Next Article in Journal
Recent Progress in Multifunctional Gas Sensors Based on 2D Materials
Next Article in Special Issue
Single-Atom Nanomaterials in Electrochemical Sensors Applications
Previous Article in Journal / Special Issue
Recent Advances in Functionalization Strategies for Biosensor Interfaces, Especially the Emerging Electro-Click: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Novel Electrochemical Sensor Based on MnO2 Nanowire Modified Carbon Paper Electrode for Sensitive Determination of Tetrabromobisphenol A

School of the Chemistry and Life Sciences, Suzhou University of Science and Technology, Suzhou 215009, China
*
Author to whom correspondence should be addressed.
Submission received: 26 June 2023 / Revised: 17 July 2023 / Accepted: 16 August 2023 / Published: 1 September 2023

Abstract

:
In this paper, a MnO2 nanowire (MnO2-NW) modified carbon paper electrode (CP) was developed as a novel electrochemical sensor for the sensitive determination of tetrabromobisphenol A (TBBPA). The MnO2 nanowire was prepared by a hydrothermal synthesis method, and the morphology and structure of MnO2 were characterized using scanning electron microscopy, X-ray diffraction and X-ray photoelectron spectroscopy. The electrochemical performance of TBBPA on MnO2-NW/CP was investigated by cyclic voltammetry, and the result confirmed that MnO2-NW/CP exhibited excellent sensitivity for the determination of TBBPA due to the high specific surface area and good electrical conductivity of the nanowire-like MnO2. Moreover, the important electrochemical factors such as pH value, incubation time and modified material proportion were systematically studied to improve the determination sensitivity. The interferences from similar structure compounds on TBBPA have also been investigated. Under the optimal conditions, MnO2-NW/CP displayed a linear range of 70~500 nM for TBBPA with a detection limit of 3.1 nM. This was superior to some electrochemical methods in reference. The work presents a novel and simple method for the determination of TBBPA.

1. Introduction

Tetrabromobisphenol A (TBBPA), an organic compound with the chemical formula C15H12Br4O2, is widely used as a reactive flame retardant in electronic equipment, furniture, plastics and textile products [1,2]. TBBPA accounts for 80% of the global demand in Asia. As a result of its massive use, release from industrial products, as well as the migration of terrestrial and marine food chains, TBBPA is continuously released into the environment [3,4,5]. TBBPA is a potential persistent organic pollutant reported to be bioaccumulative and highly toxic. Chronic exposure to TBBPA can induce immune, reproductive and neurotoxic effects on animals [6,7,8]. In 2017, the World Health Organization’s International Agency for Research on Cancer published a preliminary list of carcinogens for reference, and tetrabromobisphenol A is on the list of carcinogens in category 2A. Therefore, detecting TBBPA quickly and accurately is important for environmental monitoring and the protection of human health.
Several analytical methods have been reported for the detection of TBBPA, including chromatographic [9,10], spectroscopic [11,12] and immunoassay methods [13,14]. For example, Zhang et al. [10] developed a high-performance liquid chromatography method coupled with triple quadrupole mass spectrometry (HPLC-MS/MS) with atmospheric pressure chemical ionization (APCI) source for simultaneous detection of TBBPA and its ten derivatives in determining complicated environmental samples, including sewage sludge, river water and vegetable samples. Feng et al. [12] explored a fluorescent sensor by coating a molecularly imprinted polymer layer on CdTe quantum dots for the sensitive determination of TBBPA. Although these methods have good selectivity and high detection sensitivity, they are limited by the high costs of instrument maintenance, sophisticated operating skills, and extensive organic solvent consumption. Many researchers pay attention to electrochemical sensors due to their high efficiency, low cost, prominent sensitivity, and rapid detection speed [15,16,17]. Presently, some electrochemical sensors have been reported for the detection of TBBPA in the environment. Nonetheless, some TBBPA electrochemical sensors’ utility is limited by the non-conductive film that forms on the electrode surface due to TBBPA’s electrochemical oxidation, impairing detection stability [18].
The materials used for electrode modification play a pivotal role in mitigating the passivation induced by TBBPA oxidation products. The conception and advancement of nanomaterials have considerably influenced electrochemical sensors, with their properties largely being a function of their microstructure and form. A variety of nanomaterials, such as metal oxides, carbon-based materials, metal-organic frameworks, and nanocomposites, are being studied for TBBPA detection. Zhou et al. [19] prepared a conductive composite of carbon nanotubes@zeolitic imidazole framework-67 (CNTs@ZIF-67), which possesses an excellent adsorption capacity (92.12 mg g−1) for TBBPA. The composite was used to modify an acetylene black electrode for the sensitive determination of TBBPA in spiked rain and pool water samples with the aid of perfluorodecanoic acid. The sensor is stable, reproducible, and has a linear range of 0.01–1.5 μM TBBPA concentration, with a 4.2 nM detection limit (at S/N = 3). Recently, carbon dots (CDs) have attracted a lot of attention due to their low toxicity, biocompatibility and good electrical conductivity, and are widely used in biosensors, photocatalysis, electrocatalysis and electrochemical sensors. Guo et al. [20] developed magnetic CDs composed of carbon dots (CDs) and Fe3O4 nanoparticles via an amination reaction. A glassy carbon electrode modified with magnetic CDs and cetyltrimethylammonium bromide (CTAB) was employed as an electrochemical sensor for TBBPA detection in beverages. Magnetic CDs facilitate TBBPA oxidation, and CTAB’s hydrophobic effect can enrich TBBPA. The combined impact of Magnetic CDs and CTAB boosts electrochemical sensor performance, indicating a linear range between 1 and 1000 nM, a detection limit of 0.75 nM, and displaying benefits such as rapidity, superior sensitivity, and robust stability. MXene, an exceptional electrode modification material, is a two-dimensional transition metal carbide obtained by etching aluminum in Ti3AlC2 with hydrofluoric acid. The combination of precious metals nanoparticles and Mxene can take advantage of the extraordinary electrocatalytic properties of noble metal nanomaterials and the specific surface area and electrical conductivity of MXene. Shao et al. [21] developed a TBBPA electrochemical sensor by modifying it to a glassy carbon electrode with the MXene/Au nanocomposite. The finalized sensor exhibited excellent linearity for TBBPA concentrations between 0.05 and 10 nM. It had a detection limit of 0.0144 nM and successfully detected TBBPA in water, with recoveries ranging from 97.1% to 106% for the added standards. Lu et al. [22] reported that the composite of graphitic carbon nitride (g-C3N4) and N-butylpyridinium hexafluorophosphate (NBH) could promote the oxidation of TBBPA. The detection of TBBPA was fulfilled using a g-C3N4-NBH modified carbon paste electrode in the range of 1 nM to 30 nM and 30 nM to 500 nM with a limit of 0.4 nM.
Various metal oxides, including CuO, Fe3O4 and Fe2O3, have extensive applications in electrochemical sensors due to their distinct morphology [23,24,25]. Zhou and colleagues [23] produced different forms of CuO nanomaterials, namely nanostrips, nanowires, and microspheres. These were combined with graphene nanoplates and used to modify the surface of glassy carbon electrodes for electrochemical detection of both glucose and TBBPA. The composite made from CuO nanostrips demonstrated the most significant active surface area, the least charge transfer resistance, and the highest detection sensitivity. The designed electrochemical sensor offered sensitive TBBPA detection within a linear range of 5 to 600 nM, with an anticipated detection limit of 0.73 nM based on a signal-to-noise ratio of three. Luo et al. [24] developed an Fe3O4-activated biochar using surplus sludge, leading to the successful creation of an electrochemical sensor for TBBPA detection. The results from electrochemical tests suggested that the Fe3O4–activated biochar film possessed a more substantial active surface area, reduced charge transfer resistance, and increased TBBPA accumulation efficiency. The sensor exhibited solid linearity for TBBPA concentrations ranging from 5 to 1000 nM, a reasonably low detection limit of 3.2 nM, and was effectively utilized to determine TBBPA in water samples. Zhang et al. [25] synthesized different morphologies of Fe2O3 nanomaterials such as nanoplate, nanorod and flower-like by a hydrothermal method. The electrochemical activity of Fe2O3 nanomaterial modified electrodes was found to be closely related to Fe2O3′s morphology. The composite of flower-like Fe2O3 (f-Fe2O3) and expanded graphite (EG) exhibited the largest electrochemical active area and the lowest electron transfer resistance. The f-Fe2O3/EG modified electrode displayed the best sensing performance of TBBPA with a detection limit of 1.23 nM.
Nano manganese dioxide (MnO2) was widely used as electrode material owing to its extraordinary electrochemical performance, low cost, abundant storage, non-toxic nature and simplicity in preparation. Currently, MnO2 nanomaterials show promising potential in solar cell devices, bioapplications, sensing, dye mineralization etc., because MnO2 can be easily tuned into desired structure and morphology [26]. Different morphological MnO2 nanoparticles prepared under specific conditions display varying electrochemical characteristics, boasting qualities such as excellent electrical conductivity, stability, and sensitivity [27,28,29]. Karami-Kolmoti et al. [28] exhibited a sensitive electrochemical sensor based on a manganese dioxide nanorods/graphene oxide nanocomposite for the determination of hydroquinone in water samples. Yu et al. [29] prepared novel manganese dioxide-graphene nanosheets (MnO2-GNSs) by a one-step hydrothermal method. A MnO2-GNSs modified glassy carbon electrode was employed for the sensitive detection of hydrogen peroxide. Yakubu et al. [30] introduced a novel competitive electrochemical immunosensor built on gold-palladium bimetallic nanoparticles for quick and sensitive detection of TBBPA in environmental water. The nanoparticles were successfully synthesized and modified with amine-functionalized nanoflower-like manganese oxide. Under the best conditions, the competitive sensors demonstrated superb performance, including commendable sensitivity (LOD, 0.10 ng/mL; S/N = 3) and satisfactory accuracy (recovery rate, 84–120%). This proposed method has been applied in the analysis of TBBPA from various water sources, demonstrating the significant potential for the sensitive detection of trace amounts of TBBPA in aquatic settings. However, a non-enzymatic electrochemical sensor for the detection of TBBPA based on a MnO2 nanowire modified electrode has not been reported.
In this work, an easy-to-use and sensitive electrochemical sensor, a MnO2 nanowire modified carbon paper electrode, was fabricated for the sensitive detection of TBBPA. Carbon paper (CP) is a kind of rigid material composed of a complex and tortuous fibrous structure with a resin [31]. It is well known that the electrochemical behavior of sensors depends on their surface properties. Compared with other conventional electrodes, carbon paper electrodes possess some unique characteristics, such as adjustable size, porosity, conductivity and mechanical strength [32,33]. In addition, the use of carbon paper electrodes eliminates the tedious electrode grinding of solid electrodes, making miniaturization and portability possible and advancing on-line and in-situ measurements for electrochemical testing. The effects of pH, the usage of MnO2, and the enrichment time were also studied to enhance the detection performance of MnO2-NW/CP for the detection of TBBPA.

2. Materials and Methods

2.1. Chemicals and Instruments

The carbon paper (CP) used in this research was purchased from Toray (Japan). TBBPA and Potassium ferricyanide (K3[Fe(CN)6]) were purchased from Aladdin Reagent Co. Ltd. (Shanghai, China). Potassium permanganate, sodium dihydrogen phosphate (NaH2PO4), dibasic sodium phosphate (Na2HPO4), N, N-Dimethylformamide (DMF), anhydrous ethanol and acetonitrile were purchased from Jiangsu Qiangsheng Functional Chemistry Co. Ltd. (Nanjing, China). All the chemicals were of analytical grade and used without further purification. All of the aqueous solutions were prepared with twice-distilled water throughout the whole experiment. The experimental temperatures were all at room temperature.
All electrochemical experiments, including differential pulse voltammetry (DPV) and cyclic voltammetry (CV), were carried out in an electrochemical workstation RST 3000 with a three-electrode system. CP or modified electrode was used as the working electrode, and Ag/AgCl and Pt wire were used as the reference and auxiliary electrodes, respectively.

2.2. Fabrication of Manganese Dioxide

Manganese dioxide was prepared by two methods.
Method 1: 0.728 g of potassium permanganate was added to a mixture solution of 80 mL H2O and 15 mL ethanol and then stirred at 25 °C for 2 h. After the reaction was completed, the product was washed 3 times with deionized water and anhydrous ethanol and dried in an oven at 60 °C for 12 h [34].
Method 2: 0.45 g of potassium permanganate was dissolved in 30 mL of 0.4 M acetic acid solution, transferred to a 50 mL Teflon-lined stainless-steel autoclave (Anhui Kemi Instruments Co., Hefei, China) and kept at 140 °C for 12 h. After the reaction was completed, the product was washed 3 times with deionized water and anhydrous ethanol and dried in an oven at 60 °C for 12 h.

2.3. Preparation of Modified Electrodes

The CP was cut into 1 × 0.5 cm and ultrasonically cleaned in ethanol for five minutes. The geometric area of the carbon paper electrode in solution was controlled to be 0.5 × 0.5 cm2 with the insulating adhesive, and the upper end of the carbon paper electrode was connected to the electrochemical workstation. To excite the electrochemical properties of the CP [35], the electrode was immersed in 0.1 M sulfuric acid solution and scanned for 50 cycles from −0.2 V to 1.0 V at 100 mV/s using the CV method. Then, the electrode was cleaned with deionized water and dried at room temperature.
A total of 4.0 mg of manganese dioxide was dispersed in 2.0 mL of DMF solution with the aid of an ultrasonic bath to obtain a suspension of 2 mg/mL of manganese dioxide. Next, 7.0 μL of the above suspension was dropped onto the surface of the treated carbon paper electrode and dried at room temperature.

3. Results and Discussion

3.1. Characterization of Manganese Dioxide

Scanning electron microscope (SEM) images were used to investigate the microstructures and morphologies of MnO2 nanomaterials. Different microstructures and morphologies of MnO2 nanomaterials were found when MnO2 was prepared by different methods. As shown in Figure 1a, an irregular sheet structure was obtained when MnO2 was synthesized by reducing potassium permanganate using ethanol as a reducing agent under normal temperature and pressure at pH 7 (Method 1). MnO2 nanowire (Figure 1b) was observed when MnO2 was prepared via a hydrothermal synthesis under acidic conditions (Method 2). The length of the nanowire is approximately 6–10 μm. The obvious different morphologies were expected to result in different electrochemical performances. The crystal structure of the MnO2 nanowire (MnO2-NW) was characterized by X-ray powder diffraction (XRD). In XRD patterns of MnO2-NW (Figure 1c), the diffraction peaks at 12.78°,18.06°, 28.82°, 37.82° and 42.17° are indexed to crystal Surface of (110), (200), (310), (211) and (301), respectively, which are matched with the standard diffraction pattern of α-MnO2 (JCPDS card No.44-0141). More importantly, no other impurities are found in these XRD patterns, revealing that as-synthesized MnO2-NW is well crystallized and at high purity.
X-ray photoelectron spectroscopy (XPS) was used to study the surface composition of MnO2-NW (Figure 2). According to the full scan XPS spectrum (Figure 2a), Mn and O were observed on the surface of the sample of MnO2-NW. In the high-resolution spectrum of O 1s (Figure 2b), the peak at 529.88 eV corresponded to lattice oxygen, and the appearance of a high-intensity peak indicates that a large amount of lattice oxygen (O latt) was present in MnO2 nanowires. The characteristic peak of O abs at 531.49 eV was related to surface-adsorbed oxygen, hydroxyl groups, or defect-related oxygen species [36,37]. Figure 2c displays the XPS energy spectrum of Mn 3s. The valence state of Mn was usually determined based on the peak spacing of Mn 3s. The peak spacing of Mn 3s in MnO2-NW was 4.70 eV, indicating that most of the Mn in the sample exists in the form of +4 valence. In the XPS spectra of Mn 2p (Figure 2d), the peaks at 642.47 eV and 641.04 eV correspond to Mn4+ and Mn3+, respectively. The content of Mn3+ is relatively low, indicating that most of the Mn in MnO2-NW samples exist in the form of +4 valence [38,39].

3.2. Electrochemical Behavior of TBBPA on MnO2-NW/CP

Figure 3a shows the cyclic voltammograms of 500 nM TBBPA on CP, MnO2/CP (MnO2 nanosheet), and MnO2-NW/CP, respectively. The buffer solution was 0.10 M phosphate buffer solution (pH 6.5). As can be seen from Figure 3a, an irreversible oxidation peak at 0.60 V (vs. Ag/AgCl) was found to correspond to the irreversible oxidation of TBBPA at the working electrode surface. The comparison of the oxidation peak current of TBBPA on the above three electrodes is displayed in Figure 3b. It was shown that the oxidation peak current of TBBPA was slightly enhanced by modifying the nanosheet MnO2 on CP. In addition, the electrochemical signal of TBBPA on MnO2-NW/CP was approximately twice higher than that on the bare CP electrode. Furthermore, the oxidation potential shifted to 0.58 V. It is well understood that the electrochemical activity of nano MnO2 is shape-dependent. Compared with the MnO2 nanosheet, the MnO2 nanowire can provide a higher surface area and more abundant surface oxygen, which can result in enhanced catalytic activity. The oxidation of TBBPA was obviously promoted when the CP electrode was modified by MnO2-NW, which may be due to the larger specific surface area, better electrical conductivity and the oxidizing ability of MnO2-NW. Hence, MnO2-NW/CP was chosen for the detection of TBBPA.
CVs of MnO2-NW/CP in 0.1 M KCl solution containing 2.5 mM K3[Fe(CN)6] at various scan rates were recorded in the range of 30–150 mV/s to investigate the effective electrochemical area of the MnO2-NW/CP electrode. It was observed that the oxidation and reduction peaks of potassium ferricyanide increase with increasing scanning speed at the MnO2-NW/CP electrode, and the potential difference between the oxidation and reduction peaks becomes larger. Diffusion-controlled behavior of K3[Fe(CN)6] could be deduced from the linear relationship between oxidation peak currents ip and the square root of scan rates υ1/2 with the regression equation of ip = 0.04109 υ1/2 − 0.05549 (R2 = 0.9976) (Figure 3d). The Randles–Sevcik equation was used to calculate the effective surface area (ESA) [40].
i p = 0.4463 n F A C n F v D R T 1 / 2 = 2.69 × 10 5 n 2 / 3 A C D 1 / 2 v 1 / 2           ( 25   ° C )
where ip (A) is the oxidation or reduction peak current; n is the total number of electrons exchanged in the redox reaction; A (cm2) is the effective surface area (ESA); C (mol/cm3) is the concentration of K3[Fe(CN)6] (2.5 mM); D (cm2/s) is the diffusion coefficient of [Fe(CN)6]3− (6.5 × 10−6 cm2/s); and υ (V/s) is the scan rate. In our experiment, the slope was 0.04109, and the ESA value was calculated as 0.24 cm2. The ESA value of the bare carbon paper electrode was determined using a similar process to be 0.026 cm2. Thus, the ESA of MnO2-NW/CP was about nine times greater than that of a bare carbon paper electrode, suggesting the excellent electrochemical properties of MnO2-NW/CP.
Figure 4a shows CVs of MnO2-NW/CP in 0.1 M pH 6.5 phosphate buffer solutions containing 500 nM TBBPA at different scan rates in the range of 25–150 mV/s. The peak current in CVs is coupled with the change of concentration gradient with time. Thus, the peak currents are obviously influenced by the scan rate. This is because the scan rate determines the time between the switching potential and the peak potential. It was shown that the oxidation peak currents of TBBPA kept increasing with the increase in the scan rate in the range of 25–150 mV/s. Figure 4b shows a plot of the TBBPA oxidation peak current versus sweep rate with the linear equation ip = 0.0975υ + 0.00459 (R2 = 0.9915). Figure 4c is a plot of ip versus υ1/2 with the linear equation ip = 1.67755υ1/2−6.61462 (R2 = 0.9531). The oxidation peak currents of TBBPA were background subtracted to correct for the larger charging current contribution at the higher scan rate. The oxidation peak current of TBBPA is proportional to the potential scan rate, which is in accordance with the formula of ip = (n2F2ΓAV)/(4RT). Moreover, the independence between ip and the square root of the scan rate reveals that the oxidation of TBBPA on MnO2-NW/CP was not controlled by the diffusing process. Thus, TBBPA oxidation on a MnO2-NW/CP is an adsorption-controlled process. The oxidation peak potentials of TBBPA were found to shift toward positive potentials with the increase in scan rates, which is related to the irreversible oxidation process of TBBPA.

3.3. Optimization of Experimental Conditions

DPV was chosen to prepare a sensitive sensor for the determination of TBBPA. The experiment parameters, including the pH value of the buffer solution, the enriching time and the amount of MnO2-NW, were optimized by recording the electrochemical performance of TBBPA under different experiment conditions.
Figure 5a shows the oxidation peak currents and peak potentials of TBBPA at different pH conditions. It was found that the oxidation peak currents of TBBPA increased with increasing pH in the range of 5.5–6.5 and reached a maximum at pH 6.5. The pKa1 and pKa2 of TBBPA were reported to be 7.5 and 8.5. TBBPA was not dissociated when the pH was 6.5. The undissociated TBBPA can be adsorbed more easily on a MnO2-NW/CP electrode than the dissociated form. This resulted in the maximum oxidation peak current of TBBPA at pH 6.5. Hence, pH 6.5 was chosen as the optimum pH condition in the subsequent experiments. In the pH range of 5.8–8.0, the oxidation peak potential of TBBPA negatively linearly shifted with an increase in pH value following the equations of Ep = 0.978–0.0594 pH. This shows that protons took part in the electrochemical reaction of TBBPA. The slope −0.0594 V/pH was close to the value obtained from the Nernst equation (−0.059 V/pH), which indicated that the number of electrons was equal to the number of protons during the oxidation of TBBPA [41]. Referring to the oxidation mechanism of TBBPA on Fe3O4@SiO2@CDs-CTAB/GCE [20], the oxidation mechanism of TBBPA on MnO2-NW/CP is deduced in Figure 6.
The effect of enrichment time on the oxidation peak current at the open circuit of TBBPA (500 nM) is shown in Figure 5b. When the enrichment time was less than 5 min, the peak current increased rapidly with the increase in enrichment time. This is because more TBBPA was adsorbed on the MnO2-NW/CP electrode with increasing enrichment time. When the enrichment time was above 5 min, the oxidation peak current tended to be stable due to the adsorption saturation of TBBPA. Therefore, 5 min of enrichment time was chosen.
The effect of the modification amount of MnO2-NW was investigated in Figure 5c. It was found that the oxidation current of TBBPA increased sharply when the modification amount of MnO2-NW suspension increased from 3.0 to 7.0 µL. This might be a result of more catalyst-enhanced catalytic ability. However, when the modification amount was over 7.0 µL, the oxidation peak current decreased due to the deterioration of electrical conductivity. Therefore, 7.0 µL suspension of MnO2-NW was selected as the optimal modification amount in this work.

3.4. Electrochemical Detection of TBBPA

DPV of MnO2-NW/CP in different TBBPA concentrations was investigated under optimal experimental conditions (Figure 7). As shown in Figure 7a, the oxidation currents increased when TBBPA was added successively from 70 to 500 nM. The linear regression equation was ip = 0.00323c + 0.06304, R2 = 0.9914 (Figure 7b). The estimated detection limit is 3.1 nM based on a three-signal-to-noise ratio of 3. It could be seen that our electrochemical sensor exhibited a wide linear range and a low detection limit, which might be due to the high ESA of the MnO2-NW/CP, with plenty of active sites. Our experimental results were compared with the literature reports in Table 1, indicating that this method has some favorable advantages.
Five parallel MnO2-NW/CP electrodes were prepared for the detection of 500 nM TBBPA, and the detection relative standard deviation was 3.8%. One of the above five electrodes was further investigated five consecutive times, and the relative standard deviation was 1.9%. The results showed good reproducibility and precision of MnO2-NW/CP.

3.5. Real Samples Analysis

To estimate the potential of a MnO2-NW for the detection of TBBPA in practical applications, the fabricated electrochemical sensor was employed to detect TBBPA in lake water (Suzhou, China). Before detection, the water was filtered through a 0.45 μm membrane to remove suspended solids. No response to TBBPA was found in these real samples, indicating that the concentration of TBBPA was extremely low or that no TBBPA existed in these real samples. Then standard solutions of TBBPA were spiked into real sample solutions for recovery evaluation. The recoveries were in the range of 94.5% to 100.5%. The results indicated that MnO2-NW/CP is reliable and could be applied to detect TBBPA in the actual water sample.

4. Conclusions

In this study, a cost-effective, simple fabricated and sensitive electrochemical sensor was prepared by modifying carbon paper with a MnO2 nanomaterial. The performance of electrochemical sensors was proven to be closely related to the microscopic morphology of MnO2. The fabricated MnO2-NW/CP displayed attractive performance for the detection of TBBPA. The sensitivity and stability of MnO2-NW/CP were satisfied, and the selectivity of MnO2-NW/CP can be further enhanced by combining other materials, such as molecular imprinting technology. It is reasonable to believe that MnO2 nanowire-based electrochemical sensors might be a promising method for the effective detection of TBPPA in environment monitoring.

Author Contributions

Conceptualization, Q.Z.; methodology, Q.Z. and C.Z.; software, C.Z. and Q.W.; validation, Q.W.; investigation, C.Z.; resources, Q.Z.; data curation, C.Z. and F.Y.; writing—original draft preparation, C.Z. and F.Y.; writing—review and editing Q.Z., J.L. and D.W.; funding acquisition, Q.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (No. 20905055), Postgraduate Research and Practice Innovation Program of Jiangsu Province (No. SJCX22_1564).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lee, S.; Song, G.J.; Kannan, K.; Moon, H.B. Occurrence of PBDEs and other alternative brominated flame retardants in sludge from wastewater treatment plants in Korea. Sci. Total Environ. 2014, 470, 1422–1429. [Google Scholar] [PubMed]
  2. Shi, Z.X.; Zhang, L.; Li, J.G.; Wu, Y.N. Legacy and emerging brominated flame retardants in China: A review on food and human milk contamination, human dietary exposure and risk assessment. Chemosphere 2018, 198, 522–536. [Google Scholar] [CrossRef] [PubMed]
  3. Shi, Z.X.; Zhang, L.; Zhao, Y.F.; Sun, Z.W.; Zhou, X.Q.; Li, J.G.; Wu, Y.N. Dietary exposure assessment of Chinese population to tetrabromobisphenol-A, hexabromocyclododecane and decabrominated diphenyl ether: Results of the 5th Chinese Total Diet Study. Environ. Pollut. 2017, 229, 539–547. [Google Scholar] [CrossRef] [PubMed]
  4. Wluka, A.; Wozniak, A.; Wozniak, E.; Michalowicz, J. Tetrabromobisphenol A, terabromobisphenol S and other bromophenolic flame retardants cause cytotoxic effects and induce oxidative stress in human peripheral blood mononuclear cells (in vitro study). Chemosphere 2020, 261, 127705. [Google Scholar] [CrossRef]
  5. Zuiderveen, E.A.R.; Slootweg, J.C.; de Boer, J. Novel brominated flame retardants—A review of their occurrence in indoor air, dust, consumer goods and food. Chemosphere 2020, 255, 126816. [Google Scholar]
  6. Nakajima, A.; Saigusa, D.; Tetsu, N.; Yamakuni, T.; Tomioka, Y.; Hishinuma, T. Neurobehavioral effects of tetrabromobisphenol A, a brominated flame retardant, in mice. Toxicol. Lett. 2009, 189, 78–83. [Google Scholar] [CrossRef]
  7. Feiteiro, J.; Mariana, M.; Cairrao, E. Health toxicity effects of brominated flame retardants: From environmental to human exposure. Environ. Pollut. 2021, 285, 117475. [Google Scholar]
  8. Jagic, K.; Dvorscak, M.; Klincic, D. Analysis of brominated flame retardants in the aquatic environment: A review. Arh. Hig. Rada Toksikol. 2021, 72, 254–267. [Google Scholar] [PubMed]
  9. Gao, W.; Li, G.L.; Liu, H.; Tian, Y.; Li, W.T.; Fa, Y.; Cai, Y.Q.; Zhao, Z.S.; Yu, Y.L.; Qu, G.B.; et al. Covalent organic frameworks with tunable pore sizes enhanced solid-phase microextraction direct ionization mass spectrometry for ultrasensitive and rapid analysis of tetrabromobisphenol A derivatives. Sci. Total Environ. 2021, 764, 144388. [Google Scholar] [CrossRef]
  10. Zhang, S.; Liu, J.; Hou, X.; Zhang, H.; Zhu, Z.; Jiang, G. Sensitive method for simultaneous determination of TBBPA and its ten derivatives. Talanta 2023, 264, 124750. [Google Scholar]
  11. Zeng, L.S.; Cui, H.R.; Chao, J.L.; Huang, K.; Wang, X.; Zhou, Y.K.; Jing, T. Colorimetric determination of tetrabromobisphenol A based on enzyme-mimicking activity and molecular recognition of metal-organic framework-based molecularly imprinted polymers. Microchem. J. 2020, 187, 142. [Google Scholar] [CrossRef]
  12. Feng, J.; Tao, Y.; Shen, X.; Jin, H.; Zhou, T.; Zhou, Y.; Lee, Y.I. Highly sensitive and selective fluorescent sensor for tetrabromobisphenol-A in electronic waste samples using molecularly imprinted polymer coated quantum dots. Microchem. J. 2019, 144, 93–101. [Google Scholar] [CrossRef]
  13. Fu, H.J.; Wang, Y.; Xiao, Z.L.; Wang, H.; Li, Z.F.; Shen, Y.D.; Lei, H.T.; Sun, Y.M.; Xu, Z.L.; Hammock, B. A rapid and simple fluorescence enzyme-linked immunosorbent assay for tetrabromobisphenol A in soil samples based on a bifunctional fusion protein. Ecotoxicol. Environ. Saf. 2020, 188, 109904. [Google Scholar] [CrossRef] [PubMed]
  14. Li, Z.F.; Wang, Y.; Vasylieva, N.; Wan, D.B.; Yin, Z.H.; Dong, J.X.; Hammock, B. An Ultrasensitive Bioluminescent Enzyme Immunoassay Based on Nanobody/Nanoluciferase Heptamer Fusion for the Detection of Tetrabromobisphenol A in Sediment. Anal. Chem. 2020, 92, 10083–10090. [Google Scholar] [CrossRef]
  15. Zenasni, M.; Quintero-Jaime, A.; Salinas-Torres, D.; Benyoucef, A.; Morallón, E. Electrochemical synthesis of composite materials based on titanium carbide and titanium dioxide with poly (N-phenyl-o-phenylenediamine) for selective detection of uric acid. J. Electroanal. Chem. 2021, 895, 115481. [Google Scholar] [CrossRef]
  16. Zhang, Z.H.; Cai, R.; Long, F.; Wang, J. Development and application of tetrabromobisphenol A imprinted electrochemical sensor based on graphene/carbon nanotubes three-dimensional nanocomposites modified carbon electrode. Talanta 2015, 134, 435–442. [Google Scholar] [CrossRef]
  17. Zhou, T.T.; Feng, Y.Q.; Zhou, L.X.; Tao, Y.; Luo, D.; Jing, T.; Shen, X.L.; Zhou, Y.K.; Mei, S.R. Selective and sensitive detection of tetrabromobisphenol-A in water samples by molecularly imprinted electrochemical sensor. Sens. Actuators B Chem. 2016, 236, 153–162. [Google Scholar] [CrossRef]
  18. Wang, Y.Y.; Liu, G.S.; Hou, X.D.; Huang, Y.N.; Li, C.Y.; Wu, K.B. Assembling gold nanorods on a poly-cysteine modified glassy carbon electrode strongly enhance the electrochemical reponse to tetrabromobisphenol A. Microchim. Acta 2015, 183, 689–696. [Google Scholar] [CrossRef]
  19. Zhou, T.T.; Zhao, X.Y.; Xu, Y.H.; Tao, Y.; Luo, D.; Hu, L.Q.; Jing, T.; Zhou, Y.K.; Wang, P.; Mei, S.R. Electrochemical determination of tetrabromobisphenol A in water samples based on a carbon nanotubes@zeolitic imidazole framework-67 modified electrode. RSC Adv. 2020, 10, 2123–2132. [Google Scholar] [CrossRef]
  20. Guo, J.H.; Zhou, B.Y.; Li, S.Y.; Tong, Y.Y.; Li, Z.; Liu, M.H.; Li, Y.H.; Qu, T.X.; Zhou, Q.X. Novel electrochemical sensor from magnetic carbon dots and cetyltrimethylammonium bromide for sensitive measurement of tetrabromobisphenol A in beverages. Chemosphere 2022, 298, 134326. [Google Scholar] [CrossRef]
  21. Shao, Y.M.; Zhu, Y.; Zheng, R.; Wang, P.; Zhao, Z.Z.; An, J. Highly sensitive and selective surface molecularly imprinted polymer electrochemical sensor prepared by Au and MXene modified glassy carbon electrode for efficient detection of tetrabromobisphenol A in water. Adv. Compos. Hybrid Mater. 2022, 5, 3104–3116. [Google Scholar] [CrossRef]
  22. Zhao, Q.; Zhou, H.; Wu, W.; Wei, X.; Jiang, S.; Zhou, T.; Lu, Q. Sensitive electrochemical detection of tetrabromobisphenol A based on poly (diallyldimethylammonium chloride) modified graphitic carbon nitride-ionic liquid doped carbon paste electrode. Electrochim. Acta 2017, 254, 214–222. [Google Scholar] [CrossRef]
  23. Zhou, Q.; Zhang, Y.Y.; Zeng, T.; Wan, Q.J.; Yang, N.J. Morphology-dependent sensing performance of CuO nanomaterials. Anal. Chim. Acta 2021, 1171, 338663. [Google Scholar] [CrossRef] [PubMed]
  24. Luo, S.X.; Yang, M.Z.; Wu, Y.H.; Li, J.; Qin, J.; Feng, F. A Low Cost Fe3O4-Activated Biochar Electrode Sensor by Resource Utilization of Excess Sludge for Detecting Tetrabromobisphenol A. Micromachines 2022, 13, 115. [Google Scholar] [CrossRef]
  25. Chen, X.Y.; Zhang, Y.Y.; Li, C.; Li, C.; Zeng, T.; Wan, Q.J.; Li, Y.W.; Ke, Q.; Yang, N.J. Nanointerfaces of expanded graphite and Fe2O3 nanomaterials for electrochemical monitoring of multiple organic pollutants. Electrochim. Acta 2020, 329, 135118. [Google Scholar] [CrossRef]
  26. Baral, A.; Satish, L.; Zhang, G.Y.; Ju, S.H.; Ghosh, M.K. A Review of Recent Progress on Nano MnO2: Synthesis, Surface Modification and Applications. J. Inorg. Organomet. Polym. Mater. 2020, 31, 899–922. [Google Scholar] [CrossRef]
  27. Yu, Y.; Liu, S.L.; Ji, J.; Huang, H.B. Amorphous MnO2 surviving calcination: An efficient catalyst for ozone decomposition. Catal. Sci. Technol. 2019, 9, 5090–5099. [Google Scholar] [CrossRef]
  28. Karami-Kolmoti, P.; Beitollahi, H.; Modiri, S. Electrochemical Sensor for Simple and Sensitive Determination of Hydroquinone in Water Samples Using Modified Glassy Carbon Electrode. Biomedicines 2023, 11, 1869. [Google Scholar] [CrossRef]
  29. Huang, Z.N.; Liu, G.C.; Zou, J.; Jiang, X.Y.; Liu, Y.P.; Yu, J.G. A hybrid composite of recycled popcorn-shaped MnO2 microsphere and Ox-MWCNTs as a sensitive non-enzymatic amperometric H2O2 sensor. Microchem. J. 2020, 158, 105215. [Google Scholar] [CrossRef]
  30. Yakubu, S.; Xiao, J.X.; Gu, J.P.; Cheng, J.; Wang, J.; Li, X.S.; Zhang, Z. A competitive electrochemical immunosensor based on bimetallic nanoparticle decorated nanoflower-like MnO2 for enhanced peroxidase-like activity and sensitive detection of Tetrabromobisphenol A. Sens. Actuators B Chem. 2020, 325, 128909. [Google Scholar] [CrossRef]
  31. Torrinha, A.; Morais, S. Electrochemical (bio)sensors based on carbon cloth and carbon paper: An overview. Trends Anal. Chem. 2021, 142, 116324. [Google Scholar] [CrossRef]
  32. Li, Y.H.; Xu, Q.Z.; Li, Q.Y.; Wang, H.Q.; Huang, Y.G.; Xu, C.W. Pd deposited on MWCNTs modified carbon fiber paper as high-efficient electrocatalyst for ethanol electrooxidation. Electrochim. Acta 2014, 147, 151–156. [Google Scholar] [CrossRef]
  33. Kannan, P.; Maiyalagan, T.; Marsili, E.; Ghosh, S.; Niedziolka-Jonsson, J.; Jonsson-Niedziolka, M. Hierarchical 3-dimensional nickel-iron nanosheet arrays on carbon fiber paper as a novel electrode for non-enzymatic glucose sensing. Nanoscale 2016, 8, 843–855. [Google Scholar] [CrossRef]
  34. Sun, H.Y.; Wang, C.X.; Xu, Y.J.; Dai, D.M.; Deng, X.Y.; Gao, H.T. A Novel Electrochemical Sensor Based on A Glassy Carbon Electrode Modified with GO/MnO2 for Simultaneous Determination of Trace Cu(II) and Pb(II) in Environmental Water. ChemistrySelect 2019, 4, 11862–11871. [Google Scholar] [CrossRef]
  35. Radulescu, M.C.; Bucur, M.P.; Bucur, B.; Radu, G.L. Ester flavorants detection in foods with a bienzymatic biosensor based on a stable Prussian blue-copper electrodeposited on carbon paper electrode. Talanta 2019, 199, 541–546. [Google Scholar] [CrossRef] [PubMed]
  36. Tagsin, P.; Suksangrat, P.; Klangtakai, P.; Srepusharawoot, P.; Ruttanapun, C.; Kumnorkaew, P.; Pimanpang, S.; Amornkitbamrung, V. Electrochemical mechanisms of activated carbon, alpha-MnO2 and composited activated carbon-alpha-MnO2 films in supercapacitor applications. Appl. Surf. Sci. 2021, 570, 151056. [Google Scholar] [CrossRef]
  37. Cui, Y.; Song, H.K.; Shi, Y.Y.; Ge, P.X.; Chen, M.D.; Xu, L.L. Enhancing the Low-Temperature CO Oxidation over CuO-Based alpha-MnO2 Nanowire Catalysts. Nanomaterials 2022, 12, 2083. [Google Scholar] [CrossRef]
  38. Zhou, X.; Ye, X.X.; Wu, K.B.; Li, C.; Wang, Y.Y. Electrochemical sensing of terabromobisphenol A at a polymerized ionic liquid film electrode and the enhanced effects of anions. Ionics 2018, 24, 2843–2850. [Google Scholar] [CrossRef]
  39. Ganesh, A.; Sivakumar, T.; Sankar, G. Biomass-derived porous carbon-incorporated MnO2 composites thin films for asymmetric supercapacitor: Synthesis and electrochemical performance. J. Mater. Sci. Mater. Electron. 2022, 33, 14772–14783. [Google Scholar] [CrossRef]
  40. Shen, T.Y.; Liu, T.C.; Mo, H.Q.; Yuan, Z.C.; Cui, F.; Jin, Y.X.; Chen, X.J. Cu-based metal-organic framework HKUST-1 as effective catalyst for highly sensitive determination of ascorbic acid. RSC Adv. 2020, 10, 22881–22890. [Google Scholar] [CrossRef]
  41. Stradins, J.; Hasanli, B. Anodic voltammetry of phenol and benzenethiol derivatives.1. Influence of pH on electrooxidation potentials of substituted phenols and evaluation of pK(a) from anodic voltammetry data. J. Electroanal. Chem. 1993, 353, 57–69. [Google Scholar] [CrossRef]
  42. Shen, J.; Bian, C.; Xia, S.H.; Wu, K.B. Poly(sulfosalicylic acid)-functionalized gold nanoparticles for the detection of tetrabromobisphenol A at pM concentrations. J. Hazard. Mater. 2020, 388, 121733. [Google Scholar] [CrossRef]
  43. Li, Z.; Hu, J.Y.; Lou, Z.Z.; Zeng, L.X.; Zhu, M.S. Molecularly imprinted photoelectrochemical sensor for detecting tetrabromobisphenol A in indoor dust and water. Microchem. J. 2021, 188, 320. [Google Scholar] [CrossRef] [PubMed]
  44. Yakubu, S.; Jia, B.Y.; Guo, Y.J.; Zou, Y.M.; Song, N.H.; Xiao, J.X.; Liang, K.L.; Bu, Y.Q.; Zhang, Z. Indirect competitive-structured electrochemical immunosensor for tetrabromobisphenol A sensing using CTAB-MnO2 nanosheet hybrid as a label for signal amplification. Anal. Bioanal. Chem. 2021, 413, 4217–4226. [Google Scholar] [CrossRef] [PubMed]
Figure 1. SEM of granular MnO2 (a) and MnO2-NW (b); XRD of MnO2-NW (c).
Figure 1. SEM of granular MnO2 (a) and MnO2-NW (b); XRD of MnO2-NW (c).
Chemosensors 11 00482 g001
Figure 2. XPS figures of MnO2-NW. (a) full spectrum, (b) O 1s, (c) Mn 3s, (d) Mn 2p.
Figure 2. XPS figures of MnO2-NW. (a) full spectrum, (b) O 1s, (c) Mn 3s, (d) Mn 2p.
Chemosensors 11 00482 g002
Figure 3. (a) CVs of 500 nM TBBPA on different electrodes; (b) oxidation current of TBBPA on different electrodes; (c) CVs of MnO2−NW/CP in 2.5 mM K3[Fe(CN)6] at different scan rates; (d) The linear plot of K3[Fe(CN)6] ipa vs. υ1/2.
Figure 3. (a) CVs of 500 nM TBBPA on different electrodes; (b) oxidation current of TBBPA on different electrodes; (c) CVs of MnO2−NW/CP in 2.5 mM K3[Fe(CN)6] at different scan rates; (d) The linear plot of K3[Fe(CN)6] ipa vs. υ1/2.
Chemosensors 11 00482 g003
Figure 4. (a) CVs of 500 nM TBBPA in 0.1 M pH 6.5 PBS on MnO2−NW/CP at different scan rates; (b)The linear plot of ip vs. υ; (c)The linear plot of ip vs. υ1/2.
Figure 4. (a) CVs of 500 nM TBBPA in 0.1 M pH 6.5 PBS on MnO2−NW/CP at different scan rates; (b)The linear plot of ip vs. υ; (c)The linear plot of ip vs. υ1/2.
Chemosensors 11 00482 g004
Figure 5. (a) The effect of pH on the oxidation peak current and potential of TBBPA; (b) The effect of enrichment time on the peak current of TBBPA; (c) The effect of MnO2-NW amount on the peak current of TBBPA.
Figure 5. (a) The effect of pH on the oxidation peak current and potential of TBBPA; (b) The effect of enrichment time on the peak current of TBBPA; (c) The effect of MnO2-NW amount on the peak current of TBBPA.
Chemosensors 11 00482 g005
Figure 6. Electrochemical oxidation mechanism of TBBPA.
Figure 6. Electrochemical oxidation mechanism of TBBPA.
Chemosensors 11 00482 g006
Figure 7. (a) DPVs of MnO2−NW/CP in 0.1 M pH 6.5 PBS containing different concentrations of TBBPA; (b) The corresponding calibration plots between the peak currents and the concentrations of TBBPA.
Figure 7. (a) DPVs of MnO2−NW/CP in 0.1 M pH 6.5 PBS containing different concentrations of TBBPA; (b) The corresponding calibration plots between the peak currents and the concentrations of TBBPA.
Chemosensors 11 00482 g007
Table 1. Comparison of the experimental results with other electrodes or methods.
Table 1. Comparison of the experimental results with other electrodes or methods.
Sensor AssemblyMethodLODLinear RangeReference
CNTs@ZIF-67/PFDA/ABDPV4.2 nM10~150 nM[19]
AuNPs-PSSA/GCEDPV25 pM0.1~10 nM[42]
MI-TiO2/Au/rGODPV, CV0.51 nM1.68~100 nM[43]
CTAB-MnO2/PdI-t0.17 ng/mL0~81 ng/mL[44]
poly(PPBim-DS)/GCEDPV20 nM0.05~10 µM[39]
MnO2-NW/CPDPV3.1 nM70~500 nMThis work
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhu, C.; Wu, Q.; Yuan, F.; Liu, J.; Wang, D.; Zhang, Q. Novel Electrochemical Sensor Based on MnO2 Nanowire Modified Carbon Paper Electrode for Sensitive Determination of Tetrabromobisphenol A. Chemosensors 2023, 11, 482. https://0-doi-org.brum.beds.ac.uk/10.3390/chemosensors11090482

AMA Style

Zhu C, Wu Q, Yuan F, Liu J, Wang D, Zhang Q. Novel Electrochemical Sensor Based on MnO2 Nanowire Modified Carbon Paper Electrode for Sensitive Determination of Tetrabromobisphenol A. Chemosensors. 2023; 11(9):482. https://0-doi-org.brum.beds.ac.uk/10.3390/chemosensors11090482

Chicago/Turabian Style

Zhu, Chunmao, Qi Wu, Fanshu Yuan, Jie Liu, Dongtian Wang, and Qianli Zhang. 2023. "Novel Electrochemical Sensor Based on MnO2 Nanowire Modified Carbon Paper Electrode for Sensitive Determination of Tetrabromobisphenol A" Chemosensors 11, no. 9: 482. https://0-doi-org.brum.beds.ac.uk/10.3390/chemosensors11090482

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