. Adv. Mater. 2015, 27, 5365–5371. [Google Scholar] [CrossRef] [PubMed]
  14. Huang, M.; Mi, K.; Zhang, J.; Liu, H.; Yu, T.; Yuan, A.; Kong, Q.; Xiong, S. MOF-Derived Bi-Metal Embedded N-Doped Carbon Polyhedral Nanocages with Enhanced Lithium Storage. J. Mater. Chem. A 2017, 5, 266–274. [Google Scholar] [CrossRef]
  15. Meng, J.; Niu, C.; Xu, L.; Li, J.; Liu, X.; Wang, X.; Wu, Y.; Xu, X.; Chen, W.; Li, Q.; et al. General Oriented Formation of Carbon Nanotubes from Metal-Organic Frameworks. J. Am. Chem. Soc. 2017, 139, 8212–8221. [Google Scholar] [CrossRef]
  16. Yang, Q.; Xu, Q.; Jiang, H.L. Metal-Organic Frameworks Meet Metal Nanoparticles: Synergistic Effect for Enhanced Catalysis. Chem. Soc. Rev. 2017, 46, 4774–4808. [Google Scholar] [CrossRef]
  17. Han, L.; Yu, X.Y.; Lou, X.W. Formation of Prussian-Blue-Analog Nanocages via a Direct Etching Method and Their Conversion into Ni-Co-Mixed Oxide for Enhanced Oxygen Evolution. Adv. Mater. 2016, 28, 4601–4605. [Google Scholar] [CrossRef]
  18. Ji, W.; Xu, Z.; Liu, P.; Zhang, S.; Zhou, W.; Li, H.; Zhang, T.; Li, L.; Lu, X.; Wu, J.; et al. Metal-Organic Framework Derivatives for Improving the Catalytic Activity of the CO Oxidation Reaction. ACS Appl. Mater. Interfaces 2017, 9, 15394–15398. [Google Scholar] [CrossRef]
  19. Liu, H.; Zhang, S.; Liu, Y.; Yang, Z.; Feng, X.; Lu, X.; Huo, F. Well-Dispersed and Size-Controlled Supported Metal Oxide Nano-particles Derived from MOF Composites and Further Application in Catalysis. Small 2015, 11, 3130–3134. [Google Scholar] [CrossRef]
  20. Chen, T.; Liu, X.; Niu, L.; Gong, Y.; Li, C.; Xu, S.; Pan, L. Recent progress on metal–organic frameworkderived materials for sodium-ion battery anodes. Inorg. Chem. Front. 2020, 3, 567–582. [Google Scholar] [CrossRef]
  21. Guo, Z.; Song, L.; Xu, T.; Gao, D.; Li, C.; Hu, X.; Chen, G. CeO2-CuO bimetal oxides derived from Ce-based MOF and their difference in catalytic activities for CO oxidation. Mater. Chem. Phys. 2019, 26, 338–343. [Google Scholar] [CrossRef]
  22. Xie, X.; Huang, K.; Wu, X. Metal-organic framework derived hollow materials for electrochemical energy storage. J. Mater. Chem. A 2017, 7, 6754–6771. [Google Scholar] [CrossRef]
  23. Zhang, S.; Gao, H.; Xu, X.; Cao, R.; Yang, H.; Xu, X.; Li, J. MOF-derived CoN/N-C@SiO2 yolk-shell nanoreactor with dual active sites for highly efficient catalytic advanced oxidation processes. Chem. Eng. J. 2020, 381, 122670. [Google Scholar] [CrossRef]
  24. Phoka, S.; Laokul, P.; Swatsitang, E.; Promarak, V.; Seraphin, S.; Maensiri, S. Synthesis, structural and optical properties of CeO2 nanoparticles synthesized by a simple polyvinyl pyrrolidone (PVP) solution route. Mater. Chem. Phys. 2019, 115, 423–428. [Google Scholar] [CrossRef]
  25. Dhall, A.; Self, W. Cerium oxide nanoparticles: A brief review of their synthesis methods and biomedical applications. Antioxidants 2018, 7, 97. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Goharshadi, E.K.; Samiee, S.; Nancarrow, P. Fabrication of cerium oxide nanoparticles: Characterizations and optical properties. J. Colloid Interface Sci. 2011, 356, 473–480. [Google Scholar] [CrossRef] [PubMed]
  27. Ramachandran, R.; Len, X.; Zhao, C.; Xu, Z.; Wang, F. 2D siloxene sheets: A novel electrochemical sensor for selective dopamine detection. Appl. Mater. Today 2019, 18, 100477. [Google Scholar] [CrossRef]
  28. Zhang, Y.L.; Li, B.; Guan, W.; Wei, Y.N.; Yan, C.H.; Meng, M.J.; Pan, J.M.; Yan, Y.S. One-pot synthesis of HMF from carbohydrates over acid-base bi-functional carbonaceous catalyst supported on halloysite nanotubes. Cellulose 2020, 27, 3037–3054. [Google Scholar] [CrossRef]
  29. Zhu, Z.; Huo, P.W.; Lu, Z.Y.; Yan, Y.S.; Liu, Z.; Shi, W.D.; Li, C.X.; Dong, H.J. Fabrication of magnetically recoverable photocatalysts using g-C3N4 for effective separation of charge carriers through like-Z-scheme mechanism with Fe3O4 mediator. Chem. Eng. J. 2018, 331, 615–625. [Google Scholar] [CrossRef]
  30. Hu, S.; Ma, L.; You, J.; Li, F.; Fan, Z.; Lu, G.; Liu, D.; Gui, J. Enhanced visible light photocatalytic performance of g-C3N4 photocatalysts Co-doped with iron and phosphorus. Appl. Surf. Sci. 2014, 311, 164–171. [Google Scholar] [CrossRef]
  31. Oh, W.D.; Chang, V.W.C.; Hu, Z.T.; Goei, R.; Lim, T.T. Enhancing the catalytic activity of g-C3N4 through Me doping (Me = Cu, Co and Fe) for selective sulfathiazole degradation via redox-based advanced oxidation process. Chem. Eng. J. 2017, 323, 260–269. [Google Scholar] [CrossRef]
  32. Mo, Z.; She, X.; Li, Y.; Liu, L.; Huang, L.; Chen, Z.; Zhang, Q.; Xu, H.; Li, H. Synthesis of g-C3N4 at different temperatures for superior visible/UV photocatalytic performance and photoelectrochemical sensing of MB solution. RSC Adv. 2015, 5, 101552–101562. [Google Scholar] [CrossRef]
  33. Zou, J.; Wu, S.; Liu, Y.; Sun, Y.; Cao, Y.; Hsu, J.; Wee, A.; Jiang, J. An ultrasensitive electrochemical sensor based on 2D g-C3N4/CuO nanocomposites for dopamine detection. Carbon 2018, 130, 652–663. [Google Scholar] [CrossRef]
  34. Xavier, M.; Nair, P.; Mathew, S. Emerging trends in sensors based on carbon nitride materials. Analyst 2019, 144, 1475–1491. [Google Scholar] [CrossRef] [PubMed]
  35. Sakthivel, T.; Ramachandran, R.; Kirubakaran, K. Photocatalytic properties of copper-two dimensional graphitic carbon nitride hybrid film synthesized by pyrolysis method. J. Environ. Chem. Eng. 2018, 6, 2636–2642. [Google Scholar] [CrossRef]
  36. Ramachandran, R.; Xuan, W.; Zhao, C.; Len, X.; Sun, D.; Luo, D.; Wang, F. Enhanced electrochemical properties of cerium metal–organic framework based composite electrodes for high-performance supercapacitor application. RSC Adv. 2018, 8, 3462–3469. [Google Scholar] [CrossRef] [Green Version]
  37. Chen, X.; Yu, E.; Cai, S.; Jia, H.; Chen, J.; Liang, P. In situ pyrolysis of Ce-MOF to prepare CeO2 catalyst with obviously improved catalytic performance for toluene combustion. Chem. Eng. J. 2018, 334, 469–479. [Google Scholar] [CrossRef]
  38. Rajkumar, C.; Thirumalraj, B.; Chen, S.-M.; Chen, H.-A. A simple preparation of graphite/gelatin composite for electrochemical detection of dopamine. J. Colloid Interface Sci. 2017, 487, 149–155. [Google Scholar] [CrossRef]
  39. Saenz, H.S.C.; Saravia, L.P.H.; Selva, J.S.G.; Sukeri, A.; Montero, P.J.E.; Bertotti, M. Electrochemical dopamine sensor using a nanoporous gold microelectrode: A proof-of-concept study for the detection of dopamine release by scanning electrochemical microscopy. Microchim. Acta 2018, 185, 367. [Google Scholar] [CrossRef]
  40. Sukanya, R.; Ramki, S.; Chen, S.-M.; Karthik, R. Ultrasound treated cerium oxide/tin oxide (CeO2/SnO2) nanocatalyst: A feasible approach and enhanced electrode material for sensing of anti-inflammatory drug 5-aminosalicylic acid in biological samples. Anal. Chim. Acta 2020, 1096, 76–88. [Google Scholar] [CrossRef]
  41. Sakthivel, R.; Kubendhiran, S.; Chen, S.-M. One-pot sonochemical synthesis of marigold flower-like structured ruthenium doped bismuth sulfide for the highly sensitive detection of antipsychotic drug thioridazine in the human serum sample. J. Taiwan Inst. Chem. Eng. 2020, 111, 270–282. [Google Scholar] [CrossRef]
  42. Vali, A.; Malayeri, H.Z.; Azizi, M.; Choi, H. DPV-assisted understanding of TiO2 photocatalytic decomposition of aspirin by identifying the role of produced reactive species. Appl. Catal. B Environ. 2020, 266, 118646. [Google Scholar] [CrossRef]
  43. Moccelini, S.; Fernandes, S.; Vieira, I. Bean sprout peroxidase biosensor based on l-cysteine self-assembled monolayer for the determination of dopamine. Sens. Actuators B Chem. 2018, 133, 364–369. [Google Scholar] [CrossRef]
  44. Roychoudhury, A.; Basu, S.; Jha, S. Dopamine biosensor based on surface functionalized nanostructured nickel oxide platform. Biosens. Bioelectron. 2016, 84, 72–81. [Google Scholar] [CrossRef]
  45. Derviseevic, M.; Senel, M.; Cevik, E. Novel impedimetric dopamine biosensor based on boronic acid functional polythiophene modified electrodes. Mater. Sci. Eng. C 2017, 72, 641–649. [Google Scholar] [CrossRef]
  46. Chandra, S.; Arora, K.; Bahadur, D. Impedimetric biosensor based on magnetic nanoparticles for electrochemical detection of dopamine. Mater. Sci. Eng. B 2012, 177, 1531–1537. [Google Scholar] [CrossRef]
  47. How, G.T.S.; Pandikumar, A.; Ming, H.N.; Ngee, L.H. Highly exposed {001} factes of titanium dioxide modified with reduced graphene oxide for dopamine sensing. Sci. Rep. 2014, 4, 5044. [Google Scholar]
  48. Mani, V.; Devasenathipathy, R.; Chen, S.-M.; Kohilarani, K.; Ramachandran, R. A sensitive amperometric sensor for the determination of dopamine at graphene and bismuth nanocomposite film modified electrode. Int. J. Electrochem. Sci. 2015, 10, 1199–1207. [Google Scholar]
Scheme 1. Schematic diagram of cerium oxide (CeO2)/siloxene composite preparation.
Scheme 1. Schematic diagram of cerium oxide (CeO2)/siloxene composite preparation.
Sensors 20 04880 sch001
Scheme 2. Electrode modification process and dopamine sensing mechanism of the modified electrode.
Scheme 2. Electrode modification process and dopamine sensing mechanism of the modified electrode.
Sensors 20 04880 sch002
Figure 1. X-ray diffraction pattern of (a) cerium oxide (CeO2) and composites pyrolyzed at 350 °C, and (b) CeO2 and composites pyrolyzed at 450 °C.
Figure 1. X-ray diffraction pattern of (a) cerium oxide (CeO2) and composites pyrolyzed at 350 °C, and (b) CeO2 and composites pyrolyzed at 450 °C.
Sensors 20 04880 g001
Figure 2. Scanning electron microscope (SEM) images of (a,b) Ce-MOF, pyrolyzed at 350 °C; (c) Bare CeO2; (e) CeO2/siloxene; (g) CeO2/graphitic carbon nitride (g-C3N4), pyrolyzed at 450 °C; (d) Bare CeO2; (f) CeO2/siloxene; and (h) CeO2/g-C3N4.
Figure 2. Scanning electron microscope (SEM) images of (a,b) Ce-MOF, pyrolyzed at 350 °C; (c) Bare CeO2; (e) CeO2/siloxene; (g) CeO2/graphitic carbon nitride (g-C3N4), pyrolyzed at 450 °C; (d) Bare CeO2; (f) CeO2/siloxene; and (h) CeO2/g-C3N4.
Sensors 20 04880 g002
Figure 3. (a) Nitrogen adsorption–desorption isotherms; (b) Barrett–Joyner–Halenda (BJH) pore size distribution curves.
Figure 3. (a) Nitrogen adsorption–desorption isotherms; (b) Barrett–Joyner–Halenda (BJH) pore size distribution curves.
Sensors 20 04880 g003
Figure 4. (ac) Transmission electron microscopy (TEM) and (d) High-resolution transmission electron microscopy (HRTEM) images of the CeO2/siloxene composite.
Figure 4. (ac) Transmission electron microscopy (TEM) and (d) High-resolution transmission electron microscopy (HRTEM) images of the CeO2/siloxene composite.
Sensors 20 04880 g004
Figure 5. Energy-dispersive X-ray spectroscopy (EDS) elemental mapping of (ad) CeO2/g-C3N4 and (eh) CeO2/siloxene composites.
Figure 5. Energy-dispersive X-ray spectroscopy (EDS) elemental mapping of (ad) CeO2/g-C3N4 and (eh) CeO2/siloxene composites.
Sensors 20 04880 g005
Figure 6. (a) Cyclic voltammetry response of bare glassy carbon electrodes (GCEs) and modified GCEs in 0.4 µM concentration dopamine (DA) containing 0.1 M phosphate-buffered saline (PBS) solution; (b) Cyclic voltammetry response of CeO2/siloxene-modified GCE electrode at different scan rates in 0.4 µM concentration DA containing 0.1 M PBS; (c) Calibration curve of the anodic and cathodic peak current versus the scan rate of the CeO2/siloxene-modified electrode; (d) The cyclic voltammetry peak current versus the different amounts of CeO2/siloxene loading.
Figure 6. (a) Cyclic voltammetry response of bare glassy carbon electrodes (GCEs) and modified GCEs in 0.4 µM concentration dopamine (DA) containing 0.1 M phosphate-buffered saline (PBS) solution; (b) Cyclic voltammetry response of CeO2/siloxene-modified GCE electrode at different scan rates in 0.4 µM concentration DA containing 0.1 M PBS; (c) Calibration curve of the anodic and cathodic peak current versus the scan rate of the CeO2/siloxene-modified electrode; (d) The cyclic voltammetry peak current versus the different amounts of CeO2/siloxene loading.
Sensors 20 04880 g006
Figure 7. (a) Differential pulse voltammetry (DPV) response of the CeO2/siloxene-modified electrode in 0.1 M PBS solution containing different concentrations of DA; (b) Linear response of the oxidation current versus DA concentrations; (c) Repeatability of a modified electrode in the presence of 1 µM of DA-containing 0.1 M PBS solution; (d) Plot of the oxidation current versus the run number for five consecutive measurements.
Figure 7. (a) Differential pulse voltammetry (DPV) response of the CeO2/siloxene-modified electrode in 0.1 M PBS solution containing different concentrations of DA; (b) Linear response of the oxidation current versus DA concentrations; (c) Repeatability of a modified electrode in the presence of 1 µM of DA-containing 0.1 M PBS solution; (d) Plot of the oxidation current versus the run number for five consecutive measurements.
Sensors 20 04880 g007
Figure 8. (a) Interference analysis of the CeO2/siloxene-modified electrode in the presence of other interfering compounds; (b) Current change response to the interfering compounds; (c,d) Stability of the CeO2/siloxene-modified electrode in the presence of 1000 µL of 0.2 mM of DA containing 0.1 M PBS solution.
Figure 8. (a) Interference analysis of the CeO2/siloxene-modified electrode in the presence of other interfering compounds; (b) Current change response to the interfering compounds; (c,d) Stability of the CeO2/siloxene-modified electrode in the presence of 1000 µL of 0.2 mM of DA containing 0.1 M PBS solution.
Sensors 20 04880 g008
Figure 9. (a) DPV response of the CeO2/siloxene modified electrode before and after washing in blank and DA containing PBS solutions; (b) Current changes bar chart before and after washed the electrodes.
Figure 9. (a) DPV response of the CeO2/siloxene modified electrode before and after washing in blank and DA containing PBS solutions; (b) Current changes bar chart before and after washed the electrodes.
Sensors 20 04880 g009
Table 1. Comparison of the analytical properties of other reported dopamine (DA) electrochemical sensors.
Table 1. Comparison of the analytical properties of other reported dopamine (DA) electrochemical sensors.
ElectrodeTechniqueLimit of Detection (μM)Linear Range (μM)Ref.
Beansprout/(SAM)/AuSWV0.4789.9–2210[43]
Tyrosinase/NiO/ITOCV1.0322–100[44]
P(TBA0.50Th0.50)EIS/CV0.37.8–125[45]
PA-MWCNT/GCEEIS14.110–1000[46]
rGO/TiO2DPV1.51–35; 35–100[47]
Ag/Graphene/GCELSV5.410–800[48]
CeO2/siloxene/GCEDPV0.2920.292–7.8This work

Share and Cite

MDPI and ACS Style

Ge, C.; Ramachandran, R.; Wang, F. CeO2-Based Two-Dimensional Layered Nanocomposites Derived from a Metal–Organic Framework for Selective Electrochemical Dopamine Sensors. Sensors 2020, 20, 4880. https://0-doi-org.brum.beds.ac.uk/10.3390/s20174880

AMA Style

Ge C, Ramachandran R, Wang F. CeO2-Based Two-Dimensional Layered Nanocomposites Derived from a Metal–Organic Framework for Selective Electrochemical Dopamine Sensors. Sensors. 2020; 20(17):4880. https://0-doi-org.brum.beds.ac.uk/10.3390/s20174880

Chicago/Turabian Style

Ge, Chengjie, Rajendran Ramachandran, and Fei Wang. 2020. "CeO2-Based Two-Dimensional Layered Nanocomposites Derived from a Metal–Organic Framework for Selective Electrochemical Dopamine Sensors" Sensors 20, no. 17: 4880. https://0-doi-org.brum.beds.ac.uk/10.3390/s20174880

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