Next Article in Journal
Ex Vivo and In Silico Approaches of Tracheal Relaxation through Calcium Channel Blockade of 6-Aminoflavone and Its Toxicological Studies in Murine Models
Next Article in Special Issue
Nucleic Acid Aptamer-Based Biosensors: A Review
Previous Article in Journal
The Role of ATP-Binding Cassette Proteins in Stem Cell Pluripotency
Previous Article in Special Issue
Formulation and Optimization of Repaglinide Nanoparticles Using Microfluidics for Enhanced Bioavailability and Management of Diabetes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electrochemical Sensor for Simple and Sensitive Determination of Hydroquinone in Water Samples Using Modified Glassy Carbon Electrode

1
Department of Chemistry, Graduate University of Advanced Technology, Kerman 76311-33131, Iran
2
Environment Department, Institute of Science and High Technology and Environmental Sciences, Graduate University of Advanced Technology, Kerman 76311-33131, Iran
3
Polymer Department, Graduate University of Advanced Technology, Kerman 76311-33131, Iran
*
Author to whom correspondence should be addressed.
Biomedicines 2023, 11(7), 1869; https://doi.org/10.3390/biomedicines11071869
Submission received: 10 April 2023 / Revised: 21 June 2023 / Accepted: 25 June 2023 / Published: 30 June 2023
(This article belongs to the Special Issue Bio-Nano Interfaces: From Biosensors to Nanomedicines)

Abstract

:
This study addressed the use of manganese dioxide nanorods/graphene oxide nanocomposite (MnO2 NRs/GO) for modifying a glassy carbon electrode (GCE). The modified electrode (MnO2 NRs/GO/GCE) was used as an electrochemical sensor for the determination of hydroquinone (HQ) in water samples. Differential pulse voltammetry (DPV), cyclic voltammetry (CV), and chronoamperometry were used for more analysis of the HQ electrochemical behavior. Analyses revealed acceptable electrochemical functions with lower transfer resistance of electrons and greater conductivity of the MnO2 NRs/GO/GCE. The small peak-to-peak separation is an indication of a rapid electron transfer reaction. Therefore, this result is probably related to the effect of the MnO2 NRs/GO nanocomposite on the surface of GCE. In the concentration range of 0.5 μM to 300.0 μM with the detection limit as 0.012 μM, there was linear response between concentration of HQ and the current. The selectivity of the modified electrode was determined by detecting 50.0 μM of HQ in the presence of various interferent molecules. At the end, the results implied the acceptable outcome of the prepared electrode for determining HQ in the water samples.

1. Introduction

The existence of any form of life on earth is dependent on water, which is vital. Water contamination is the primary cause of the water crisis. Water contamination is primarily concentrated in areas of industrialization, urbanization, agricultural activities, etc. The level of water contamination is affected by the abundance of pollutants, their ecological impact, and the intended use of the water. Water contamination is a significant issue that has adverse effects on both the ecosystem and human health. Exposure to water contamination over an extended period is a leading factor in causing various health problems and diseases. Additionally, it can disrupt ecological environments, including plant and animal life, and affect food chains. Protecting water systems to maintain public health and detecting contamination are crucial factors in achieving water security [1,2].
Hydroquinone (HQ) (1,4-dihydroxybenze), as one of the major isomers of di-hydroxybenzene, has widespread application for preparation of photo-stabilizers, dyes, cosmetics, plasticizers, drugs, and pesticides. In fact, HQ has an extensive use in producing food additives, antioxidants, as well as hair dyes. However, since dihydroxybenze isomers have lower degradability and great toxicity in an ecological environment (according to the World Health Organization (WHO)), HQ is toxic to several aquatic organisms even at concentrations below 1 mg L−1 (9.1 µmol L−1) [3]. A main topic, which has been largely investigated in the environmental pollutants, is how to selectively detect dihydroxybenzene [4,5]. Hence, experts in the field have provided several techniques of HQ detection such as surface-enhanced Raman spectroscopy [6], spectrophotometry [7], capillary electrophoresis [8], chemiluminescence [9], calorimetric sensor [10], gas chromatography-mass spectrometry [11], and high-performance liquid chromatography [12]. However, a number of the mentioned techniques suffer from complex operations, costly equipment, hard separation procedure, and high toxic organic solvents, but since HQ has an electroactivity property, electrochemical approaches could be used for its detection [13]. Hence, the development of simple and sensitive electrochemical sensors for the determination of HQ has received considerable attention [14,15,16,17,18,19,20,21,22].
Electrochemical methods have been increasingly attractive for features such as faster responses, simplified operations, higher sensitivity, good selectivity, insignificant content of the samples, and inexpensiveness [23,24,25,26,27,28,29,30]. Nonetheless, direct oxidation of HQ on the bare electrode adsorbs the oxidized products over the electrodes’ surfaces, leading to the unstable response of the electrode and electrode fouling [31]. Additionally, one of the challenges resulting from overlapping their oxidation–reduction peak has been the direct detection of the dihydroxybenze isomers on the conventional electrodes such as GCEs [32,33].
It is well known that electrochemical performances can be tuned by tailoring the material compositions and surface properties of modified electrodes [34,35,36]. Moreover, analytical response due to the use of the modifiers is generally greater and thus the sensitivity of the electrode would largely improve. Hence, the well-organized modified electrodes must be immediately developed for detecting the analytes [37,38,39].
In recent years, the use of nanomaterials in various fields has progressed considerably [40,41,42,43,44,45,46,47,48]. On the one hand, scholars have utilized nanostructures as the modification materials because they have satisfactory features, including larger surface areas, and smaller sizes [49,50,51,52]. The design and fabrication of the nanomaterial-based electrochemical sensors by amplifying the signal and reducing overvoltage play an important role in electroanalysis. Graphene oxide (GO), one of the single layers of graphite oxide, was initially created by the graphite treatment with robust aqueous oxidizing agents. Studies have shown very good mechanical strength, higher mobility of the charge carriers, greater thermal conductivity, faster electron transport, and higher surface areas of the two-dimensional matter. These indicate its satisfactory features for possible uses in numerous areas [53,54,55]. Furthermore, it is possible to adjust graphene’s surface properties using chemical modifications, resulting in simple application in composite materials. Some studies have also referred to the synergetic effect of graphene and inorganic particles, resulting in the very good features and greater performance of the graphene-supported hybrids [56,57,58]. Manganese dioxide (MnO2), as one of the major functional materials, displays a higher surface area, inexpensiveness, stronger catalytic activities, as well as acceptable biocompatibility [59,60,61]. It should be noted that integrating the MnO2 nanoparticles (NPs) into the GO sheets potentially would increase the specific surface area and accelerate the transfer of electrons, leading to multiple channels and higher conductivity to diffuse the electrolyte ions [62,63]. So far, in some previous works, a GO/MnO2 nanocomposite with various morphologies of MnO2 has been used as an electrode material in the design and fabrication of electrochemical sensors for the determination of various analytes, and they showed good performance [22,64,65,66].
Herein, we developed the MnO2 NRs/GO/GCE for the sensitive detection of HQ in water samples. The results demonstrated that MnO2 NRs and GO had a good synergistic effect on the electrochemical oxidation of HQ. The other advantages of the sensor are excellent reproducibility, repeatability, stability, and high selectivity.

2. Materials and Methods

2.1. Instrumentation and Materials

According to the research design, we employed a General-Purpose Electrochemical System (GPES) 4.9 and Potentiostat/Galvanostat (Autolab PGSTAT302N, made in The Netherlands) for performing the electrochemical experiments.
All the electrochemical studies were performed at 25 ± 1 °C. A three electrode assembly was employed in the experiment in a 15 mL borosilicate glass cell containing GCE as a working electrode, which was bought from Azar electrode Co. (made in Urmia, Iran), a Pt wire as the counter electrode, and an Ag/AgCl (3 M KCl) reference electrode. The pH was also measured and buffer solution was prepared using a digital pH meter (Metrohm AG, Herisau, Switzerland, pH Lab 713). Deionized water used in each experiment was also taken from a Millipore Direct-Q® 8 UV (ultra-violet) (Millipore, Germany). The morphological and elemental analyses of the prepared materials were carried out by a MIRA3 SEM (Tescan, Brno, Czech Republic).
It should be noted that the precursors for synthesizing the MnO2 NRs/GO nanocomposite, HQ, as well as other chemicals were also of analytical grade. It is noted that they were received from Merck and Sigma-Aldrich chemical companies. Furthermore, phosphate buffer solutions (0.1 M, PBS) of various levels of pH were provided via mixing the suitable contents of sodium hydroxide (NaOH) solution and phosphoric acid (H3PO4), also known as orthophosphoric acid under the pH-meter.

2.2. Synthesis of MnO2 NRs/GO Nanocomposite

For preparation of MnO2 NRs/GO nanocomposite, 20 mg GO powder was dispersed in deionized water (30 mL) under ultrasonic condition to form a homogeneous suspension. Afterwards, 0.316 gr of KMnO4 was poured into the GO suspension and 3.0 M HCl with strong stirring and continued stirring for half an hour. Next, this mixture was transported into the Teflon-lined autoclave at 160 °C for 6 h. Then, cooling was performed for collecting a black product through centrifugation and washed with ethanol and deionized water many times. After 12 h, an oven was used to dry MnO2 NRs/GO at 60 °C.

2.3. Modified Electrode Preparation

For the modification of GCE by using MnO2 NRs/GO nanocomposite, the drop-casting technique was employed. Then, 1 mg of the synthesized MnO2 NRs/GO was distributed in 1 mL of the deionized water in ultrasonication conditions for 30 min and the cast was provided by putting the 4 µL water/MnO2 NRs/GO nanocomposite suspensions on GCE and drying at ambient temperature.
The surface areas of the MnO2 NRs/GO/GCE and the bare electrode were obtained by CV using 1.0 mM K3Fe(CN)6 at several scan rates. By using the Randles–Sevcik equation, the value of the surface area for MnO2 NRs/GO/GCE (0.119 cm2) was 3.8 times greater than bare GCE.

2.4. Preparation of Water Specimens

River water and tap water were also sampled, filtrated with a membrane filter, and poured into 0.1 M PBS (pH = 7.0). At last, the HQ contents were measured in the water specimens using the as-developed protocol according to standard addition method.

3. Results and Discussions

3.1. Characterization of MnO2 NRs/GO Nanocomposite by Field Emission Scanning Electron Microscopy (FE-SEM)

FE-SEM analyses were conducted for studying the morphology and structural properties of the GO and MnO2 NRs/GO nanocomposite (Figure 1). The FE-SEM image of GO clearly shows that GO has a sheet structure (Figure 1a). In addition, the FE-SEM image of the nanocomposite (Figure 1b) represented the distribution and deposition of the MnO2 rod-shaped nanostructures on the GO sheets.

3.2. Electrochemical Behaviors of HQ on the MnO2 NRs/GO/GCE

Studies revealed that electrochemical behaviors of HQ on the modified electrodes were followed by exchanges of protons. Therefore, the DPV technique was used for determining the effects of pH on the electrochemical response of MnO2 NRs/GO/GCE for HQ determination in various pHs of the PBS in ranges between 2.0 and 9.0, and concluded that a neutral medium would be more suitable to detect HQ electrochemically (Figure 2).
Figure 3 presents the typical cyclic voltametric response of different electrodes in 0.1 M PBS (pH = 7.0) in the absence and presence of 200.0 μM HQ. According to Curve a, no peak was observed on the surface of modified electrode in the absence of HQ.
The results of electrochemical responses of 200.0 μM HQ in 0.1 M PBS (pH = 7.0) at the surface of bare GCE (b), MnO2 NRs/GCE (c), GO/GCE (d), and MnO2 NRs/GO/GCE (e) are shown in Table 1. Analysis indicated an enhancement in the cathodic and anodic peak currents by modifying the electrode surface and decrease of over potential. According to the results, MnO2 NRs/GO/GCE developed the greatest cathodic and anodic peak currents, demonstrating the maximum activities of the electrode surface for HQ redox reaction.

3.3. Scan Rate Exploration

For studying the mechanisms required for the electrocatalytic redox reaction, the effects of the scanning rate (υ) on the reaction of 120.0 µM HQ on MnO2 NRs/GO/GCE ranged between 5 and 400 mVs−1. Considering Figure 4, as the scan rate enhanced, a linear increase in the cathodic and anodic peak currents (Ipc, Ipa) was observed. The inset of Figure 4 depicts the plots of the redox peak currents as a function of the square root of the scan rate (υ1/2) for HQ. The figure represents a linear increase in the redox peak currents with the square root of the scanning rates for different scan rates. Ultimately, analyses indicated the control of the electrode process through diffusion.

3.4. Chronoamperometric Determinations

In this step, chronoamperometry was employed for HQ on the MnO2 NRs/GO/GCE surface (Figure 5) and the potential of the working electrode was set at 0.25 V for chronoamperometric measurements of various concentrations of HQ on the modified electrode surface. Then, through chronoamperometry, the diffusion coefficient (D) of HQ was specified in the aqueous solution by the Cottrell equation:
I = nFACbD1/2π−1/2t−1/2
where C represents the known concentration, D stands for the apparent diffusion coefficient, and A refers to the electrode area. Experimental plots of I vs. t −1/2 were also utilized for distinct concentrations of HQ (see Figure 5A) and slopes were drawn of the final straight lines versus the HQ concentration (Figure 5B). Finally, D was calculated according to the slope of the resulting plots with the Cottrell equation (8.9 × 10−6 cm2/s).

3.5. Calibration Plot

Differential pulse voltammetry (DPV) in the optimum experimental conditions was employed to quantitatively analyze the concentration of HQ in the PBS (0.1 M, pH = 7.0) on MnO2 NRs/GO/GCE. Figure 6 depicts the voltammograms for distinct concentrations of HQ. Considering the figure, as the HQ concentration enhanced, an increase in the peak currents was observed. Furthermore, the peak current linearly was related to the concentration of HQ in ranges between 0.5 and 300.0 µM (see Figure 6 (Inset)). For calibration curve of the modified electrode with an acceptable determination coefficient of 0.9996 and the lowest limit of detection of 0.012 µM, see the inset of Figure 6. The comparison of MnO2 NRs/GO/GCE sensor with other sensors for the determination of HQ is listed in Table 2. It can be seen that the MnO2 NRs/GO/GCE offers a proper linear range and a lower detection limit than some modified electrodes. Hence, our proposed method gives a straight and faster means of HQ detection in the samples.

3.6. Interference Studies

As a general principle, the relative error in the measurement is controlled at approximately ±5% and is considered to have no interference. To evaluate the selectivity of the fabricated MnO2 NRs/GO/GCE sensor, the influence of organic molecules and several common ions were assayed for 50.0 µM HQ. The results suggested that 120-fold of Pb2+, NH4+, Mg2+, Ca2+, NO3, Na+, Cl, K+, Al3+, and Fe3+ ions, 50-fold of citric acid, vitamin B12, CH3COOH and 5-fold excess of uric acid did not interfere with the determination of HQ (the relative errors were within ±5%). However, catechol, dopamine, and ascorbic acid in equal molar concentrations showed interference.

3.7. Stability, Repeatability, and Reproducibility Studies

In order to determine the stability of the MnO2 NRs/GO/GCE, its electrocatalytic response to 50.0 μM HQ in 0.1 M PBS was monitored every day (1 to 14 day intervals). The electrode presented consistent voltametric responses to HQ during the 14-day storage period. About 96.1% of initial response current was maintained after 14 days of its consecutive use and this result denoted the admissible stability of the proposed modified electrode (MnO2 NRs/GO/GCE).
Repeatability and reproducibility of the proposed MnO2 NRs/GO-based sensor have been evaluated using voltametric studies in 0.1 M PBS (pH = 7.0) containing 50.0 µM HQ. The MnO2 NRs/GO/GCE sensor presented appreciable repeatability with relative standard deviation (R.S.D) of 4.9% for 6 repetitive measurements performed using an individua electrode.
To check the reproducibility, the five MnO2 NRs/GO/GCE were applied in the determination of HQ. Experiments with 3.3% (%RSD) for 50 µM HQ in 0.1 M PBS (pH = 7.0) were done, which showed the acceptable reproducibility of the fabricated sensor for determine HQ.

3.8. Application of the MnO2 NRs/GO/GCE Sensor for the Determination of HQ in Real Samples (River Water and Tap Water)

The capability of the electrocatalytic oxidation of HQ in the real samples was studied by voltammetry in the water samples (river water and tap water). Table 3 reports the results. As seen, the method recoveries ranged between 97.3 and 101.4%, reflecting the capability of MnO2 NRs/GO/GCE for voltametric detection of HQ with acceptable reproducibility.

4. Conclusions

In this study, a sensitive electrochemical sensor was made based on the MnO2 NRs/GO/GCE to detect HQ. It was produced through a simple drop-casting of the MnO2 NRs/GO nanocomposite dispersion over the GCE surface. As the synergistic effects of GO and MnO2 NRs were applied, HQ’s redox peak currents largely enhanced in comparison to the bare GCE. According to the findings, MnO2 NRs/GO/GCE had very good sensing functions to determine HQ, with a linear range (0.5 µM–300.0 µM) and lower LOD of 0.012 µM. Finally, the MnO2 NRs/GO/GCE could be satisfactorily utilized for HQ detection in the real samples. The other advantages of the sensors are excellent reproducibility, repeatability, stability, and high selectivity.

Author Contributions

Formal analysis, P.K.-K.; Writing—original draft, P.K.-K., H.B. and S.M.; Supervision, H.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Apetrei, C.; Iticescu, C.; Georgescu, L.P. Multisensory system used for the analysis of the water in the lower area of River Danube. Nanomaterials 2019, 9, 891. [Google Scholar] [CrossRef] [Green Version]
  2. Ghrissi, H.; Veloso, A.C.; Marx, Í.M.; Dias, T.; Peres, A.M. A potentiometric electronic tongue as a discrimination tool of water-food indicator/contamination bacteria. Chemosensors 2021, 9, 143. [Google Scholar] [CrossRef]
  3. Martoni, L.V.; Gomes, N.O.; Prado, T.M.; Calegaro, M.L.; Oliveira, O.N., Jr.; Machado, S.A.; Raymundo-Pereira, P.A. Carbon spherical shells in a flexible photoelectrochemical sensor to determine hydroquinone in tap water. J. Environ. Chem. Eng. 2022, 10, 107556. [Google Scholar] [CrossRef]
  4. Nordlund, J.J.; Grimes, P.E.; Ortonne, J.P. The safety of hydroquinone. J. Eur. Acad. Dermatol. Venereol. 2006, 20, 781–787. [Google Scholar] [CrossRef] [PubMed]
  5. Li, Z.; Zhou, W.; Zhang, L.; Jia, Y.; Liu, Y. Simultaneous determination of catechol and hydroquinone using gold nanoparticles/poly (p-aminobenzenesulfonic acid)/multi-walled carbon nanotubes modified glassy carbon electrode. J. Electrochem. Soc. 2022, 169, 107504. [Google Scholar] [CrossRef]
  6. Cabrera-Alonso, R.; Guevara, E.; Ramírez-Elías, M.G.; Moncada, B.; González, F.J. Surface-enhanced Raman scattering of hydroquinone assisted by gold nanorods. J. Nanophotonics 2019, 13, 036006. [Google Scholar] [CrossRef]
  7. Bhanger, M.I.; Niaz, A.; Shah, A.; Rauf, A. Ultra-trace level determination of hydroquinone in waste photographic solutions by UV–vis spectrophotometry. Talanta 2007, 72, 546–553. [Google Scholar]
  8. Lin, Y.H.; Yang, Y.H.; Wu, S.M. Experimental design and capillary electrophoresis for simultaneous analysis of arbutin, kojic acid and hydroquinone in cosmetics. J. Pharm. Biomed. Anal. 2007, 44, 279–282. [Google Scholar] [CrossRef]
  9. Zhao, L.; Lv, B.; Yuan, H.; Zhou, Z.; Xiao, D. A sensitive chemiluminescence method for determination of hydroquinone and catechol. Sensors 2007, 7, 578–588. [Google Scholar] [CrossRef] [Green Version]
  10. Wolf, A.; Hüttl, R.; Wolf, G. Enzymatic reactions for the calorimetric detection of phenolic compounds. J. Therm. Anal. Calorim. 2000, 61, 37–44. [Google Scholar] [CrossRef]
  11. Jurica, K.; Karaconji, I.B.; Šegan, S.; Milojković-Opsenica, D.; Kremer, D. Quantitative analysis of arbutin and hydroquinone in strawberry tree (Arbutus unedo L., Ericaceae) leaves by gas chromatography-mass spectrometry. Arh. Hig. Rada Toksikol. 2015, 66, 197–202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Gao, W.; Legido-Quigley, C. Fast and sensitive high performance liquid chromatography analysis of cosmetic creams for hydroquinone, phenol and six preservatives. J. Chromatogr. A 2011, 1218, 4307–4311. [Google Scholar] [CrossRef] [PubMed]
  13. Cotchim, S.; Promsuwan, K.; Dueramae, M.; Duerama, S.; Dueraning, A.; Thavarungkul, P.; Limbut, W. Development and application of an electrochemical sensor for hydroquinone in pharmaceutical products. J. Electrochem. Soc. 2020, 167, 155528. [Google Scholar] [CrossRef]
  14. Park, J.; Kim, J.; Min, A.; Choi, M.Y. Fabrication of nonenzymatic electrochemical sensor based on Zn@ZnO core-shell structures obtained via pulsed laser ablation for selective determination of hydroquinone. Environ. Res. 2022, 204, 112340. [Google Scholar] [CrossRef] [PubMed]
  15. Qi, H.; Zhang, C. Simultaneous determination of hydroquinone and catechol at a glassy carbon electrode modified with multiwall carbon nanotubes. Electroanalysis 2005, 17, 832–838. [Google Scholar] [CrossRef]
  16. Xia, Y.; Wang, K.; Shi, Y.; Gui, X.; Lv, C.; Tao, H. Reduced graphene oxide cross-linked L-cysteine modified glassy carbon electrode for detection of environmental pollutant of hydroquinone. FlatChem 2021, 25, 100214. [Google Scholar] [CrossRef]
  17. Chen, T.; Xu, J.; Arsalan, M.; Sheng, Q.; Zheng, J.; Cao, W.; Yue, T. Controlled synthesis of Au@ Pd core-shell nanocomposites and their application for electrochemical sensing of hydroquinone. Talanta 2019, 198, 78–85. [Google Scholar] [CrossRef]
  18. Zhao, D.M.; Zhang, X.H.; Feng, L.J.; Jia, L.; Wang, S.F. Simultaneous determination of hydroquinone and catechol at PASA/MWNTs composite film modified glassy carbon electrode. Colloids Surf. B Biointerfaces 2009, 74, 317–321. [Google Scholar] [CrossRef]
  19. Chen, T.W.; Palanisamy, S.; Chen, S.M.; Velusamy, V.; Liu, Y.H.; Tseng, T.W.; Liu, X. Sensitive and low-potential electrochemical detection of hydroquinone using a nanodiamond modified glassy carbon electrode. Int. J. Electrochem. Sci. 2017, 12, 8021–8032. [Google Scholar] [CrossRef]
  20. Zou, J.; Ma, J.; Zhang, Y.; Huang, L.; Wan, Q. A hydroquinone sensor based on a new nanocrystals modified electrode. J. Chem. Technol. Biotechnol. 2014, 89, 259–264. [Google Scholar] [CrossRef]
  21. Chen, L.; Tang, Y.; Wang, K.; Liu, C.; Luo, S. Direct electrodeposition of reduced graphene oxide on glassy carbon electrode and its electrochemical application. Electrochem. Commun. 2011, 13, 133–137. [Google Scholar] [CrossRef]
  22. Gan, T.; Sun, J.; Huang, K.; Song, L.; Li, Y. A graphene oxide–mesoporous MnO2 nanocomposite modified glassy carbon electrode as a novel and efficient voltammetric sensor for simultaneous determination of hydroquinone and catechol. Sens. Actuators B Chem. 2013, 177, 412–418. [Google Scholar] [CrossRef]
  23. Roshanfekr, H. A simple specific dopamine aptasensor based on partially reduced graphene oxide–Au NPs composite. Prog. Chem. Biochem. Res. 2023, 6, 79–88. [Google Scholar]
  24. Beitollahi, H.; Tajik, S.; Karimi-Maleh, H.; Hosseinzadeh, R. Application of a 1-benzyl-4-ferrocenyl-1H-[1,2,3]-triazole/carbon nanotube modified glassy carbon electrode for voltammetric determination of hydrazine in water samples. Appl. Organomet. Chem. 2013, 27, 444–450. [Google Scholar] [CrossRef]
  25. Ariavand, S.; Ebrahimi, M.; Foladi, E. Design and construction of a novel and an efficient potentiometric sensor for determination of sodium ion in urban water samples. Chem. Methodol. 2022, 6, 886–904. [Google Scholar]
  26. Mohanraj, J.; Durgalakshmi, D.; Rakkesh, R.A.; Balakumar, S.; Rajendran, S.; Karimi-Maleh, H. Facile synthesis of paper based graphene electrodes for point of care devices: A double stranded DNA (dsDNA) biosensor. J. Colloid Interface Sci. 2020, 566, 463–472. [Google Scholar] [CrossRef]
  27. Hasanpour, F.; Taei, M.; Fouladgar, M.; Salehi, M. Au nano dendrites/composition optimized nd-dopped cobalt oxide as an efficient electrocatalyst for ethanol oxidation. J. Appl. Organomet. Chem. 2022, 2, 188–196. [Google Scholar]
  28. Guo, Q.; Zhang, M.; Zhou, G.; Zhu, L.; Feng, Y.; Wang, H.; Hou, H. Highly sensitive simultaneous electrochemical detection of hydroquinone and catechol with three-dimensional N-doping carbon nanotube film electrode. J. Electroanal. Chem. 2016, 760, 15–23. [Google Scholar] [CrossRef]
  29. Li, J.; Liu, C.Y.; Cheng, C. Electrochemical detection of hydroquinone by graphene and Pt-graphene hybrid material synthesized through a microwave-assisted chemical reduction process. Electrochim. Acta 2011, 56, 2712–2716. [Google Scholar] [CrossRef]
  30. Wang, H.; Li, R.; Li, Z. Nanohybrid of Co3O4 and histidine-functionalized graphene quantum dots for electrochemical detection of hydroquinone. Electrochim. Acta 2017, 255, 323–334. [Google Scholar] [CrossRef]
  31. Li, S.J.; Xing, Y.; Wang, G.F. A graphene-based electrochemical sensor for sensitive and selective determination of hydroquinone. Microchim. Acta 2012, 176, 163–168. [Google Scholar] [CrossRef]
  32. Nasr, B.; Abdellatif, G.; Canizares, P.; Sáez, C.; Lobato, J.; Rodrigo, M.A. Electrochemical oxidation of hydroquinone, resorcinol, and catechol on boron-doped diamond anodes. Environ. Sci. Technol. 2005, 39, 7234–7239. [Google Scholar]
  33. Hu, S.; Wang, Y.; Wang, X.; Xu, L.; Xiang, J.; Sun, W. Electrochemical detection of hydroquinone with a gold nanoparticle and graphene modified carbon ionic liquid electrode. Sens. Actuators B Chem. 2012, 168, 27–33. [Google Scholar] [CrossRef]
  34. Bijad, M.; Karimi-Maleh, H.; Farsi, M.; Shahidi, S.A. An electrochemical-amplified-platform based on the nanostructure voltammetric sensor for the determination of carmoisine in the presence of tartrazine in dried fruit and soft drink samples. J. Food Meas. Charact. 2018, 12, 634–640. [Google Scholar] [CrossRef]
  35. Hosseini Fakhrabad, A.; Sanavi Khoshnood, R.; Abedi, M.R.; Ebrahimi, M. Fabrication a composite carbon paste electrodes (CPEs) modified with Multi-Wall Carbon Nano-Tubes (MWCNTs/N, N-Bis (salicyliden)-1,3-propandiamine) for determination of lanthanum (III). Eurasian Chem. Commun. 2021, 3, 627–634. [Google Scholar]
  36. Zhang, Z.; Karimi-Maleh, H. Label-free electrochemical aptasensor based on gold nanoparticles/titanium carbide MXene for lead detection with its reduction peak as index signal. Adv. Compos. Hybrid Mater. 2023, 6, 68. [Google Scholar] [CrossRef]
  37. Peyman, H. Design and fabrication of modified DNA-Gp nano-biocomposite electrode for industrial dye measurement and optical confirmation. Prog. Chem. Biochem. Res. 2022, 5, 391–405. [Google Scholar]
  38. Buledi, J.A.; Mahar, N.; Mallah, A.; Solangi, A.R.; Palabiyik, I.M.; Qambrani, N.; Karimi, F.; Vasseghian, Y.; Karimi-Maleh, H. Electrochemical quantification of mancozeb through tungsten oxide/reduced graphene oxide nanocomposite: A potential method for environmental remediation. Food Chem. Toxicol. 2022, 161, 112843. [Google Scholar] [CrossRef] [PubMed]
  39. Mehdizadeh, Z.; Shahidi, S.A.; Ghorbani-HasanSaraei, A.; Limooei, M.B.; Bijad, M. Monitoring of amaranth in drinking samples using voltammetric amplified electroanalytical sensor. Chem. Methodol. 2022, 6, 246–252. [Google Scholar]
  40. Husain, Z.; Raheman, A.S.; Ansari, K.B.; Pandit, A.B.; Khan, M.S.; Qyyum, M.A.; Lam, S.S. Nano-sized mesoporous biochar derived from biomass pyrolysis as electrochemical energy storage supercapacitor. Mater. Sci. Technol. 2022, 5, 99–109. [Google Scholar] [CrossRef]
  41. Janitabar Darzi, S.; Bastami, H. Au decorated mesoporous TiO2 as a high performance photocatalyst towards crystal violet dye. Adv. J. Chem. A 2022, 5, 22–30. [Google Scholar]
  42. Kavade, R.; Khanapure, R.; Gawali, U.; Patil, A.; Patil, S. Degradation of methyl orange under visible light by ZnO-Polyaniline nanocomposites. J. Appl. Organomet. Chem. 2022, 2, 101–112. [Google Scholar]
  43. Sheikhshoaie, I.; Rezazadeh, A.; Ramezanpour, S. Removal of Pb (II) from aqueous solution by gel combustion of a new nano sized Co3O4/ZnO composite. Asian J. Nanosci. Mater. 2022, 5, 336–345. [Google Scholar]
  44. Dormanesh, R.; Mohseni, S. Green synthesis of zinc oxide nanoparticles and their application in ring opening of epoxides. Prog. Chem. Biochem. Res. 2021, 4, 403–413. [Google Scholar]
  45. Muhi-Alden, Y.Y.; Saleh, K.A. Removing methylene blue dye from industrial wastewater using polyacrylonitrile/iron oxide nanocomposite. Eurasian Chem. Commun. 2021, 3, 755–762. [Google Scholar]
  46. Tallapaneni, V.; Mude, L.; Pamu, D.; Karri, V.V.S.R. Formulation, characterization and in vitro evaluation of dual-drug loaded biomimetic chitosan-collagen hybrid nanocomposite scaffolds. J. Med. Chem. Sci. 2022, 5, 1059–1074. [Google Scholar]
  47. Abadi, M.A.A.; Masrournia, M.; Abedi, M.R. Simultaneous extraction and preconcentration of benzene, toluene, ethylbenzene and xylenes from aqueous solutions using magnetite-graphene oxide composites. Chem. Methodol. 2021, 5, 11–20. [Google Scholar]
  48. Khazaeli, P.; Alaei, M.; Khaksarihadad, M.; Ranjbar, M. Preparation of PLA/chitosan nanoscaffolds containing cod liver oil and experimental diabetic wound healing in male rats study. J. Nanobiotechnol. 2020, 18, 176. [Google Scholar] [CrossRef]
  49. Tajik, S.; Lohrasbi-Nejad, A.; Mohammadzadeh Jahani, P.; Askari, M.B.; Salarizadeh, P.; Beitollahi, H. Co-detection of carmoisine and tartrazine by carbon paste electrode modified with ionic liquid and MoO3/WO3 nanocomposite. J. Food Meas. Charact. 2022, 16, 722–730. [Google Scholar] [CrossRef]
  50. Mohammadi, S.; Beitollahi, H.; Mohadesi, A. Electrochemical behaviour of a modified carbon nanotube paste electrode and its application for simultaneous determination of epinephrine, uric acid and folic acid. Sens. Lett. 2013, 11, 388–394. [Google Scholar] [CrossRef]
  51. Karimi-Maleh, H.; Fakude, C.T.; Mabuba, N.; Peleyeju, G.M.; Arotiba, O.A. The determination of 2-phenylphenol in the presence of 4-chlorophenol using nano-Fe3O4/ionic liquid paste electrode as an electrochemical sensor. J. Colloid Interface Sci. 2019, 554, 603–610. [Google Scholar] [CrossRef] [PubMed]
  52. Beitollahi, H.; Sheikhshoaie, I. Novel nanostructure-based electrochemical sensor for simultaneous determination of dopamine and acetaminophen. Mater. Sci. Eng. C 2012, 32, 375–380. [Google Scholar] [CrossRef]
  53. Roy, S.; Soin, N.; Bajpai, R.; Misra, D.S.; McLaughlin, J.A.; Roy, S.S. Graphene oxide for electrochemical sensing applications. J. Mater. Chem. 2011, 21, 14725–14731. [Google Scholar] [CrossRef]
  54. Ardakani, M.M.; Taleat, Z.; Beitollahi, H.; Salavati-Niasari, M.; Mirjalili, B.B.F.; Taghavinia, N. Electrocatalytic oxidation and nanomolar determination of guanine at the surface of a molybdenum (VI) complex–TiO2 nanoparticle modified carbon paste electrode. J. Electroanal. Chem. 2008, 624, 73–78. [Google Scholar] [CrossRef]
  55. Ratinac, K.R.; Yang, W.; Gooding, J.J.; Thordarson, P.; Braet, F. Graphene and related materials in electrochemical sensing. Electroanalysis 2011, 23, 803–826. [Google Scholar] [CrossRef]
  56. Almarbd, Z.Z.; Abbass, N.M. Synthesis and characterization of TiO2, Ag2O, and graphene oxide nanoparticles with polystyrene as a nonocomposites and some of their applications. Eurasian Chem. Commun. 2022, 4, 1033–1043. [Google Scholar]
  57. Garkani-Nejad, F.; Beitollahi, H.; Alizadeh, R. Sensitive determination of hydroxylamine on ZnO nanorods/graphene oxide nanosheets modified graphite screen printed electrode. Anal. Bioanal. Electrochem. 2017, 9, 134–144. [Google Scholar]
  58. Li, J.; Liu, J.; Tan, G.; Jiang, J.; Peng, S.; Deng, M.; Liu, Y. High-sensitivity paracetamol sensor based on Pd/graphene oxide nanocomposite as an enhanced electrochemical sensing platform. Biosens. Bioelectron. 2014, 54, 468–475. [Google Scholar] [CrossRef]
  59. Dai, Y.; Huang, J.; Zhang, H.; Liu, C.C. Highly sensitive electrochemical analysis of tunnel structured MnO2 nanoparticle-based sensors on the oxidation of nitrite. Sens. Actuators B Chem. 2019, 281, 746–750. [Google Scholar] [CrossRef]
  60. Beitollahi, H.; Tajik, S.; Di Bartolomeo, A. Application of MnO2 nanorod–ionic liquid modified carbon paste electrode for the voltammetric determination of sulfanilamide. Micromachines 2022, 13, 598. [Google Scholar] [CrossRef]
  61. Lu, J.; Zhang, X.; Zhang, X.; Liu, N.; Li, H.; Yu, Z.; Yan, X. Electrochemical sensor for mercuric chloride based on graphene-MnO2 composite as recognition element. Electrochim. Acta 2015, 174, 221–229. [Google Scholar] [CrossRef]
  62. Obodo, R.M.; Onah, E.O.; Nsude, H.E.; Agbogu, A.; Nwanya, A.C.; Ahmad, I.; Ezema, F.I. Performance evaluation of graphene oxide based Co3O4@GO, MnO2@GO and Co3O4/MnO2@GO electrodes for supercapacitors. Electroanalysis 2020, 32, 2786–2794. [Google Scholar] [CrossRef]
  63. Li, L.; Du, Z.; Liu, S.; Hao, Q.; Wang, Y.; Li, Q.; Wang, T. A novel nonenzymatic hydrogen peroxide sensor based on MnO2/graphene oxide nanocomposite. Talanta 2010, 82, 1637–1641. [Google Scholar] [CrossRef]
  64. Kumar, P.; Khan, M.Q.; Khan, R.A.; Ahmad, K.; Kim, H. Hydrothermal synthesis of MnO2/reduced graphene oxide composite for 4-Nitrophenol sensing applications. Inorganics 2022, 10, 219. [Google Scholar] [CrossRef]
  65. Tian, Y.; Deng, P.; Wu, Y.; Li, J.; Liu, J.; Li, G.; He, Q. MnO2 nanowires-decorated reduced graphene oxide modified glassy carbon electrode for sensitive determination of bisphenol A. J. Electrochem. Soc. 2020, 167, 046514. [Google Scholar] [CrossRef]
  66. Lei, J.; Lu, X.; Wang, W.; Bian, X.; Xue, Y.; Wang, C.; Li, L. Fabrication of MnO2/graphene oxide composite nanosheets and their application in hydrazine detection. RSC Adv. 2012, 2, 2541–2544. [Google Scholar] [CrossRef]
Figure 1. FE-SEM image of GO (a) and MnO2 NRs/GO nanocomposite (b).
Figure 1. FE-SEM image of GO (a) and MnO2 NRs/GO nanocomposite (b).
Biomedicines 11 01869 g001
Figure 2. Plot of the oxidation peak current of 200.0 μM HQ as a function of pH solution on MnO2 NRs/GO/GCE in 0.1 M PBS at different pH value (2.0–9.0).
Figure 2. Plot of the oxidation peak current of 200.0 μM HQ as a function of pH solution on MnO2 NRs/GO/GCE in 0.1 M PBS at different pH value (2.0–9.0).
Biomedicines 11 01869 g002
Figure 3. CVs of the (a) MnO2 NRs/GO/GCE in 0.1 M PBS (pH = 7.0) in the absence of HQ), (b–e) bare GCE, MnO2 NRs/GCE, GO/GCE and MnO2 NRs/GO/GCE in 0.1 M PBS (pH = 7.0) with 200.0 μM HQ. The scan rate was equal to 50 mV s−1.
Figure 3. CVs of the (a) MnO2 NRs/GO/GCE in 0.1 M PBS (pH = 7.0) in the absence of HQ), (b–e) bare GCE, MnO2 NRs/GCE, GO/GCE and MnO2 NRs/GO/GCE in 0.1 M PBS (pH = 7.0) with 200.0 μM HQ. The scan rate was equal to 50 mV s−1.
Biomedicines 11 01869 g003
Figure 4. CVs of 120.0 μM HQ on MnO2 NRs/GO/GCE in 0.1 M PBS (pH = 7.0) at the various scanning rates (ν); (Curves a–n: (a) 5 mV/s, (b) 10 mV/s, (c) 20 mV/s, (d) 30 mV/s, (e) 40 mV/s, (f) 50 mV/s, (g) 60 mV/s, (h) 70 mV/s, (i) 80 mV/s, (j) 90 mV/s, (k) 100 mV/s, (l) 200 mV/s, (m) 300 mV/s, and (n) 400 mV/s. Inset: The plot of Ipa and Ipc vs. the square root of the scanning rate (υ1/2).
Figure 4. CVs of 120.0 μM HQ on MnO2 NRs/GO/GCE in 0.1 M PBS (pH = 7.0) at the various scanning rates (ν); (Curves a–n: (a) 5 mV/s, (b) 10 mV/s, (c) 20 mV/s, (d) 30 mV/s, (e) 40 mV/s, (f) 50 mV/s, (g) 60 mV/s, (h) 70 mV/s, (i) 80 mV/s, (j) 90 mV/s, (k) 100 mV/s, (l) 200 mV/s, (m) 300 mV/s, and (n) 400 mV/s. Inset: The plot of Ipa and Ipc vs. the square root of the scanning rate (υ1/2).
Biomedicines 11 01869 g004
Figure 5. Chronoamperograms for the different concentrations of HQ on the MnO2 NRs/GO/GCE in 0.1 M PBS (pH = 7.0) in ranges between 0.1 and 2.0 mM (Curves a–d: (a) 0.1 mM, (b) 0.5 mM, (c) 1.5 mM, and (d) 2.0 mM). Insets: I-plots vs. t−1/2 for chronoamperograms a–d (A) and the slope from the straight lines vs. the concentration of HQ (B).
Figure 5. Chronoamperograms for the different concentrations of HQ on the MnO2 NRs/GO/GCE in 0.1 M PBS (pH = 7.0) in ranges between 0.1 and 2.0 mM (Curves a–d: (a) 0.1 mM, (b) 0.5 mM, (c) 1.5 mM, and (d) 2.0 mM). Insets: I-plots vs. t−1/2 for chronoamperograms a–d (A) and the slope from the straight lines vs. the concentration of HQ (B).
Biomedicines 11 01869 g005
Figure 6. DPVs for diverse concentrations of HQ on MnO2 NRs/GO/GCE in 0.1 M PBS (pH = 7.0) in ranges from 0.5 to 300.0 μM (Curves a–n: (a) 0.5 μM, (b) 5.0 μM, (c) 10.0 μM, (d) 20.0 μM, (e) 30.0 μM, (f) 40.0 μM, (g) 50.0 μM, (h) 60.0 μM, (i) 70.0 μM, (j) 80.0 μM, (k) 90.0 μM, (l) 100.0 μM, (m) 200.0 μM, and (n) 300.0 μM). Inset: the related linear calibration curve of the peak current vs. concentration of HQ.
Figure 6. DPVs for diverse concentrations of HQ on MnO2 NRs/GO/GCE in 0.1 M PBS (pH = 7.0) in ranges from 0.5 to 300.0 μM (Curves a–n: (a) 0.5 μM, (b) 5.0 μM, (c) 10.0 μM, (d) 20.0 μM, (e) 30.0 μM, (f) 40.0 μM, (g) 50.0 μM, (h) 60.0 μM, (i) 70.0 μM, (j) 80.0 μM, (k) 90.0 μM, (l) 100.0 μM, (m) 200.0 μM, and (n) 300.0 μM). Inset: the related linear calibration curve of the peak current vs. concentration of HQ.
Biomedicines 11 01869 g006
Table 1. Comparison the electrochemical responses of 200.0 μM HQ in 0.1 M PBS (pH = 7.0) at the surface of different electrodes.
Table 1. Comparison the electrochemical responses of 200.0 μM HQ in 0.1 M PBS (pH = 7.0) at the surface of different electrodes.
ElectrodeAnodic Peak Current (µA)Anodic Peak Potential (mV)Cathodic Peak Current (µA)Cathodic Peak Potential (mV)
Bare GCE5.4290−4.0−140
MnO2 NRs/GCE8.5240−6.5−130
GO/GCE11.7215−9.9−115
MnO2 NRs/GO/GCE15.0200−13.4−100
Table 2. Comparison the sensing performances toward the detection of HQ between the existing electrochemical sensors and the proposed MnO2 NRs/GO/GCE sensor.
Table 2. Comparison the sensing performances toward the detection of HQ between the existing electrochemical sensors and the proposed MnO2 NRs/GO/GCE sensor.
Electrochemical SensorAnalytical MethodsDynamic Linear RangeLimit of DetectionRef.
Zinc @ zinc oxide core-shell/glassy carbon electrodeCyclic voltammetry10.0 to 90.0 µM0.10443 µM[14]
Glassy carbon electrode modified with multiwall carbon nanotubesDifferential pulse voltammetry1.0 × 10−6 M to 1.0 × 10−4 M7.5 × 10−7 M[15]
Reduced graphene oxide cross-linked L-cysteine/glassy carbon electrodeDifferential pulse voltammetry2.0 to 160.0 µM1.5 µM[16]
Au@Pd nanocomposites/glassy carbon electrodeDifferential pulse voltammetry4.0 to 5000.0 µM0.63 µM[17]
Poly-amidosulfonic acid and multi-wall carbon nanotubes composite electropolymerization on glassy carbon electrodeDifferential pulse voltammetry6.0 × 10−6 to 4.0 × 10−4 M1.0 × 10−6 M[18]
Nanodiamond/glassy carbon electrodeDifferential pulse voltammetry1.0 to 78.0 μM0.19 μM[19]
CuS nanocrystals/chitosan/glassy carbon electrodeCyclic voltammetry4.5 µM to 4.5 mM1.5 µM[20]
Electrodeposition of reduced graphene oxide on glassy carbon electrodeDifferential pulse voltammetry6.0 to 200.0 μM0.2 μM[21]
GO–mesoporous MnO2 nanocomposite/glassy carbon electrodeDifferential pulse voltammetry0.01 to 0.7 µM7.0 nM[22]
MnO2 NRs/GO/GCEDifferential pulse voltammetry0.5 to 300.0 μM0.012 µMThis Work
Table 3. HQ detection in the real samples with the MnO2 NRs/GO/GCE (concentration in µM (n = 5)).
Table 3. HQ detection in the real samples with the MnO2 NRs/GO/GCE (concentration in µM (n = 5)).
SampleSpikedFoundRecovery (%)R.S.D. (%)
River water0---
5.04.998.03.5
7.07.1101.41.9
Tap
Water
0---
5.55.7103.62.8
7.57.397.32.1
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

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. https://doi.org/10.3390/biomedicines11071869

AMA Style

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(7):1869. https://doi.org/10.3390/biomedicines11071869

Chicago/Turabian Style

Karami-Kolmoti, Parisa, Hadi Beitollahi, and Sina Modiri. 2023. "Electrochemical Sensor for Simple and Sensitive Determination of Hydroquinone in Water Samples Using Modified Glassy Carbon Electrode" Biomedicines 11, no. 7: 1869. https://doi.org/10.3390/biomedicines11071869

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