Next Article in Journal
Stochastic Analysis of the Gas Flow at the Gas Diffusion Layer/Channel Interface of a High-Temperature Polymer Electrolyte Fuel Cell
Previous Article in Journal
IMU-Assisted 2D SLAM Method for Low-Texture and Dynamic Environments
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Determination of Bisphenol A in Beverages by an Electrochemical Sensor Based on Rh2O3/Reduced Graphene Oxide Composites

1
Key Laboratory for Environmental Factors Control of Agro-Product Quality Safety, Agro-Environmental Protection Institute, Ministry of Agriculture, Tianjin 300191, China
2
CAS Key Laboratory of Biobased Materials, Qingdao Institute of Bioenergy and Bioprocess Technology, Chinese Academy of Sciences, Qingdao 266101, China
*
Author to whom correspondence should be addressed.
Appl. Sci. 2018, 8(12), 2535; https://doi.org/10.3390/app8122535
Submission received: 8 November 2018 / Revised: 23 November 2018 / Accepted: 27 November 2018 / Published: 7 December 2018
(This article belongs to the Section Nanotechnology and Applied Nanosciences)

Abstract

:
A novel electrochemical sensor, based on a Rh2O3–reduced graphene oxide (rGO) composite modified carbon electrode, has been developed for detecting bisphenol A (BPA) in beverages. The prepared Rh2O3/rGO and its precursor materials were characterized by scanning electron microscope (SEM) and X-ray diffraction (XRD). Under optimum conditions, the sensor presented good electrochemical performance for analyzing BPA, with a linear range of 0.6–40 μM, detection limit of 0.12 μM, good reproducibility, and excellent stability. The good performance can be attributed to the combination of the good catalytic properties of Rh2O3 and good conductivity of rGO. The sensor is directly used for detecting BPA in the residual solutions of four beverages after simple filtration, with satisfactory recoveries of 93–99%.

1. Introduction

Bisphenol A (BPA), one of the most produced and used chemicals, has been widely used for preparing industrial polymers, such as polycarbonate, epoxy resin, and thermosensitive paper [1]. Furthermore, it is often involved in various food contact materials (i.e., canned packaging, bottles, and lacquer coating). Its widespread usage, especially in these food contact materials, has inevitably lead to its exposure to human beings via food and drinking water [2,3]. As an endocrine-disrupting chemical (EDC) [4,5], it can destroy natural or synthetic compounds with endocrine function by simulating or blocking endogenous hormones [6]. To date, a temporary tolerable daily intake (t-TDI) of 4 μg kg−1 b.w. day−1 for BPA has been proposed by the European Food Safety Authority (EFSA) [7]. Legislation against BPA usage in some infant food package and bottles has also been enacted, in many countries all over the world [8].
For monitoring the pollution levels and ecological risks of BPA, many methods based on capillary electrophoresis [9,10,11], gas chromatography coupled with mass spectrometry (GC–MS) [12,13,14,15], high performance liquid chromatography mass spectrometry (LC–MS) [16,17], enzyme linked immune sorbent assay (ELISA) [18,19,20], and electrochemical methods [21,22,23,24] have been developed. Pretreatment procedures, such as solid-phase extraction [24,25,26], liquid-liquid extraction [27,28], and immunoaffinity chromatography [29,30], are often involved when using GC–MS, HPLC–MS, and ELISA for analyzing BPA in complex matrices [31,32]. Although these common methods can’t do without complementary pretreatments and expensive equipment, they are still the reliable strategies for analyzing environmental BPA.
Compared with the methods mentioned above, electrochemical methods, with the advantages of easy preparation, high sensitivity, simplicity for operators, and in-situ monitoring [33], are often proposed as one of the preferable techniques for routine analysis. In order to analyze trace levels of chemicals, improving the response signals of substances is a vital part of these electrochemical methods. Modification of the electrodes with various functional materials is one of the most recommended strategies. For example, electrodes modified with graphene nanomaterials present excellent electric conductivity as a result of their good conductivity [33,34,35]. Molybdenum–selenide (MoSe2) electrodes showed much more sensitivity to BPA, as a result of increasing surface active sites [36]. Noble metal oxides, such as RuO2, PdO, IrO2, and Rh2O3, are widely used in heterogeneous catalysis systems. Rh2O3 has been proven to be an excellent catalyst for the oxidation of CO [37,38] and the deposition of H2O2 [39] and N2O [40], but it is still rarely reported for sensors.
In this paper, a Rh2O3–GO composite nanomaterial was chosen, and a novel electrochemical sensor was fabricated with a modified glassy carbon electrode (GCE) for excellent catalytic performance and conductivity. The modified glassy carbon electrode (GCE), coated with a Rh2O3–rGO membrane with excellent stability and selectivity, was synthesized and was then used for analyzing BPA dissolved in drinks for the first time after simple filtration.

2. Materials and Methods

2.1. Materials and Reagents

Poly (allylamine hydrochloride) (Mw = 15,000) was purchased from Shanghai Macklin Biochemical Co., Ltd. (Shanghai, China). Rhodium (III) chloride trihydrate (RhCl3·3H2O), formaldehyde solution (HCHO, 37%), sodium hypochlorite (NaClO), bisphenol A (BPA), phosphotungstic acid (H3[PW12O4]), graphite powders, potassium hydroxide (KOH), sodium hydroxide (NaOH), potassium permanganate (KMnO4), sulfuric acid (H2SO4, 98%), sodium nitrate (NaNO3), and phosphoric acid (H3PO4) were purchased from Sinopharm Group Chemical Reagent Co. Ltd. (Shanghai, China). All the chemicals were analytical grade. BPA was dissolved in ethanol at 1 mM and kept at −4 °C. Phosphate buffer solution (PBS, 0.1 M, pH = 7.0) was used reduced, and a supporting electrolyte (containing 0.1 mM phosphotungstic acid) was used.

2.2. Instruments

All electrochemical experiments were carried out on a CHI660e electrochemical workstation (Chenhua Instruments, Shanghai, China) with a conventional three-electrode cell. A saturated Ag/AgCl electrode was used as reference electrode, platinum wire as a counter electrode, and a bare glassy carbon electrode (GCE) (or modified GCE) as the working electrode.
The morphology of the prepared nanocomposites was characterized by scanning electron microscopy (SEM, Hitachi S-4800, Tokyo, Japan) and X-ray diffraction (XRD, DX-2700 Bruker D8 Advance).

2.3. Synthesis of Rh2O3 Nanoparticles (Rh2O3–NPs)

Rh2O3 nanoparticles were prepared, according to the method in the literature [41] with slight modification. Typically, RhCl3 (63 mg) and poly (allylamine hydrochloride) (200 mg) were dissolved in 100 mL of water. The pH of the solution was adjusted to 7.0 by KOH. After that, 10 mL of a HCHO solution was added to the solution. Then, the reaction solution was transferred to a teflon-lined high-pressure vessel and heated at 120 °C for 6 h. Then, 15 mL of NaClO solution was added to the mixture and mechanically stirred for 72 h at room temperature. Finally, the suspension was centrifuged and then washed with water several times, to eliminate the residues and obtain Rh2O3–NPs.
Graphene oxide (GO) was prepared, according to previous reports [42]. Briefly, graphite (0.5 g) and NaNO3 (0.5 g) were added into concentrated H2SO4 (23 mL) in an ice bath, followed by slow addition of 3 g KMnO4. The mixture was heated at 35 °C for two hours, and then 40 mL of H2O was added. After that, the reaction temperature was raised to 95 °C, kept for 30 min, and then followed by the addition of 100 mL H2O and 20 mL of H2O2. The final suspension was filtered, washed with 1.2 M of HCl and ultrapure water three times, and dried at 60 °C to obtain graphene oxide (GO).

2.4. Preparation of Reduced Graphene Oxide–Rhodium Nanoparticle Electrode (Rh2O3–rGO/GCE)

The schematic diagram of the preparation and action mechanism of the Rh2O3–GO/GCE is given as Scheme 1. The Rh2O3–NP solution was prepared by dispersing the above NPs into 10 mL water, then ultrasound treatment for 2 h. The GO suspension was dispersed into water to prepare 1 mg mL−1 GO solution. Then, 100 μL of Rh2O3 nanoparticle solution was dispersed into 10 mL GO solution (1 mg mL−1). After ultrasound treatment for 30 min, the Rh2O3–GO nanoparticle composites were synthesized successfully. Before modifying the electrode, the GCE was polished to a mirror-like finish using 0.3 mm alumina slurry, and sonicated in distilled water and ethanol. The Rh2O3–GO solution (5 μL) was coated on the GCE surface and left to dry in air. Reduction of GO/GCE to rGO/GCE and Rh2O3–GO/GCE to Rh2O3–rGO/GCE was performed by cyclic voltammetry (CV) at a reduced potential of −1.2 V for 120 s in phosphate buffer solution.

2.5. Real Sample Preparation

Four beverage samples and one thermal paper sample were purchased from a local supermarket, to evaluate the performance of Rh2O3–rGO/GCE. After simple filtration for removing the fruit flesh, 800 μL beverage was mixed with PBS solution (9.2 mL) and directly analyzed by the proposed sensors. Thermal paper (1 g) was cut to pieces and extracted by ultrasound with 100 mL ethanol for 2 h. Then, 100 μL of extraction agent was mixed with PBS (9.9 mL) and analyzed by the proposed sensors. Afterward, the samples were spiked with BPA at three levels (5, 10 and 15 μM), and treated, following the above procedures, for recovery evaluation.

3. Results

3.1. Material Characterization

Pure GO exhibited a typical silk fold surface (Figure 1a), corresponding to a much higher surface area than GO. As for Rh2O3–NP, it displayed uniform three-dimensional characteristics composed of network structures (Figure 1b). The Rh2O3 nanoparticles were wrapped by the gauze-like GO, while a relatively smoother surface was observed in the Rh2O3–GO composite, compared with that of the Rh2O3–NPs (Figure 1c). In XRD spectra of rGO (Figure 1d), the peak at 2θ = 28.4° belonged to the characteristic reflection of graphite (002). The wide diffraction peaks at 2θ = 40° and 70° can be attributed to the diffractions of (100) and (214) planes of Rh2O3 (Figure 1d). Totally, the XRD pattern of Rh2O3–rGO indicated the successful preparation of the Rh2O3–rGO composite.

3.2. Electrochemical Properties of the Modified Electrodes

Electrochemical characterizations of the electrodes were investigated using [Fe(CN)6]3−/4− as probes. The cyclic voltammograms (CVs) of different electrodes in 1.0 mM [Fe(CN)6]3−/4− solution containing 0.1 M KCl at a scan rate of 100 mV s−1 are shown in Figure 2A. The smaller peak-to-peak separations (Ep) of Rh2O3–rGO/GCE (90 mV) than that of bare GCE (149 mV) and rGO/GCE (148 mV) can be referred to the faster electron transfer on the surface of Rh2O3–rGO/GCE. Additionally, the redox current of rGO/GCE was much larger than GCE. When Rh2O3–NPs were modified on rGO, the peak current increased rapidly. This indicated that Rh2O3–NPs should be involved for the enhanced electron transfer rate.

3.3. Electrocatalytic Behaviors of BPA

The CVs were performed in 0.1 M PBS (pH = 7.0) solutions containing 20 μM BPA, to investigate the electrochemical behaviors of the prepared electrodes at a scan rate of 100 mV s−1. There was only a relatively small oxidation peak (4.5 μA) at the potential of 0.57 V (Figure 2B, curve a). After coating with rGO (curve b), the peak current dramatically increased, due to the good adsorption efficiency of BPA onto rGO. Rh2O3–rGO/GCE (curve c) presented the greatest peak current of 30.5 μA, at the peak potential of about 0.52 V. Though peak potential shifted from 0.57 to 0.52 V, the results still indicated that Rh2O3–rGO could accelerate the electrocatalytic oxidation of BPA. This enhancement can be attributed to both the effective specific surface area of rGO, and the good electrocatalytic properties of Rh2O3–NPs.

3.4. Effects of Solution pH

The electrochemical responses of BPA on prepared Rh2O3–rGO/GCE under different pH were also studied by CV (Figure 3). The oxidation peak current increased gradually with pH from 5.5 to 7.0, and then it began to decrease with the pH continually increasing to 8.5. So, in the following studies, a pH of 7.0 was selected for the supporting electrolyte.

3.5. Effects of Scan Rate

To assess the electrocatalytic behaviors of BPA at Rh2O3–rGO/GCE, scan rates from 10 to 270 mV s−1 were studied (Figure 4). Oxidation peak current (I, μA) of BPA was linearly increased with the scan rate (Ʋ, mV s−1) with R2 = 0.9938, and the relationship can be expressed by the linear regression Equation (1).
I = 0.0489 × Ʋ + 2.7458,
The oxidation of BPA on the Rh2O3–rGO/GCE surface was an adsorption-controlled process. Meanwhile, the oxidation peak potential (E, V) increased linearly with natural logarithm of the scan rate (lnƲ, mV s−1) with R2 = 0.9931, and the Equation (2) is as follows:
E = 0.0246 × lnƲ + 0.4543,
As for an adsorption-controlled and totally irreversible electrode process, E was defined by Equation (3) [43].
E = Eθ + (RT/αnF) × ln((RTkθ)/αnF) + (RT/αnF) × lnƲ,
where α, kθ, n, υ, Eθ, R, T and F represent transfer coefficient, standard rate constant of the reaction, electron transfer number (involved in rate determining step), scan rate, formal redox potential, the gas constant, the absolute temperature, and the Faraday constant, respectively. According to Equation (2), the value of RT/αnF was 0.0246. Generally, α is assumed to be 0.5 in a totally irreversible electrode process [44]. Therefore, n was calculated to be 2.09, indicating that the electron transfer number for BPA oxidation is 2.

3.6. Analytical Performance of the Rh2O3–rGO/GCE

As shown in Figure 5, a linear relationship between oxidation current and BPA concentration within the range of 0.6–40 μM could be expressed as:
I = 3.1569 × C + 0.1002,
where I is oxidation peak current of BPA (μA), and C is the concentration of BPA in solution (μM). The limit of detection (LOD) was estimated to be 0.12 μM using the Equation (5)
LOD = (k × SD)/b,
where k is a constant, SD is the standard deviation of the intercept, and b is the slope of the calibration plot. In this paper, the values of k, SD, and b are 3, 0.1315 and 3.1569, respectively. Compared with some other electrodes, based on gold and copper oxide–graphene nanoparticles [42,45], the Rh2O3–rGO/GCE presented comparable performance in similar linear detection ranges (Table 1).
The stability, reproducibility, and selectivity of the Rh2O3–rGO/GCE were also evaluated. To test the stability of prepared electrode, Rh2O3–rGO/GCE was stored at 4 °C and examined three times per week by CV. After two weeks, the peak current retained 77.2% of the initial response. The Rh2O3–GO solution displayed excellent stability, being deposited at 4 °C within 30 days. Using this Rh2O3–GO solution to prepare Rh2O3–rGO/GCE, the response current for detecting BPA still remained within 98% of the initial currents. The good stability and reproducibility indicated the sensor should be appropriate for routine measurements of real samples. The relative standard deviation (RSD) for detecting BPA (n = 7) by one as-prepared sensor was less than 2.73%. The RSD for seven Rh2O3–rGO/GCEs fabricated in parallel was less than 3.22%.
The i–t curves of bisphenol A, by the addition of different inorganic ions and BPA analogs, are shown in Figure 6. The concentrations of inorganic ions, BPA, and BPA analogs are 200 µM, 20 µM, and 20 µM, respectively. Common inorganic ions (Ba2+, Cu2+, Ca2+, K+, Na+, OH, SO42− and Cl) have no effect on analyzing BPA, even when their levels are 10 times higher than BPA. Meanwhile, no obvious interferences from BPA analogs have been observed, such as phenol, hydroquinone, bisphenol B, bisphenol F, and bisphenol S. The RSDs for different interferences were all less than 3.53%.

3.7. Sample Analysis

The real samples spiked with BPA at 0–15 μM were successfully analyzed by the prepared Rh2O3–rGO/GCE. Before analysis, these drinks were filtered with an 0.45 μM hydrophilic membrane to remove the suspended particles. The detailed procedure has been described in Section 2.5. The recoveries of BPA from four kinds of beverages and thermal paper ranged from 93% to 98% (Table 2), indicating that the prepared Rh2O3–rGO modified electrode should be a useful sensor for analyzing BPA in aqueous solutions containing relatively complex interferences.

4. Conclusions

In this work, a Rh2O3–rGO modified GCE has been successfully fabricated for detecting BPA, which presents good catalytic performance. The enhanced electrocatalytic property can be attributed to the synergism of rGO and Rh2O3–NPs. Wide linear range, low detection limit, and good stability for detecting BPA have been obtained for this prepared Rh2O3–rGO/GCE sensor. More excitingly, BPA in some drinks could be directly and successfully detected after simple filtration. As a whole, a Rh2O3–rGO/GCE sensor, combining the catalytic properties of Rh2O3 and the conductivity of rGO, should be a useful supplemental strategy to common methods based on GC–MS and HPLC–MS, especially for rapid field detection.

Author Contributions

Conceptualization, Z.Z. and R.S.; formal analysis, R.S.; funding acquisition, A.L. and Z.Z.; investigation, X.Y.; methodology, A.L.; supervision, Z.Z.; validation, X.Y.; writing—original draft, A.L.; writing—review and editing, M.X.

Funding

This work was supported by the China Postdoctoral Science Foundation (2016M602210 and 2018T110715) and the Open Fund supported by Key Laboratory for Environmental Factors Control of Agro-product Quality Safety, Ministry of Agriculture.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hoekstra, E.J.; Simoneau, C. Release of bisphenol A from polycarbonate: A review. Crit. Rev. Food Sci. Nutr. 2013, 53, 386–402. [Google Scholar] [CrossRef] [PubMed]
  2. Niu, Y.; Zhang, J.; Wu, Y.; Shao, B. Analysis of bisphenol A and alkylphenols in cereals by automated on-line solid-phase extraction and liquid chromatography tandem mass spectrometry. J. Agric. Food Chem. 2012, 60, 6116–6122. [Google Scholar] [CrossRef] [PubMed]
  3. Makris, K.C.; Andra, S.S.; Jia, A.; Herrick, L.; Christophi, C.A.; Snyder, S.A.; Hauser, R. Association between water consumption from polycarbonate containers and bisphenol A intake during harsh environmental conditions in summer. Environ. Sci. Technol. 2013, 47, 3333–3343. [Google Scholar] [CrossRef] [PubMed]
  4. Latif, U.; Qian, J.J.; Can, S.; Dickert, F.L. Biomimetic Receptors for Bioanalyte Detection by Quartz Crystal Microbalances—From Molecules to Cells. Sensors 2014, 14, 23419–23438. [Google Scholar] [CrossRef] [PubMed]
  5. Rosenfeld, C.S. Bisphenol A and phthalate endocrine disruption of parental and social behaviors. Front. Neurosci. 2015, 9, 57. [Google Scholar] [CrossRef] [PubMed]
  6. Schug, T.T.; Janesick, A.; Blumberg, B.; Heindel, J.J. Endocrine disrupting chemicals and disease susceptibility. J. Steroid Biochem. 2011, 127, 204–215. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Bacle, A.; Thevenot, S.; Grignon, C.; Belmouaz, M.; Bauwens, M.; Teychene, B.; Venisse, N.; Migeot, V.; Dupuis, A. Determination of bisphenol A in water and the medical devices used in hemodialysis treatment. Int. J. Pharm. 2016, 505, 115–121. [Google Scholar] [CrossRef] [PubMed]
  8. Pouokam, G.B.; Ajaezi, G.C.; Mantovani, A.; Orisakwe, O.E.; Frazzoli, C. Use of Bisphenol A-containing baby bottles in Cameroon and Nigeria and possible risk management and mitigation measures: Community as milestone for prevention. Sci. Total Environ. 2014, 481, 296–302. [Google Scholar] [CrossRef]
  9. Mei, S.; Wu, D.; Jiang, M.; Lu, B.; Lim, J.M.; Zhou, Y.K.; Lee, Y.I. Determination of trace bisphenol A in complex samples using selective molecularly imprinted solid-phase extraction coupled with capillary electrophoresis. Microchem. J. 2011, 98, 150–155. [Google Scholar] [CrossRef]
  10. Zeng, J.; Kuang, H.; Hu, C.; Shi, X.; Yan, M.; Xu, L.; Wang, L.; Xu, C.; Xu, G. Effect of Bisphenol A on Rat Metabolic Profiling Studied by Using Capillary Electrophoresis Time-of-Flight Mass Spectrometry. Environ. Sci. Technol. 2013, 47, 7457–7465. [Google Scholar] [CrossRef]
  11. Zhang, X.F.; Zhu, D.; Huang, C.P.; Sun, Y.H.; Lee, Y.I. Sensitive detection of bisphenol A in complex samples by in-column molecularly imprinted solid-phase extraction coupled with capillary electrophoresis. Microchem. J. 2015, 121, 1–5. [Google Scholar] [CrossRef]
  12. Ahn, Y.G.; Shin, J.H.; Kim, H.Y.; Khim, J.; Lee, M.K.; Hong, J. Application of solid-phase extraction coupled with freezing-lipid filtration clean-up for the determination of endocrine-disrupting phenols in fish. Anal. Chim. Acta 2007, 603, 67–75. [Google Scholar] [CrossRef] [PubMed]
  13. Thomson, B.M.; Grounds, P.R. Bisphenol A in canned foods in New Zealand: An exposure assessment. Food Addit. Contam. 2005, 22, 65–72. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Stuart, J.D.; Capulong, C.P.; Launer, K.D.; Pan, X. Analyses of phenolic endocrine disrupting chemicals in marine samples by both gas and liquid chromatography-mass spectrometry. J. Chromatogr. A 2005, 1079, 136–145. [Google Scholar] [CrossRef] [PubMed]
  15. Jin, X.L.; Jiang, G.B.; Huang, G.L.; Liu, J.F.; Zhou, Q.F. Determination of 4-octylphenol, 4-nonylphenol and bisphenol A in surface waters from the Haihe River in Tianjin by gas chromatography–mass spectrometry with selected ion monitoring. Chemosphere 2004, 56, 1113–1119. [Google Scholar] [CrossRef] [PubMed]
  16. Shao, B.; Han, H.; Tu, X.; Huang, L. Analysis of alkylphenol and bisphenol A in eggs and milk by matrix solid phase dispersion extraction and liquid chromatography with tandem mass spectrometry. J. Chromatogr. B 2007, 850, 412–416. [Google Scholar] [CrossRef] [PubMed]
  17. Shao, B.; Han, H.; Li, D.; Ma, Y.; Tu, X.; Wu, Y. Analysis of alkylphenol and bisphenol A in meat by accelerated solvent extraction and liquid chromatography with tandem mass spectrometry. Food Chem. 2007, 105, 1236–1241. [Google Scholar] [CrossRef]
  18. Lei, Y.J.; Fang, L.Z.; Akash, M.S.H.; Liu, Z.M.; Shi, W.X.; Chen, S.Q. Development and comparison of two competitive ELISAs for the detection of bisphenol A in human urine. Anal. Methods 2013, 5, 6106–6113. [Google Scholar] [CrossRef]
  19. Ohkuma, H.; Abe, K.; Ito, M.; Kokado, A.; Kambegawa, A.; Maeda, M. Development of a highly sensitive enzyme-linked immunosorbent assay for bisphenol A in serum. Analyst 2002, 127, 93–97. [Google Scholar] [CrossRef]
  20. Maiolini, E.; Ferri, E.; Pitasi, A.L.; Montoya, A.; Di, G.M.; Errani, E.; Girotti, S. Bisphenol A determination in baby bottles by chemiluminescence enzyme-linked immunosorbent assay, lateral flow immunoassay and liquid chromatography tandem mass spectrometry. Analyst 2013, 139, 318–324. [Google Scholar] [CrossRef]
  21. Watabe, Y.; Kondo, T.; Imai, H.; Morita, M.; Tanaka, N.; Hosoya, K. Reducing Bisphenol A Contamination from Analytical Procedures to Determine Ultralow Levels in Environmental Samples Using Automated HPLC Microanalysis. Anal. Chem. 2004, 76, 105–109. [Google Scholar] [CrossRef]
  22. Xue, J.Q.; Li, D.W.; Qu, L.L.; Long, Y.T. Surface-imprinted core-shell Au nanoparticles for selective detection of bisphenol A based on surface-enhanced Raman scattering. Anal. Chim. Acta 2013, 777, 57–62. [Google Scholar] [CrossRef] [PubMed]
  23. Fan, H.; Li, Y.; Wu, D.; Ma, H.; Mao, K.; Fan, D.; Du, B.; Li, H.; Wei, Q. Electrochemical bisphenol A sensor based on N-doped graphene sheets. Anal. Chim. Acta 2012, 711, 24–28. [Google Scholar] [CrossRef] [PubMed]
  24. Yin, H.S.; Zhou, Y.L.; Ai, S.Y.; Chen, Q.P.; Zhu, X.B.; Liu, X.G.; Zhu, L.S. Sensitivity and selectivity determination of BPA in real water samples using PAMAM dendrimer and CoTe quantum dots modified glassy carbon electrode. J. Hazard. Mater. 2010, 174, 236–243. [Google Scholar] [CrossRef] [PubMed]
  25. Chang, C.M.; Chou, C.C.; Lee, M.-R. Determining leaching of bisphenol A from plastic containers by solid-phase microextraction and gas chromatography–mass spectrometry. Anal. Chim. Acta 2005, 539, 41–47. [Google Scholar] [CrossRef]
  26. Sadeghi, M.; Nematifar, Z.; Fattahi, N.; Pirsaheb, M.; Shamsipur, M. Determination of Bisphenol A in Food and Environmental Samples Using Combined Solid-Phase Extraction–Dispersive Liquid–Liquid Microextraction with Solidification of Floating Organic Drop Followed by HPLC. Food Anal. Method 2016, 9, 1814–1824. [Google Scholar] [CrossRef]
  27. Liu, S.H.; Xie, Q.L.; Chen, J.; Sun, J.Z.; He, H.; Zhang, X.K. Development and comparison of two dispersive liquid–liquid microextraction techniques coupled to high performance liquid chromatography for the rapid analysis of bisphenol A in edible oils. J. Chromatogr. A 2013, 1295, 16–23. [Google Scholar] [CrossRef] [PubMed]
  28. Wu, H.L.; Li, G.L.; Liu, S.C.; Hu, N.; Geng, D.D.; Chen, G.; Sun, Z.W.; Zhao, X.E.; Xia, L.; You, J.M. Monitoring the contents of six steroidal and phenolic endocrine disrupting chemicals in chicken, fish and aquaculture pond water samples using pre-column derivatization and dispersive liquid–liquid microextraction with the aid of experimental design methodology. Food Chem. 2016, 192, 98–106. [Google Scholar]
  29. Yao, K.; Wen, K.; Shan, W.; Xie, S.; Peng, T.; Wang, J.; Jiang, H.; Shao, B. Development of an immunoaffinity column for the highly sensitive analysis of bisphenol A in 14 kinds of foodstuffs using ultra-high-performance liquid chromatography tandem mass spectrometry. J. Chromatogr. B 2018, 1080, 50–58. [Google Scholar] [CrossRef]
  30. Li, K.H.; Zhu, L.X.; Liu, R.R.; Meng, W. Determination of bisphenol A in beer by immunoaffinity column tandem high-performance liquid chromatography. In Proceedings of the ICITMI, Shenzhen, China, 12–13 September 2015; pp. 230–234. [Google Scholar]
  31. Yan, W.; Li, Y.; Zhao, L.X.; Lin, J.M. Determination of Estrogens and Bisphenol A in Bovine Milk by Automated on-line C30 Solid-phase Extraction Coupled with High-performance Liquid Chromatography-mass Spectrometry. J. Chromatogr. A 2009, 1216, 7539–7545. [Google Scholar] [CrossRef]
  32. Liu, M.; Hashi, Y.; Pan, F.; Yao, J.; Song, G.; Lin, J.M. Automated on-line liquid chromatography-photodiode array-mass spectrometry method with dilution line for the determination of bisphenol A and 4-octylphenol in serum. J. Chromatogr. A 2006, 1133, 142–148. [Google Scholar] [CrossRef] [PubMed]
  33. Messaoud, N.B.; Ghica, M.E.; Dridi, C.; Ali, M.B.; Brett, C.M.A. Electrochemical sensor based on multiwalled carbon nanotube and gold nanoparticle modified electrode for the sensitive detection of bisphenol A. Sens. Actuators B Chem. 2017, 253, 513–522. [Google Scholar] [CrossRef]
  34. Wang, L.; Zhang, Y.J.; Wu, A.G.; Wei, G. Designed graphene-peptide nanocomposites for biosensor applications: A review. Anal. Chim. Acta 2017, 985, 24–40. [Google Scholar] [CrossRef] [PubMed]
  35. Wang, L.; Wu, A.G.; Wei, G. Graphene-based aptasensors: From molecule–interface interactions to sensor design and biomedical diagnostics. Analyst 2018, 143, 1526–1543. [Google Scholar] [CrossRef] [PubMed]
  36. Shi, R.G.; Liang, J.; Zhao, Z.Z.; Liu, Y.; Liu, A.F. In Situ Determination of Bisphenol A in Beverage Using a Molybdenum Selenide/Reduced Graphene Oxide Nanoparticle Composite Modified Glassy Carbon Electrode. Sensors 2018, 18, 1660. [Google Scholar] [CrossRef] [PubMed]
  37. Song, W.; Jansen, A.P.; Degirmenci, V.; Ligthart, D.A.; Hensen, E.J. A computational study of the mechanism of CO oxidation by a ceria supported surface rhodium oxide layer. Chem. Commun. 2013, 49, 3851–3853. [Google Scholar] [CrossRef]
  38. Ligthart, D.A.; van Santen, R.A.; Hensen, E.J. Supported rhodium oxide nanoparticles as highly active CO oxidation catalysts. Angew. Chem. Int. Ed. 2011, 50, 5306–5310. [Google Scholar] [CrossRef]
  39. Kim, Y.L.; Ha, Y.; Lee, N.S.; Kim, J.G.; Baik, J.M.; Lee, C.; Yoon, K.; Lee, Y.; Kim, M.H. Hybrid architecture of rhodium oxide nanofibers and ruthenium oxide nanowires for electrocatalysts. J. Alloys Compd. 2016, 663, 574–580. [Google Scholar] [CrossRef]
  40. Verónica, R.P.; Sonia, P.E.; María, J.L.G.; Agustín, B.L. Preparation, characterisation and N2O decomposition activity of honeycomb monolith-supported Rh/Ce0.9Pr0.1O2 catalysts. Appl. Catal. B Environ. 2011, 107, 18–25. [Google Scholar]
  41. Bai, J.; Han, S.H.; Peng, R.L.; Zeng, J.H.; Jiang, J.X.; Chen, Y. Ultrathin Rhodium Oxide Nanosheet Nanoassemblies: Synthesis, Morphological Stability, and Electrocatalytic Application. ACS Appl. Mater. Interfaces 2017, 9, 17195–17200. [Google Scholar] [CrossRef]
  42. Shi, R.G.; Liang, J.; Zhao, Z.S.; Liu, A.F.; Tian, Y. An electrochemical bisphenol A sensor based on one step electrochemical reduction of cuprous oxide wrapped graphene oxide nanoparticles modified electrode. Talanta 2017, 169, 37–43. [Google Scholar] [CrossRef] [PubMed]
  43. Laviron, E. Adsorption, autoinhibition and autocatalysis in polarography and in linear potential sweep voltammetry. J. Electroanal. Chem. Interfacial Electrochem. 1974, 52, 355–393. [Google Scholar] [CrossRef]
  44. Lin, Y.Q.; Liu, K.Y.; Liu, C.Y.; Yin, L.; Kang, Q.; Li, L.B.; Li, B. Electrochemical sensing of bisphenol A based on polyglutamic acid/amino-functionalised carbon nanotubes nanocomposite. Electrochim. Acta 2014, 133, 492–500. [Google Scholar] [CrossRef]
  45. Zhou, L.; Wang, J.P.; Li, D.J.; Li, Y.B. An electrochemical aptasensor based on gold nanoparticles dotted graphene modified glassy carbon electrode for label-free detection of bisphenol A in milk samples. Food Chem. 2014, 162, 34–40. [Google Scholar] [CrossRef] [PubMed]
  46. Kanagavalli, P.; Kumar, S.S. Stable and Sensitive Amperometric Determination of Endocrine Disruptor Bisphenol A at Residual Metal Impurities Within SWCNT. Electroanalysis 2018, 30, 445–452. [Google Scholar] [CrossRef]
  47. Li, H.Y.; Wang, W.; Lv, Q.; Xi, G.C.; Bai, H.; Zhang, Q. Disposable paper-based electrochemical sensor based on stacked gold nanoparticles supported carbon nanotubes for the determination of bisphenol A. Electrochem. Commun. 2016, 68, 104–107. [Google Scholar] [CrossRef]
  48. Zhan, T.R.; Song, Y.; Tan, Z.W.; Hou, W.G. Electrochemical bisphenol A sensor based on exfoliated Ni2Al-layered double hydroxide nanosheets modified electrode. Sens. Actuators B Chem. 2017, 238, 962–971. [Google Scholar] [CrossRef]
Scheme 1. Schematic diagram for electrochemical synthesis of the Rh2O3–rGO electrode, and the electrochemical oxidation process of BPA by Rh2O3–rGO/GCE.
Scheme 1. Schematic diagram for electrochemical synthesis of the Rh2O3–rGO electrode, and the electrochemical oxidation process of BPA by Rh2O3–rGO/GCE.
Applsci 08 02535 sch001
Figure 1. SEM images of (a) GO, (b) Rh2O3 and (c) Rh2O3–GO; (d) XRD patterns of rGO, Rh2O3 and Rh2O3–rGO nanoparticles.
Figure 1. SEM images of (a) GO, (b) Rh2O3 and (c) Rh2O3–GO; (d) XRD patterns of rGO, Rh2O3 and Rh2O3–rGO nanoparticles.
Applsci 08 02535 g001
Figure 2. (A) CVs of (a) bare GCE, (b) rGO/GCE, and (c) Rh2O3–rGO/GCE in the solution of 1.0 mM [Fe(CN)6]3−/4− (1:1) and 0.1 M KCl; (B) CVs of (a) bare GCE, (b) rGO/GCE, and (c) Rh2O3–rGO/GCE in the solution of 20 μM BPA.
Figure 2. (A) CVs of (a) bare GCE, (b) rGO/GCE, and (c) Rh2O3–rGO/GCE in the solution of 1.0 mM [Fe(CN)6]3−/4− (1:1) and 0.1 M KCl; (B) CVs of (a) bare GCE, (b) rGO/GCE, and (c) Rh2O3–rGO/GCE in the solution of 20 μM BPA.
Applsci 08 02535 g002
Figure 3. (a) CVs of Rh2O3–rGO/GCE detected BPA with different pH; (b) effects of pH on the peak current of BPA at Rh2O3–rGO/GCE.
Figure 3. (a) CVs of Rh2O3–rGO/GCE detected BPA with different pH; (b) effects of pH on the peak current of BPA at Rh2O3–rGO/GCE.
Applsci 08 02535 g003
Figure 4. (a) CVs of Rh2O3–rGO/GCE detected 20 μM BPA with scan rate from 10 to 270 mV s−1; (b) Dependence of the oxidation peak current on the scan rate; (c) The relationship between E and ln Ʋ.
Figure 4. (a) CVs of Rh2O3–rGO/GCE detected 20 μM BPA with scan rate from 10 to 270 mV s−1; (b) Dependence of the oxidation peak current on the scan rate; (c) The relationship between E and ln Ʋ.
Applsci 08 02535 g004
Figure 5. (a) CVs of Rh2O3–rGO/GCE with BPA concentration from 0.6 to 40 μM; (b) Relationship between the oxidation peak current and the concentration of BPA (from 0.6 to 40 μM).
Figure 5. (a) CVs of Rh2O3–rGO/GCE with BPA concentration from 0.6 to 40 μM; (b) Relationship between the oxidation peak current and the concentration of BPA (from 0.6 to 40 μM).
Applsci 08 02535 g005
Figure 6. The i–t curves of bisphenol A by the addition of (a) different inorganic ions and (b) BPA analogs.
Figure 6. The i–t curves of bisphenol A by the addition of (a) different inorganic ions and (b) BPA analogs.
Applsci 08 02535 g006
Table 1. Comparison of the Rh2O3–rGO and other sensors for BPA determination.
Table 1. Comparison of the Rh2O3–rGO and other sensors for BPA determination.
SensorsLinear Range (μM)LOD (μM)Reference
Residual metal impurities within SWCNT electrode10–1007.3[46]
Gold nanoparticles supported carbon nanotubes electrode0.87–870.13[47]
Gold nanoparticles dotted graphene electrode0.01–100.005[45]
Exfoliated Ni2Al layered double hydroxide nanosheets electrode0.02–1.510.007[48]
Copper oxide and graphene electrode0.1–800.053[42]
Rhodium oxide and graphene electrode0.6–400.12This work
Table 2. The results of determined BPA in practical samples.
Table 2. The results of determined BPA in practical samples.
SampleAdded (μM)Found (μM)Recovery (%)
Beverage 103.22 ± 0.09 1-
58.10 ± 0.1898
1012.66 ± 0.4394
1517.22 ± 0.6293
Beverage 204.69 ± 0.09-
59.46 ± 0.1895
1014.34 ± 0.3797
1519.49 ± 0.4399
Beverage 303.96 ± 0.04-
58.63 ± 0.0993
1013.85 ± 0.2198
1517.80 ± 0.4993
Beverage 400-
54.86 ± 0.1097
109.57 ± 0.3096
1514.73 ± 0.2498
Thermal paper06.46 ± 0.23-
511.33 ± 0.1897
1013.85 ± 0.2198
1520.95 ± 0.5497
1 Average ± Standard deviation.

Share and Cite

MDPI and ACS Style

Shi, R.; Yuan, X.; Liu, A.; Xu, M.; Zhao, Z. Determination of Bisphenol A in Beverages by an Electrochemical Sensor Based on Rh2O3/Reduced Graphene Oxide Composites. Appl. Sci. 2018, 8, 2535. https://doi.org/10.3390/app8122535

AMA Style

Shi R, Yuan X, Liu A, Xu M, Zhao Z. Determination of Bisphenol A in Beverages by an Electrochemical Sensor Based on Rh2O3/Reduced Graphene Oxide Composites. Applied Sciences. 2018; 8(12):2535. https://doi.org/10.3390/app8122535

Chicago/Turabian Style

Shi, Rongguang, Xiaoli Yuan, Aifeng Liu, Mengmeng Xu, and Zongshan Zhao. 2018. "Determination of Bisphenol A in Beverages by an Electrochemical Sensor Based on Rh2O3/Reduced Graphene Oxide Composites" Applied Sciences 8, no. 12: 2535. https://doi.org/10.3390/app8122535

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