Next Article in Journal
Generating Self-Shaped 2D Aluminum Oxide Nanopowders
Next Article in Special Issue
Disposal Immunosensor for Sensitive Electrochemical Detection of Prostate-Specific Antigen Based on Amino-Rich Nanochannels Array-Modified Patterned Indium Tin Oxide Electrode
Previous Article in Journal
Box–Behnken Design: Optimization of Proanthocyanidin-Loaded Transferosomes as an Effective Therapeutic Approach for Osteoarthritis
Previous Article in Special Issue
Fabrication of an Organofunctionalized Talc-like Magnesium Phyllosilicate for the Electrochemical Sensing of Lead Ions in Water Samples
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Covalent Organic Frameworks-TpPa-1 as an Emerging Platform for Electrochemical Sensing

School of Chemistry and Materials Science, Ludong University, Yantai 264025, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(17), 2953; https://doi.org/10.3390/nano12172953
Submission received: 2 August 2022 / Revised: 19 August 2022 / Accepted: 22 August 2022 / Published: 26 August 2022

Abstract

:
Covalent organic frameworks (COFs) are a new type of metal-free porous architecture with a well-designed pore structure and high stability. Here an efficient electrochemical sensing platform was demonstrated based on COFs TpPa-1 constructed by 1,3,5-triformylphloroglucinol (Tp) with p-phenylenediamine (Pa-1), which possesses abundant nitrogen and oxo-functionalities. COFs TpPa-1 exhibited good water dispersibility and strong adsorption affinities for Pd2+ and thus was used as loading support to modify Pd2+. The Pd2+-modified COFs TpPa-1 electrode (Pd2+/COFs) showed high electrocatalytic activity for both hydrazine oxidation reaction and nitrophenol reduction reaction. In addition, TpPa-1-derived nitrogen-doped carbon presented high activity for the electro-oxidation of reduced glutathione (GSH), and sensitive electrochemical detection of GSH was achieved. The presented COFs TpPa-1 can be utilized as a precursor as well as support for anchoring electro-active molecules and nanoparticles, which will be useful for electrochemical sensing and electrocatalysis.

Graphical Abstract

1. Introduction

The synthesis and construction of efficient electrode materials to gain high sensitivity and selectivity is crucial for electroanalysis. The successful development of nonenzymatic electrochemical sensors depends critically on electrode materials with excellent electrocatalytic properties because many analytes are electrochemically inactive when using traditional electrodes. Up to now, different categories of materials, including noble metals, metal oxides, and carbons (such as CNT and graphene), have been applied in electrochemical sensing [1]. Recently, porous materials have demonstrated advantages in constructing electrochemical sensors due to their high specific surface areas and powerful adsorption and loading capability [2].
Metal–organic frameworks (MOFs) and covalent organic frameworks (COFs), as two emerging families of crystalline porous materials, have been explored in electroanalysis and electrocatalysis [3,4]. Compared to conventional porous materials such as zeolites, porous oxides, and carbons, MOFs and COFs show some intriguing features for electrochemical applications, including inter-connected porosity, tunable intra-framework chemical functionality, and high surface area [5]. Wang et al. synthesized graphene aerogel (GA) and metal–organic framework (MOF) composites via the in situ growth of the MOF UiO-66-NH2 crystal on the GA matrix. The interface showed high sensitivity and selectivity for the electrochemical sensing of multiple heavy-metal ions. GA was employed as the backbone for UiO-66-NH2, which also improved the electrical conductivity of the composites by accelerating the electron transfer in the matrix. UiO-66-NH2 provided a binding site for heavy-metal ions due to the interaction between hydrophilic species and metal cations [6]. Xu et al. developed a facile furazolidone sensor based on COFs [7]. In this work, 1,3,5-tris-(4-amino-phenyl) benzene (TAPB) and terephthaldicarboxaldehyde (TPA) were used as monomers to synthesize the COF, which was then self-assembled on the surface of the NH2-CNTs via a simple and rapid one-pot strategy at room temperature. The prepared COF@NH2-CNT composites exhibited a high surface area and excellent conductivity, thus achieving high sensitive detection of furazolidone. However, most MOFs are not chemically stable in water. If the MOF is unstable in the aqueous buffer used for electroanalysis, the MOF coating modified on the electrode surface may suffer from dissolving or converting into MOF-derived metal hydroxides or oxides [8]. Unlike MOFs, COFs are regularly integrated into metal-free crystalline organic structures through strong covalent bonds between the organic building blocks, leading to more robust frameworks and relatively lower densities. Many COFs possess the sp2 carbon networks with a high degree of p-conjunction and thus lead to at least a modest electronic conductivity, which facilitates the electron transfer rate. Furthermore, the presence of electronegative atoms (N and O) and π bonds within COFs can allow more specific target analyte recognition through the formation of hydrogen bonds or π-π bonds, thus improving the sensitivity and selectivity of COF-based electrochemical sensors [7]. These unique characteristics of COFs make them promising materials for electrochemical sensing.
COFs-derived porous carbons are also considered ideal platforms for electrochemical sensing due to their high conductivity, porosity, chemical and thermal stability, high electrocatalytic activity, and powerful supporting performances for anchoring molecules and nanomaterials. Most intrinsic electrocatalytic activities are contributed to the oxo-functionalities and edge-plane-like sites (defects) present on the surface of carbon materials [9]. COFs are often rich with nonmetal elements such as N, B, S, and P, which can induce heteroatom doping into derived carbons. Heteroatom doping forms charge delocalization and changes in the electronic structure due to the difference in atomic size, bond length, and coordination between the carbon atoms and dopants, leading to abundant defect sites [10]. N-doping can increase the surface polarity and wettability of the electrodes and also improve the electrocatalytic activity and adsorption [11,12]. Here COFs TpPa-1 constructed by 1,3,5-triformylphloroglucinol (Tp) with p-phenylenediamine (Pa-1) possess abundant nitrogen and oxo-functionalities, which was demonstrated for Pd2+ immobilization and fabricating nitrogen-doped porous carbon. The presented electrocatalysts showed high responses for hydrazine oxidation, nitrophenol reduction, and glutathione (GSH) oxidation.

2. Experimental

2.1. Chemicals and Solutions

GSH, palladium chloride (PdCl2), hydrazine (wt. 80%), graphite powder, liquid paraffin, p-nitrophenol (PNP), and o-nitrophenol (ONP) were purchased from Sigma-Aldrich (St. Louis, MO, USA). COFs-TpPa-1 was purchased from Nanjing XFNano Materials Tech Co., Ltd. (Nanjing, China) The pore size of COFs-TpPa-1 is in the range of 1.5–1.8 nm with a BET specific surface area of ~1360 m2/g. All of the other chemicals were of analytical reagent grade and used without further purification.

2.2. Apparatus

The morphologies were characterized by scanning electron microscopy (SEM) (Hitachi SU8010, Tokyo, Japan) and transmission electron microscopy (TEM) (JEM-2100, JEOL, Tokyo, Japan). The X-ray photoelectron spectroscopy (XPS) was recorded on a Thermo ESCALAB 250 Xi spectrometer (Waltham, MA, USA) fitted with a monochromatic Al Kα X-ray source. X-ray diffraction patterns were collected by a Bruker D8 (Billerica, MA, USA) advance powder X-ray Cu Ka radiation diffractometer. All of the electrochemical experiments were conducted on a CHI 750E electrochemical workstation with a conventional three-electrode system consisting of a modified working electrode, platinum coil auxiliary electrode, and Ag/AgCl (saturated KCl) reference electrode. Electrochemical impedance spectra (EIS) were carried out in 0.1 M KCl containing 5.0 mM Fe (CN)63−/4− in the frequency range of 1 MHz to 0.1 Hz at 0.25 V.

2.3. Electrode Preparation and Modification

2.3.1. Preparation of Pd2+/COFs/GCE

Prior to modification, the glassy carbon electrode (GCE) (3 mm diameter, 0.07 cm2) was successively polished with 1, 0.3, and 0.05 µm alumina paste to a mirror finish and then rinsed with deionized water followed by ultrasonic treatment in water and ethanol, respectively. The COFs-modified GCE (COFs/GCE) was prepared by dropping 10.0 µL COFs aqueous suspension (2.0 mg/mL) on a cleaned GCE and dried at room temperature. In order to anchor Pd2+, the prepared COFs/GCE was dipped in 10 mM PdCl2 solution for 30 min, which was then washed with water and left to dry at room temperature (Pd2+/COFs/GCE). Pd2+/GCE was prepared by soaking the GCE in Pd2+ solution by electroless deposition. The electroactive surface areas of the different electrodes could be calculated by the Randles–Sevcik equation using the CV responses of the corresponding electrodes in 0.1 M KCl solution in the presence of 5.0 mM K3Fe (CN)6 at different scan rates at 298 K (Figures S1–S3).

2.3.2. Preparation of PCs/CPE

The nitrogen-doped porous carbon materials (PCs) were prepared by calcination. A number of COFs were calcined at 800 °C (heating rate of 10 °C min−1) for 80 min under nitrogen flow in the tubular furnace, and black carbonized materials were obtained after cooling to room temperature naturally. To prepare the PCs modified electrode, 3.0 mg PCs were ultrasonically dispersed in 1.0 mL water for 1 h. The carbon paste electrode (CPE) was prepared by thoroughly mixing 1 mg of the graphite powder and 300 µL of liquid paraffin using an agate mortar and pestle for 40 min. Then, a suitable amount of the above mixture was filled firmly into one glass tube. The electrical contact was made by inserting a copper wire into the paste in the inner hole of the tube. Finally, 5.0 µL of 3.0 mg/mL PC suspension was dropped onto the freshly prepared CPE surface and dried at room temperature (PCs/CPE).

3. Results and Discussion

3.1. Electrochemical Sensing of Hydrazine

3.1.1. Characterization of Pd2+/COFs/GCE

The surface morphology and elemental composition of Pd2+/COFs/GCE were investigated by SEM and XPS. As shown in Figure 1, the SEM image of Pd2+/COFs/GCE shows the rough and inter-connected porous surface. The morphology is mainly composed of complicated long strips, indicating that COFs were modified on the GCE surface. XPS was also carried out to explore the surface composition and chemical states (Figure 2). The characteristic peaks of C, N, O, and Pd elements are found in the XPS survey spectrum (Figure S4). The O1s XPS spectrum exhibits four peaks centered at 530.79, 531.96, 533.18, and 535.29 eV (Figure 2a), which can be attributed to C−OH, C=O, C−O, or H−O, and chemisorbed oxygen in carboxylic groups or water (H-O-H), respectively [13]. The abundant oxo functionalities present on COFs can improve their dispersion in water, which is beneficial for the modification of COFs on the electrode surface. The N1s XPS spectra show two characteristic peaks at 400.06 and 403.53 eV (Figure 2c), which are assigned to the C=N and C-N groups, respectively [14]. Figure 2b displays the XPS Pd3d spectra of Pd2+/COFs, including two peaks at 338.10 and 343.40 eV, corresponding to the peak of the low energy band (Pd3d 3/2) and high energy band (Pd3d 5/2), respectively. The Pd2+ species might be attached to COFs via N coordination [15]. The results confirm the existence of Pd on the composite surface, and the Pd2+ is successfully loaded on the COFs. The XPS data showed that the atomic amount of Pd is 0.16%. In order to investigate the stability of COFs, the XRD patterns of COFs, Pd2+/COFs, and Pd2+/COFs soaked in 0.1 M NaOH were performed (Figure 3). The characteristic peaks of these samples observed in XRD patterns were in accord with the reported literature [15,16] and remained unchanged, indicating that Pd2+ modification and NaOH solution soaking did not change the crystalline structure. The results showed that COF-TpPa-1 possessed good chemical stability. It was also found that COF-TpPa-1 was still stable even though soaked in 1.0 mol/L HNO3, 6.0 mol/L NaOH, and HCl for three days [16].

3.1.2. Electrochemical Oxidation of Hydrazine

Figure 4 shows the CVs of COFs/GCE, Pd2+/GCE, and Pd2+/COFs/GCE in 0.1 M NaOH in the absence and presence of 5.0 mM hydrazine at a scan rate of 100 mV s−1. At COFs/GCE, the hydrazine oxidation began at the potential of 0.25 V (Figure 4a) and reached the oxidation peak at about 0.75 V. However, for Pd2+/GCE, two oxidation peaks were observed at −0.25 and 0.75 V, respectively (Figure 4b). Pd2+ was reduced to Pd metal by electroless deposition when GCE was immersed in a Pd2+ solution [17]. The first oxidation wave was ascribed to the electrocatalytic activity of Pd, while the second one originated from the bare GCE. Pd2+/COFs/GCE showed three oxidation peaks at −0.50, −0.33, and 0 V (Figure 4c), and no oxidation peak was found at about 0.75 V. This indicates that Pd2+/COFs/GCE demonstrated the lowest oxidation over-potential among these modified electrodes, suggesting that the combination of Pd2+ and COFs formed new catalytic sites and thus enhanced the oxidation of hydrazine. The different electrodes were also investigated by EIS (Figure S5). The electron transfer rate of different electrodes follows the order: Bare GCE > Pd2+/COFs/GCE > COFs/GCE. The higher charge transfer resistance of COFs/GCE is attributed to the lower electrical conductivity.

3.1.3. Analytical Performances for Hydrazine

Figure 5 shows the amperometric current–time response of Pd2+/COFs/GCE to the successive additions of hydrazine in 0.1 M NaOH. Upon each successive addition of different concentrations of hydrazine, a stair-shaped slot was raised rapidly and then reached a steady-state oxidation current within 2 s. The inset (a) of Figure 5 shows a linear response over a concentration range from 0.5 μM to 1.6 mM with a sensitivity of 10 μA μM−1 cm−2 and a detection limit of 0.2 μM (S/N = 3). Table S1 summarizes the sensing performances of hydrazine with Pd2+/COFs/GCE and other modified electrodes presented in the literature. The results show that Pd2+/COFs/GCE demonstrated an excellent analytical performance for hydrazine, owing to the high electrocatalytic activity.
In order to explore the selectivity of Pd2+/COFs/GCE, the interference test was also examined at Pd2+/COFs/GCE with 50.0 μM hydrazine in 0.1 M NaOH containing 0.5 mM of the possible interference substances such as KCl, NaCl, Na2SO4, NaBr, CaCl2, NaAc, glucose, (NH4)2SO4, NaNO3, and NaNO2 (Figure S6). It was found that 100-fold interferents almost did not interfere with the determination of hydrazine. Moreover, a remarkable current response with the addition of hydrazine was observed again after adding the mentioned interference species. The results indicate that our proposed Pd2+/COFs/GCE sensor displays excellent selectivity towards the electrochemical oxidation of hydrazine.
The long-term stability of the sensor was further tested by a continuous operation. After 1000s of continuous i-t running for 10.0 μM hydrazine, the electrode still retained 89% of its initial value. In addition, ten successive analyses of 10.0 μM hydrazine using the same electrode showed a relative standard deviation (RSD) value of 4.5%. A reproducible experiment was also evaluated by using five parallel electrodes for the same solution, and the RSD of the current response was found to be 6.4%, suggesting an acceptable reproducibility.

3.1.4. Sample Analysis of Hydrazine

To confirm the practical applicability of the proposed sensor, the validity of Pd2+/COFs/GCE for the determination of hydrazine spiked in real samples was tested by a standard addition method (Figure S7). The prepared sensor presented acceptable recoveries of hydrazine in lake water samples (Table S2).

3.2. Electrochemical Sensing of ONP and PNP

3.2.1. Electrochemical Behaviors of ONP and PNP

The electrochemical behaviors of ONP and PNP were investigated at bare GCE (a), COFs/GCE (b), and Pd2+/COFs/GCE (c) by CV at a scan rate of 50 mV s−1 in 0.1 M PBS (pH = 7.0) (Figure 6). The CVs of these electrodes in 0.1 M PBS are shown in Figure S8. For PNP, three redox peaks were observed, with one main irreversible cathodic peak located at −0.79 V (R1) and a pair of reversible redox peaks centered at about 0.08 V (R2) and 0.14 V (O1) at bare GCE and COFs/GCE. Pd2+/COFs/GCE also exhibited the same redox peaks, but the peak of R1 shifted to −0.56 V. However, at bare GCE and COFs/GCE, upon the addition of ONP, four redox peaks, including the three peaks and an additional oxidation peak (O2) appeared at −0.68 (R1), −0.30 (R2), −0.25 (O1), and −0.23 V (O2), respectively. Pd2+/COFs/GCE demonstrated the same redox peaks but a much lower R1 peak (−0.54 V). This indicated that Pd2+/COFs/GCE presented superior electrocatalytic activity towards the R1 reduction process of ONP and PNP, which was ascribed to the combination of the high catalytic performance of Pd2+ and efficient supporting effect of COFs.
When the potential was scanned negatively, both ONP and PNP could be reduced to hydroxylaminophenol through a four-electron and four-proton transfer process (R1). Reversible transformation (O1/R2) between hydroxylaminophenol and nitrosophenol then occurred with the potential scanning back. In addition, the phenolic group of ONP gave further oxidation through a proton and an electron transfer process (O2). The detailed reaction mechanism was proposed as Scheme 1 [18]:
To further investigate the electrochemical performance of Pd2+/COFs/GCE, the effect of the scan rate on the redox peak current (Ipc) was studied (Figure S9). The results showed that the peak currents of ONP and PNP were quadratic functions of the square root of the scan rate (ν1/2), which suggested that the reactions of ONP and PNP on the surface of Pd2+/COFs/GCE were mixed controlled processes (surface-controlled at lower scan rates but diffusion-controlled at higher scan rates).
Next, the effect of the pH on the cathodic peak current and potential (R1) of ONP and PNP (500 μM) was studied by DPV on the Pd2+/COFs/GCE in 0.1 M PBS (Figure S10). The reduction peak current of both ONP and PNP increased firstly and then decreased and reached the maximum at pH = 7.0. In the basic solution, the protonation of PNP and ONP decreases, and the electron density of the nitro group increases due to the decrease in proton concentration, thus hampering the electrochemical reduction of PNP and ONP [19]. Therefore, pH 7.0 was selected for further electrochemical experiments.

3.2.2. Analytical Performances for ONP and PNP

Under optimal conditions, the DPV response of Pd2+/COFs/GCE for different concentrations of ONP and PNP in 0.1 M PBS (pH = 7.0) was investigated, respectively (Figure 7a,b). The reduction peak current increased gradually with the concentration of ONP and PNP and exhibited a log–log linear relationship in the range of 5.0 μM–2.0 mM. The detection limit of ONP and PNP was estimated to be 1.6 and 1.3 μM (S/N = 3), respectively. Pd2+/COFs/GCE was also applied for the simultaneous detection of ONP and PNP. Figure 7c clearly shows the well-separated reduction peak for ONP and PNP. The calibration curves reveal the favorable linear relationships between the logarithm of peak current and the logarithm of concentrations (5.0 μM–2.0 mM), with 1.8 and 0.9 μM for ONP and PNP, respectively. The electrochemical sensing performances based on Pd2+/COFs/GCE and other reported sensors are summarized in Table S3. The results show that the Pd2+/COFs/GCE sensor exhibits comparable performances with the reported methods for the detection of ONP and PNP.
The antifouling property of different electrodes was investigated by successive CVs in the ONP and PNP solution, respectively (Figures S11–S13). The results showed that COFs/GCE exhibited the best antifouling property because of the hydrophilicity of COFs. For ONP, Pd2+/COFs/GCE showed improved antifouling properties than GCE in the first six cycles. However, for PNP, both Pd2+/COFs/GCE and bare GCE exhibited lower antifouling properties.

3.3. Electrochemical Sensing of GSH

3.3.1. Characterization of PC

The morphology of the PCs was characterized by SEM and TEM images. The SEM image of PCs shows the overlapping structure of stratified lamellae (Figure S14a). As shown in Figure S14b, the flake morphology can be observed in the TEM image of PCs. Furthermore, XPS characterization was employed to investigate the elemental compositions and chemical states of PCs. Figure 8 shows the XPS spectra of PCs containing C, N, and O elements. The C 1s spectra suggest the coexistence of C-C (283.60 eV), C=C (284.50 eV), C=O (287.70 eV), and π-π (292.30 eV) in PCs (Figure 8a) [20]. The N1s XPS spectra show two characteristic peaks at 397.10 and 400.00 eV (Figure 8b), which are attributed to the C=N and C-N groups, respectively [10]. Figure 8c displays the O 1s XPS spectrum of PCs, including two peaks at 531.30 and 532.20 eV, representing the C=O and C-O bonds, respectively [21]. The XPS spectra suggest that a large number of functional groups are present on the surface of the PCs.

3.3.2. Electrochemical Oxidation of GSH

GSH is an important antioxidant in protecting cells from oxidative damage by oxidizing to form glutathione disulfifide (GSSG), and the balance of GSH and GSSG acts as an indicator of many human illnesses such as several cancers, Parkinson’s disease, Alzheimer’s, and HIV [22]. Metal and non-metal electrocatalysts have been explored for the non-enzymatic detection of biological thiols [23,24,25]. Non-metal electrocatalysts consist of redox-active organic molecules [23,26] and carbon materials, including carbon nanotube [27], ordered mesoporous carbon [28], graphene oxide [29], graphene oxide nanoribbons [30], and derived carbon [31]. In fact, the oxo-functionalities present on these carbons contribute to the electrocatalytic activity towards thiols. Recently, nitrogen-doped carbon has proven effective in not only improving the electrical conductivity but also promoting defect sites in the carbon network for better catalytic activity towards GSH [32]. Here, the N-doped PCs with abundant oxo-functionalities were explored for the electro-oxidation of GSH. Figure 9 shows the CV response for the electrochemical oxidation of GSH at PCs/CPE in 0.1 M PBS (pH 7.0) at a scan rate of 50 mV s−1. As can be seen, in the absence of GSH, no anodic peak was observed on the PCs/CPE. However, in the presence of GSH, it could be noticed that a remarkable anodic current emerged at about 0.10 V, and the anodic peak was observed at 0.25 V, which could be ascribed to some oxygen-containing functional groups on PCs promoting the oxidation of GSH [26]. Furthermore, there was a very small anodic peak at 0.28 V, and this could be attributed to the edge plane-defective sites on PCs [33]. Furthermore, no cathodic current peak appeared in the CV curves, confirming that the GSH oxidation was a fully irreversible process. The results suggest that the N-doping and the oxo-functionalities present on PCs enhance the rate of electron transfer and exhibit efficient electrocatalytic activity for the oxidation of GSH.
The pH value is a crucial factor for the electrochemical oxidation of GSH, which was examined in 0.1 M PBS with various pH values containing 5.0 mM GSH on the PCs/CPE (Figure S15). The anodic peak current of GSH increased firstly and then decreased as the solution pH value increased from 3.0 to 5.0 and reached a maximum at pH = 4.0. Accordingly, pH = 4.0 is selected as the optimal condition for further detection. GSH is an important tripeptide consisting of L-cysteine, L-glutamine, and glycine with pKa1 = 2.12 (carboxylic acid of glutaminic acid), pKa2 = 3.59 (carboxylic acid of glycine), pKa3 = 8.75 (thiol of cysteine), and pKa4 = 9.65 (ammonium of glutaminic acid) [34]. Gilbertson et al. investigated the interaction and catalytic oxidation mechanisms between GO and GSH. It was found that the synergism between the adjacent epoxide and hydroxyl groups on the GO surface contributed to the catalytic oxidation of GSH [35]. It was speculated that the pH value of the solution is related to both the activity of oxo-functionalities and the ionization of GSH.

3.3.3. Analytical Performances for GSH

Under the optimal conditions, the amperometric responses of the PCs/CPE were conducted by successive injections of various concentrations of GSH into the stirred 0.1 M PBS (pH = 4.0) solution (Figure 10). Upon each addition of 200 μM GSH, a well stair-shaped amperometric response was raised quickly and reached a steady-state oxidation current within 5 s. The inset (a) of Figure 10 shows the response current of the PCs/CPE to various concentrations of GSH, and it is clear that the current increased rapidly with the addition of GSH. The enhancement of response current is ascribed to the excellent electrocatalytic property of the PCs/CPE. Figure 10b displays the linear correlation between the GSH oxidation current and concentration ranging from 5.0 μM to 1.2 mM and 1.2 to 6.4 mM, respectively. The limit of detection was calculated to be 2 μM (S/N = 3). Nevertheless, as the GSH concentration increased gradually, the sensitivity of the prepared sensor slightly decreased owing to the binding of the sulfur moiety to the electrode surface caused by the oxidation of GSH [26].
The current responses of the PCs/CPE to various concentrations of GSH in the static 0.1 M PBS (pH = 4.0) were also conducted using the amperometric technique. The inset of Figure S16 shows that the amperometric current increased linearly with the concentrations of GSH in a wide range from 5.0 μM to 10.0 mM. The limit of detection was also calculated to be 1.0 μM (S/N = 3). Table S4 summarizes the electrochemical sensing performances of the PCs/CPE and other reported sensors. The results indicate that the PCs/CPE exhibits excellent performances compared with other reported electrodes for the detection of GSH.

3.3.4. Sample Analysis of GSH

To further assess the practical applications of the proposed PCs/CPE biosensor, the validity of PCs/CPE for the determination of GSH in spiked ample was investigated by both successive and non-successive addition methods (Figures S17 and S18). The recoveries of GSH ranged from 92% to 99% and 92% to 107%, respectively (Tables S5 and S6).

4. Conclusions

In this work, COFs TpPa-1 was explored as an efficient supporting material and precursor to construct electrocatalytic interfaces for electrochemical sensing, and environmental pollutants and biological thiol were sensitively detected. COFs TpPa-1 possesses abundant oxo-functionalities and nitrogen elements, which endow good water dispersibility, powerful adsorption ability, and enhanced electrical conductivity. The results show that COFs TpPa-1 will be a potential material in synthesizing functional materials for electrocatalysis and electroanalysis.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano12172953/s1, Figure S1: The CVs of COFs/GCE in 0.1 M KCl solution in the presence of 5 mM K3Fe(CN)6 at different scan rates; Figure S2: The CVs of Pd2+/COFs/GCE in 0.1 M KCl solution in the presence of 5 mM K3Fe(CN)6 at different scan rates; Figure S3: The CVs of PCs/GCE in 0.1 M KCl solution in the presence of 5 mM K3Fe(CN)6 at different scan rates; Figure S4: XPS spectra of Pd2+/COFs; Figure S5: EIS of PCs /GCE, Bare GCE, COFs/GCE, and Pd2+/COFs/GCE in 0.1 M KCl solution in the presence of 5 mM K3Fe(CN)6; Figure S6: Amperometric curve of Pd2+/COFs/GCE with 50 μM hydrazine and 500 μM of interferences such as KCl, NaCl, Na2SO4, NaBr, CaCl2, NaAc, glucose, (NH4)2SO4, NaNO3, and NaNO2. (Electrolyte: 0.1 M NaOH); Figure S7: Amperometric curve of Pd2+/COFs/GCE for the detection of hydrazine spiked in water samples at −0.1 V. (Electrolyte: 0.1 M NaOH); Figure S8: CVs of Bare GCE, COFs/GCE, and Pd2+/COFs/GCE in 0.1 M PBS at a scan rate of 50 mV s−1 (pH = 7.0); Figure S9: CVs of 500 μM of (a) ONP and (b) PNP on Pd2+/COFs/GCE with different scan rates (20–200 mV s−1) in 0.1 M PBS (PH = 7.0). (Inset: the corresponding calibration plot.); Figure S10: DPV curves of 500 μM of (a) ONP and (b) PNP on Pd2+/COFs/GCE in 0.1 M PBS (pH = 7.0) with different pH values (1–5: 6.5, 7.0, 7.5, 8.0, and 8.5). Plots of Ipc vs. pH for (c) ONP and (d) PNP; Figure S11: Successive CVs of Bare GCE (a), COFs/GCE (b), and Pd2+/COFs/GCE (c) in the presence of 500 μM nitrophenol ONP at a scan rate of 50 mV s−1 in 0.1 M PBS (pH = 7); Figure S12: Successive CVs of Bare GCE (a), COFs/GCE (b), and Pd2+/COFs/GCE (c) in the presence of 500 μM nitrophenol PNP at a scan rate of 50 mV s−1 in 0.1 M PBS (pH = 7.0); Figure S13: Normalized peak current of bare GCE, COFs/GCE, and Pd2+/COFs/GCE in the presence of 500 μM nitrophenol (a: ONP, b: PNP) at a scan rate of 50 mV s−1 in 0.1 M PBS (pH = 7.0) for ten cycles; Figure S14: SEM (a) and TEM (b) images of PCs; Figure S15: CVs of 5 mM of GSH on (a) PCs/CPE in 0.1 M PBS with different pH values (1–5: 3.0, 3.5, 4.0, 4.5 and 5.0). Plots of Ipc vs. pH for (b) GSH; Figure S16: Amperometric responses of PCs/CPE to additions of various concentrations of GSH in 0.1 M PBS (pH = 4.0) at 0.25 V. (Inset: the corresponding calibration plot.); Figure S17: Amperometric responses of PCs/CPE to successive additions of various concentrations of GSH spiked in serum in 0.1 M PBS (pH = 4.0) at 0.25 V; Figure S18: Amperometric responses of PCs/CPE for the detection of GSH spiked in human serum in 0.1 M PBS (pH = 4.0) at 0.25 V; Table S1: Comparison of the Pd2+/COFs/GCE for different parameters with previously reported hydrazine electrodes; Table S2: Determination of hydrazine spiked in water samples; Table S3: Comparison of different modified GCE for detection of ONP and PNP; Table S4; Comparison of different modified electrodes for electrochemical sensing of GSH; Table S5; Determination of GSH spiked in human serum by successive additions;Table S6; Determination of GSH spiked in human serum by a non-successive addition method. References [36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68] were cited in the Supplementary Materials.

Author Contributions

Conceptualization, B.Y.; Data curation, B.Y., G.L. and C.X.; Formal Analysis, B.Y. and G.L.; Funding acquisition, B.Y.; Investigation, G.L., S.C. and L.G.; Methodology, B.Y.; Project administration, B.Y. and C.X.; Supervision, B.Y.; Validation, B.Y.; Visualization, B.Y. and G.L.; Writing—original draft, B.Y. and G.L.; Writing—review & editing, B.Y. and C.X. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Foundation of Shandong Province, China (No.ZR2020MB058).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The authors confirm that the data supporting the findings of this study are available within the article [and/or] its supplementary materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liu, H.Y.; Weng, L.Y.; Yang, C. A review on nanomaterial-based electrochemical sensors for H2O2, H2S and NO inside cells or released by cells. Microchim. Acta 2017, 184, 1267–1283. [Google Scholar] [CrossRef]
  2. Liu, C.S.; Li, J.J.; Pang, H. Metal-organic framework-based materials as an emerging platform for advanced electrochemical sensing. Coordin. Chem. Rev. 2020, 410, 213222. [Google Scholar] [CrossRef]
  3. Yan, Y.; He, T.; Zhao, B.; Qi, K.; Liu, H.F.; Xia, B.Y. Metal/covalent-organic frameworks-based electrocatalysts for water splitting. J. Mater. Chem. A 2018, 6, 15905–15926. [Google Scholar] [CrossRef]
  4. Yuan, B.Q.; Liu, D.; Yin, H.J.; Zhang, D.J. Materials for Electroanalysis Based on Advanced Frameworks. Front. Chem. 2021, 3, 638338. [Google Scholar] [CrossRef]
  5. Chuang, C.H.; Kung, C.W. Metal-organic frameworks toward electrochemical sensors: Challenges and opportunities. Electroanalysis 2020, 32, 1885–1895. [Google Scholar] [CrossRef]
  6. Lu, M.X.; Deng, Y.J.; Luo, Y.; Lv, J.P.; Li, T.B.; Xu, J.; Chen, S.W.; Wang, J.Y. Graphene aerogel–metal–organic framework-based electrochemical method for simultaneous detection of multiple heavy-metal ions. Anal. Chem. 2019, 91, 888–895. [Google Scholar] [CrossRef] [PubMed]
  7. Sun, Y.F.; Waterhouse, G.I.; Xu, L.H.; Qiao, X.G.; Xu, Z.X. Three-dimensional electrochemical sensor with covalent organic framework decorated carbon nanotubes signal amplification for the detection of furazolidone. Sens. Actuators B Chem. 2020, 321, 128501. [Google Scholar] [CrossRef]
  8. He, W.H.; Ifraemov, R.; Raslin, A.; Hod, I. Room-Temperature Electrochemical Conversion of Metal-Organic Frameworks into Porous Amorphous Metal Sulfides with Tailored Composition and Hydrogen Evolution Activity. Adv. Funct. Mater. 2018, 28, 1707244. [Google Scholar] [CrossRef]
  9. Beitollahi, H.; Movahedifar, F.; Tajik, S.; Jahani, S. A review on the effects of introducing CNTs in the modification process of electrochemical sensors. Electroanalysis 2019, 31, 1195–1203. [Google Scholar] [CrossRef]
  10. Hu, X.W.; Long, Y.; Fan, M.Y.; Yuan, M.; Zhao, H.; Ma, J.T.; Dong, Z.P. Two-dimensional covalent organic frameworks as self-template derived nitrogen-doped carbon nanosheets for eco-friendly metal-free catalysis. Appl. Catal. B Environ. 2019, 244, 25–35. [Google Scholar] [CrossRef]
  11. Yang, L.; Cai, Z.; Hao, L.; Ran, L.L.; Xu, X.; Dai, Y.; Pan, S.Y.; Jing, B.J.; Zhou, J.L. Increase of structural defects by N doping in MoS2 cross-linked with N-doped CNTs/carbon for enhancing charge transfer in oxygen reduction. Electrochim. Acta 2018, 283, 448–458. [Google Scholar] [CrossRef]
  12. Sun, F.; Liu, X.Y.; Wu, H.B.; Wang, L.J.; Gao, J.H.; Li, H.X.; Lu, Y.F. In situ high-level nitrogen doping into carbon nanospheres and boosting of capacitive charge storage in both anode and cathode for a high-energy 4.5 V full-carbon lithium-ion capacitor. Nano Lett. 2018, 18, 3368–3376. [Google Scholar] [CrossRef] [PubMed]
  13. Wang, J.; Yin, D.G.; Guo, X.D.; Luo, Z.Y.; Tao, L.Y.; Ren, J.J.; Zhang, Y. Fabrication of a Covalent Organic Framework-Based Heterojunction via Coupling with ZnAgInS Nanosphere with High Photocatalytic Activity. Langmuir 2022, 38, 4680–4691. [Google Scholar] [CrossRef]
  14. Zhong, X.; Lu, Z.P.; Liang, W.; Hu, B.W. The magnetic covalent organic framework as a platform for high-performance extraction of Cr (VI) and bisphenol a from aqueous solution. J. Hazard. Mater. 2020, 393, 122353. [Google Scholar] [CrossRef] [PubMed]
  15. Wang, F.N.; Zhang, J.X.; Shao, Y.H.; Jiang, H.; Liu, Y.F.; Chen, R.Z. Pd nanoparticles loaded on two-dimensional covalent organic frameworks with enhanced catalytic performance for phenol hydrogenation. Ind. Eng. Chem. Res. 2020, 59, 18489–18499. [Google Scholar] [CrossRef]
  16. Li, Z.D.; Zhang, H.Q.; Xiong, X.H.; Luo, F. U (VI) adsorption onto covalent organic frameworks-TpPa-1. J. Solid State Chem. 2019, 277, 484–492. [Google Scholar] [CrossRef]
  17. Choi, H.C.; Shim, M.; Bangsaruntip, S. Spontaneous reduction of metal ions on the sidewalls of carbon nanotubes. J. Am. Chem. Soc. 2002, 124, 9058–9059. [Google Scholar] [CrossRef]
  18. Wang, Q.Z.; Li, R.; Zhao, Y.J.; Zhe, T.T.; Bu, T.; Liu, Y.N.; Sun, X.Y.; Hu, H.F.; Zhang, M.; Zheng, X.H. Surface morphology-controllable magnetic covalent organic frameworks: A novel electrocatalyst for simultaneously high-performance detection of p-nitrophenol and o-nitrophenol. Talanta 2020, 219, 121255. [Google Scholar] [CrossRef]
  19. Afzali, F.; Zavar, M.H.A.; Rounaghi, G.; Ashraf, N.J.E.A. Gold digital versatile disc platform modified with nano-porous mercury/gold amalgam as a solid-state disposable electrochemical sensor for detection of para-nitrophenol. Electrochim. Acta 2016, 209, 654–660. [Google Scholar] [CrossRef]
  20. Chen, X.; Wang, X.; Fang, D. A review on C1s XPS-spectra for some kinds of carbon materials. Fuller. Nanotub. Carbon Nanostruct. 2020, 28, 1048–1058. [Google Scholar] [CrossRef]
  21. Cao, S.; Yang, J.X.; Li, J.; Shi, K.; Li, X.K. Preparation of oxygen-rich hierarchical porous carbon for supercapacitors through the co-carbonization of pitch and biomass. Diamond Relat. Mater. 2019, 96, 118–125. [Google Scholar] [CrossRef]
  22. Townsend, D.M.; Tew, K.D.; Tapiero, H.J.B. The importance of glutathione in human disease. Biomed. Pharmacother. 2003, 57, 145–155. [Google Scholar] [CrossRef]
  23. Harfield, J.C.; Batchelor-McAuley, C.; Compton, R.G. Electrochemical determination of glutathione: A review. Analyst 2012, 137, 2285–2296. [Google Scholar] [CrossRef] [PubMed]
  24. Hanko, M.; Švorc, Ľ.; Planková, A.; Mikuš, P. Overview and recent advances in electrochemical sensing of glutathione—A review. Anal. Chim. Acta 2019, 1062, 1–27. [Google Scholar] [CrossRef] [PubMed]
  25. Nesakumar, N.; Berchmans, S.; Alwarappan, S. Chemically modified carbon based electrodes for the detection of reduced glutathione. Sens. Actuators B Chem. 2018, 264, 448–466. [Google Scholar] [CrossRef]
  26. Yuan, B.Q.; Xu, C.Y.; Zhang, R.C.; Lv, D.H.; Li, S.J.; Zhang, D.J.; Liu, L.; Fernandez, C. Bioelectronics, Glassy carbon electrode modified with 7, 7, 8, 8-tetracyanoquinodimethane and graphene oxide triggered a synergistic effect: Low-potential amperometric detection of reduced glutathione. Biosens. Bioelectron. 2017, 96, 1–7. [Google Scholar] [CrossRef]
  27. Lee, P.; Compton, R.G. Selective electrochemical detection of thiol biomarkers in saliva using multiwalled carbon nanotube screen-printed electrodes. Sens. Actuators B Chem. 2015, 209, 983–988. [Google Scholar] [CrossRef]
  28. Zhou, M.; Ding, J.; Guo, L.P.; Shang, Q.K. Electrochemical behavior of L-cysteine and its detection at ordered mesoporous carbon-modified glassy carbon electrode. Anal. Chem. 2007, 79, 5328–5335. [Google Scholar] [CrossRef]
  29. Yuan, B.Q.; Zeng, X.Y.; Xu, C.Y.; Liu, L.; Ma, Y.H.; Zhang, D.J.; Fan, Y. Electrochemical modification of graphene oxide bearing different types of oxygen functional species for the electro-catalytic oxidation of reduced glutathione. Sens. Actuators B Chem. 2013, 184, 15–20. [Google Scholar] [CrossRef]
  30. Wu, S.; Lan, X.Q.; Huang, F.F.; Luo, Z.Z.; Ju, H.X.; Meng, C.G.; Duan, C.Y. Selective electrochemical detection of cysteine in complex serum by graphene nanoribbon. Biosens. Bioelectron. 2012, 32, 293–296. [Google Scholar] [CrossRef]
  31. Zhang, T.; Li, H.D.; Liu, M.X.; Zhou, H.; Zhang, Z.C.; Yu, C.; Wang, C.Y.; Wang, G.X. Improved the specificity of peroxidase-like carbonized polydopamine nanotubes with high nitrogen doping for glutathione detection. Sens. Actuators B Chem. 2021, 341, 129987. [Google Scholar] [CrossRef]
  32. Wang, Y.; de Carvalho, N.A.; Tan, S.S.; Gilbertson, L.M. Leveraging electrochemistry to uncover the role of nitrogen in the biological reactivity of nitrogen-doped graphene. Environ. Sci. Nano 2019, 6, 3525–3538. [Google Scholar] [CrossRef]
  33. Ndamanisha, J.C.; Bai, J.; Qi, B.; Guo, L.P. Application of electrochemical properties of ordered mesoporous carbon to the determination of glutathione and cysteine. Anal. Biochem. 2009, 386, 79–84. [Google Scholar] [CrossRef] [PubMed]
  34. Gotoh, T.; Iguchi, H.; Kikuchi, K.I. Separation of glutathione and its related amino acids by nanofiltration. Biochem. Eng. J. 2004, 19, 165–170. [Google Scholar] [CrossRef]
  35. Wang, Y.; Basdogan, Y.; Zhang, T.Y.; Lankone, R.S.; Wallace, A.N.; Fairbrother, D.H.; Keith, J.A.; Gilbertson, L.M. Interfaces, Unveiling the synergistic role of oxygen functional groups in the graphene-mediated oxidation of glutathione. ACS Appl. Mater. Interfaces 2020, 12, 45753–45762. [Google Scholar] [CrossRef]
  36. Wang, L.; Teng, Q.Q.; Sun, X.T.; Chen, Y.T.; Wang, Y.M.; Wang, H.; Zhang, Y.F. Facile synthesis of metal-organic frameworks/ordered mesoporous carbon composites with enhanced electrocatalytic ability for hydrazine. J. Colloid Interface Sci. 2018, 512, 127–133. [Google Scholar] [CrossRef]
  37. Wang, L.; Meng, T.J.; Jia, H.X.; Feng, Y.; Gong, T.; Wang, H.; Zhang, Y.F. Electrochemical study of hydrazine oxidation by leaf-shaped copper oxide loaded on highly ordered mesoporous carbon composite. J Colloid Interface Sci. 2019, 549, 98–104. [Google Scholar] [CrossRef]
  38. Gharani, M.; Bahari, A.; Ghasemi, S. Preparation of MoS2-reduced graphene oxide/Au nanohybrid for electrochemical sensing of hydrazine. J. Mater. Sci. Mater. Electron. 2021, 32, 7765–7777. [Google Scholar] [CrossRef]
  39. Yao, Y.Q.; Han, X.H.; Yang, X.H.; Zhao, J.; Chai, C.P. Detection of Hydrazine at MXene/ZIF-8 Nanocomposite Modified Electrode. Chin. J. Chem. 2021, 39, 330–336. [Google Scholar] [CrossRef]
  40. Zhang, X.J.; Zheng, J.B. Amperometric hydrazine sensor based on the use of a gold nanoparticle-modified nanocomposite consisting of porous polydopamine, multiwalled carbon nanotubes and reduced graphene oxide. Mikrochim. Acta. 2020, 187, 89. [Google Scholar] [CrossRef]
  41. Vishnu, N.; Kumar, A.S.; Badhulika, S. Selective in-situ derivatization of intrinsic nickel to nickel hexacyanoferrate on carbon nanotube and its application for electrochemical sensing of hydrazine. J. Electroanal. Chem. 2019, 837, 60–66. [Google Scholar] [CrossRef]
  42. Li, S.; Feng, W.S.; Gao, X.H.; Guo, A.M.; Li, H.J. Copper-based materials derived from metal-organic frameworks for electrochemical sensing of hydrazine. Micro Nano Lett. 2021, 16, 478–483. [Google Scholar] [CrossRef]
  43. Asadi, F.; Azizi, S.N.; Ghasemi, S. Preparation of Ag nanoparticles on nano cobalt-based metal organic framework (ZIF-67) as catalyst support for electrochemical determination of hydrazine. J. Mater. Sci. Mater. Electron. 2019, 30, 5410–5420. [Google Scholar] [CrossRef]
  44. Avanes, A.; Hasanzadeh-Karamjavan, M.; Shokri-Jarcheloo, G. Electrocatalytic oxidation and amperometric determination of hydrazine using a carbon paste electrode modified with beta-nickel hydroxide nanoplatelets. Mikrochim Acta 2019, 186, 441. [Google Scholar] [CrossRef] [PubMed]
  45. Duan, C.Q.; Dong, Y.H.; Sheng, Q.L.; Zheng, J.B. A high-performance non-enzymatic electrochemical hydrazine sensor based on NiCo2S4 porous sphere. Talanta 2019, 198, 23–29. [Google Scholar] [CrossRef]
  46. Ghasemi, S.; Hosseini, S.R.; Hasanpoor, F.; Nabipour, S. Amperometric hydrazine sensor based on the use of Pt-Pd nanoparticles placed on reduced graphene oxide nanosheets. Mikrochim. Acta 2019, 186, 601. [Google Scholar] [CrossRef]
  47. Maleki, A.; Rezaee, R.; Daraei, H.; Shahmoradi, B.; Amini, N. Fabrication of a sensitive electrochemical sensor to environmental pollutant of hydrazine in real water samples based on synergistic catalysis of Ag@C core–shell and polyalizarin yellow R. J. Alloys Compd. 2018, 763, 997–1004. [Google Scholar] [CrossRef]
  48. Shahid, M.M.; Rameshkumar, P.; Basirunc, W.J.; Wijayantha, U.; Chiu, W.S.; Khiew, P.S.; Huang, N.M. An electrochemical sensing platform of cobalt oxide@gold nanocubes interleaved reduced graphene oxide for the selective determination of hydrazine. Electrochim. Acta 2018, 259, 606–616. [Google Scholar] [CrossRef]
  49. Mousavi-Majd, A.; Ghasemi, S.; Hosseini, S.R. Zeolitic imidazolate framework derived porous ZnO/Co3O4 incorporated with gold nanoparticles as ternary nanohybrid for determination of hydrazine. J. Alloys Compd. 2022, 896, 162922. [Google Scholar] [CrossRef]
  50. Vinoth, S.; Sampathkumar, P.; Giribabu, K.; Pandikumar, A. Ultrasonically assisted synthesis of barium stannate incorporated graphitic carbon nitride nanocomposite and its analytical performance in electrochemical sensing of 4-nitrophenol, Ultrasonics Sonochemistry. Ultrason. Sonochem. 2020, 62, 104855. [Google Scholar] [CrossRef]
  51. Zhou, Y.Y.; Zhao, J.; Li, S.H.; Guo, M.J.; Fan, Z. An electrochemical sensor for the detection of p-nitrophenol based on a cyclodextrin-decorated gold nanoparticle-mesoporous carbon hybrid. Analyst 2019, 144, 4400–4406. [Google Scholar] [CrossRef] [PubMed]
  52. Ansari, A.A.; Alam, M.; Ali, M.A. Nanostructured CeO2: Ag platform for electrochemically sensitive detection of nitrophenol. Colloids Surf. A. 2021, 613, 126116. [Google Scholar] [CrossRef]
  53. Suresh, R.; Giribabu, K.; Manigandan, R.; Kumar, S.P.; Munusamy, S.; Muthamizh, S.; Narayanan, V. Polyaniline Nanorods: Synthesis, Characterization, and Application for the Determination ofpara-Nitrophenol. Anal. Lett. 2015, 49, 269–281. [Google Scholar] [CrossRef]
  54. Alam, M.M.; Asiri, A.M.; Rahman, M.M. Electrochemical detection of 2-Nitrophenol using a glassy carbon electrode modified with BaO Nanorods. Chem. Asian J. 2021, 16, 1475–1485. [Google Scholar] [CrossRef] [PubMed]
  55. Nejati, K.; Asadpour-Zeynali, K.; Rezvani, Z.; Peyghami, R. Determination of 2-nitrophenol by electrochemical synthesized Mg/Fe layered double hydroxide sensor. Int. J. Electrochem. Sci. 2014, 9, 5222–5234. [Google Scholar]
  56. Yin, H.S.; Zhou, Y.L.; Ai, S.Y.; Cui, L.; Zhu, L.S. Electrochemical determination of 2-Nitrophenol in Water Samples using Mg-Al-SDS Hydrotalcite-Like Clay modified glassy carbon electrode. Electroanalysis 2010, 22, 1136–1142. [Google Scholar] [CrossRef]
  57. Li, J.H.; He, L.Z.; Jiang, J.B.; Xu, Z.F.; Liu, M.Q.; Liu, X.; Tong, H.X.; Liu, Z.; Qian, D. Facile syntheses of bimetallic Prussian blue analogues (KxM [Fe(CN)6]•nH2O, M = Ni, Co, and Mn) for electrochemical determination of toxic 2-nitrophenol. Electrochim. Acta 2020, 353, 136579. [Google Scholar] [CrossRef]
  58. Chu, L.; Han, L.; Zhang, X.L. Electrochemical simultaneous determination of nitrophenol isomers at nano-gold modified glassy carbon electrode. J. Appl. Electrochem. 2011, 41, 687–694. [Google Scholar] [CrossRef]
  59. Zhang, J.J.; Cui, S.Q.; Ding, Y.P.; Yang, X.X.; Guo, K.; Zhao, J.T. Bioelectronics, Two-dimensional mesoporous ZnCo2O4 nanosheets as a novel electrocatalyst for detection of o-nitrophenol and p-nitrophenol. Biosens. Bioelectron. 2018, 112, 177–185. [Google Scholar] [CrossRef]
  60. Hryniewicz, B.M.; Orth, E.S.; Vidotti, M. Enzymeless PEDOT-based electrochemical sensor for the detection of nitrophenols and organophosphates. Sens. Actuators B Chem. 2018, 257, 570–578. [Google Scholar] [CrossRef]
  61. Wu, W.T.; Chen, X.; Jiao, Y.T.; Fan, W.T.; Liu, Y.L.; Huang, W.H. Versatile Construction of Biomimetic Nanosensors for Electrochemical Monitoring of Intracellular Glutathione. Electroanal. Chem. 2022, 134, e202115820. [Google Scholar]
  62. Zhao, L.Z.; Zhao, L.; Miao, Y.; Zhang, C.X. Selective electrochemical determination of glutathione from the leakage of intracellular GSH contents in HeLa cells following doxorubicin-induced cell apoptosis. Electrochim. Acta 2016, 206, 86–98. [Google Scholar] [CrossRef] [Green Version]
  63. Lee, P.T.; Goncalves, L.M.; Compton, R.G. Electrochemical determination of free and total glutathione in human saliva samples. Sens. Actuators B Chem. 2015, 221, 962–968. [Google Scholar] [CrossRef]
  64. Yuan, B.Q.; Zhang, R.C.; Jiao, X.X.; Li, J.; Shi, H.Z.; Zhang, D.J. Amperometric determination of reduced glutathione with a new Co-based metal-organic coordination polymer modified electrode. Electrochem. Commun. 2014, 40, 92–95. [Google Scholar] [CrossRef]
  65. Kannappan, S.; Bhat, L.R.; Nesakumar, N.; Babu, K.J.; Kulandaisamy, A.J.; Rayappan, J.B.B. Design and development of a non-enzymatic electrochemical biosensor for the detection of Glutathione. Electroanalysis 2022, 34, 1–12. [Google Scholar] [CrossRef]
  66. Abbas, M.N.; Saeed, A.A.; Ali, M.B.; Errachid, A.; Zine, N.; Baraket, A.; Singh, B. Biosensor for the oxidative stress biomarker glutathione based on SAM of cobalt phthalocyanine on a thioctic acid modified gold electrode. J. Solid State Electrochem. 2019, 23, 1129–1144. [Google Scholar] [CrossRef]
  67. Xu, H.Y.; Xiao, J.J.; Liu, B.; Griveau, S.; Bedioui, F. Enhanced electrochemical sensing of thiols based on cobalt phthalocyanine immobilized on nitrogen-doped graphene. Biosens. Bioelectron. 2015, 66, 438–444. [Google Scholar] [CrossRef]
  68. Saranya, S.; Geetha, B.; Deepa, P.N. Simultaneous detection of glutathione, threonine, and glycine at electrodeposited RuHCF/rGO–modified electrode. Ionics 2019, 25, 5537–5550. [Google Scholar] [CrossRef]
Figure 1. SEM image of Pd2+/COFs/GCE.
Figure 1. SEM image of Pd2+/COFs/GCE.
Nanomaterials 12 02953 g001
Figure 2. XPS (a) O1s, (b) Pd3d and (c) N1s spectra of Pd2+/COFs.
Figure 2. XPS (a) O1s, (b) Pd3d and (c) N1s spectra of Pd2+/COFs.
Nanomaterials 12 02953 g002
Figure 3. XRD patterns of COFs, Pd2+/COFs and Pd2+/COFs soaked in 0.1 M NaOH for 24 h.
Figure 3. XRD patterns of COFs, Pd2+/COFs and Pd2+/COFs soaked in 0.1 M NaOH for 24 h.
Nanomaterials 12 02953 g003
Figure 4. CVs of COFs/GCE (a), Pd2+/GCE (b), Pd2+/COFs/GCE (c) in the absence (dotted line) and presence (solid line) of 5 mM hydrazine at a scan rate of 100 mV s−1, respectively. (Electrolyte: 0.1 M NaOH).
Figure 4. CVs of COFs/GCE (a), Pd2+/GCE (b), Pd2+/COFs/GCE (c) in the absence (dotted line) and presence (solid line) of 5 mM hydrazine at a scan rate of 100 mV s−1, respectively. (Electrolyte: 0.1 M NaOH).
Nanomaterials 12 02953 g004
Figure 5. Amperometric responses of the Pd2+/COFs/GCE to successive additions of various concentrations of hydrazine in 0.1 M NaOH at −0.1 V. (Inset: (a) the corresponding calibration plot, (b) the amplified response curve.).
Figure 5. Amperometric responses of the Pd2+/COFs/GCE to successive additions of various concentrations of hydrazine in 0.1 M NaOH at −0.1 V. (Inset: (a) the corresponding calibration plot, (b) the amplified response curve.).
Nanomaterials 12 02953 g005
Figure 6. CVs of bare GCE, COFs/GCE, Pd2+/COFs/GCE in the presence of 500 μM nitrophenol ((a): ONP, (b): PNP) at a scan rate of 50 mV s−1 in 0.1 M PBS (pH = 7.0).
Figure 6. CVs of bare GCE, COFs/GCE, Pd2+/COFs/GCE in the presence of 500 μM nitrophenol ((a): ONP, (b): PNP) at a scan rate of 50 mV s−1 in 0.1 M PBS (pH = 7.0).
Nanomaterials 12 02953 g006
Scheme 1. Electrochemical reaction mechanism of PNP and ONP at the Pd2+/COFs/GCE.
Scheme 1. Electrochemical reaction mechanism of PNP and ONP at the Pd2+/COFs/GCE.
Nanomaterials 12 02953 sch001
Figure 7. DPV of Pd2+/COFs/GCE with individual addition of (a) ONP and (b) PNP (5, 10, 20, 50, 100, 200, 500, 1000, and 2000 μM) into 0.1 M PBS (pH = 7.0), respectively. Inset: Linear relationship between logarithm of peak current and logarithm of concentrations of (a) ONP and (b) PNP. DPV of Pd2+/COFs/GCE with simultaneous addition of the same concentration of (c) ONP and PNP (5–2000 μM) into 0.1 M PBS (pH = 7.0). Inset: Linear relationship between logarithm of peak current and logarithm of concentrations of (c) ONP and PNP.
Figure 7. DPV of Pd2+/COFs/GCE with individual addition of (a) ONP and (b) PNP (5, 10, 20, 50, 100, 200, 500, 1000, and 2000 μM) into 0.1 M PBS (pH = 7.0), respectively. Inset: Linear relationship between logarithm of peak current and logarithm of concentrations of (a) ONP and (b) PNP. DPV of Pd2+/COFs/GCE with simultaneous addition of the same concentration of (c) ONP and PNP (5–2000 μM) into 0.1 M PBS (pH = 7.0). Inset: Linear relationship between logarithm of peak current and logarithm of concentrations of (c) ONP and PNP.
Nanomaterials 12 02953 g007
Figure 8. XPS (a) C, (b) N and (c) O spectra of PCs.
Figure 8. XPS (a) C, (b) N and (c) O spectra of PCs.
Nanomaterials 12 02953 g008
Figure 9. CV of PCs/CPE in the absence (dotted line) and presence (solid line) of 5 mM GSH in 0.1 M PBS (pH = 7.0). Scan rate: 50 mV s−1.
Figure 9. CV of PCs/CPE in the absence (dotted line) and presence (solid line) of 5 mM GSH in 0.1 M PBS (pH = 7.0). Scan rate: 50 mV s−1.
Nanomaterials 12 02953 g009
Figure 10. Amperometric responses of the PCs/CPE to various concentrations of GSH in 0.1 M PBS (pH = 4.0) at 0.25 V. (Inset: (a) the amplified response curve, (b) the calibration curve for the different concentrations of GSH).
Figure 10. Amperometric responses of the PCs/CPE to various concentrations of GSH in 0.1 M PBS (pH = 4.0) at 0.25 V. (Inset: (a) the amplified response curve, (b) the calibration curve for the different concentrations of GSH).
Nanomaterials 12 02953 g010
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, G.; Yuan, B.; Chen, S.; Gan, L.; Xu, C. Covalent Organic Frameworks-TpPa-1 as an Emerging Platform for Electrochemical Sensing. Nanomaterials 2022, 12, 2953. https://doi.org/10.3390/nano12172953

AMA Style

Li G, Yuan B, Chen S, Gan L, Xu C. Covalent Organic Frameworks-TpPa-1 as an Emerging Platform for Electrochemical Sensing. Nanomaterials. 2022; 12(17):2953. https://doi.org/10.3390/nano12172953

Chicago/Turabian Style

Li, Gang, Baiqing Yuan, Sidi Chen, Liju Gan, and Chunying Xu. 2022. "Covalent Organic Frameworks-TpPa-1 as an Emerging Platform for Electrochemical Sensing" Nanomaterials 12, no. 17: 2953. https://doi.org/10.3390/nano12172953

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