Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Graphene oxide anchored ferric vanadate acid for the electrochemical detection of bisphenol A in tap water and urine

Huan Taoab, Feiling Xieb, Zhi Li*bd, Liangzhuo Qu*b, Qingbing Zha*bc and Mingshan Zhubd
aDepartment of Clinical Laboratory Medical, The First Affiliated Hospital of Jinan University, Guangzhou 510630, China
bDepartment of Medical Laboratory, The Fifth Affiliated Hospital of Jinan University (Heyuan Shenhe People's Hospital), Heyuan, 517000, China. E-mail: zhili@jnu.edu.cn; quliangzhuo@163.com; zhaqingbb@sina.com
cCenter of Reproductive Medicine, The First Affiliated Hospital of Jinan University, Guangzhou 510630, China
dCollege of Environment and Climate, Jinan University, Guangzhou 511443, China

Received 4th June 2025 , Accepted 12th July 2025

First published on 22nd July 2025


Abstract

The gradual emissions of bisphenol A (BPA) pose significant health risks by interfering with hormonal regulation, which urgently require to develop an efficient and rapid detection method. Here, an electrochemical sensor using graphene oxide anchored ferric vanadate acid (GO/FeVO4) nanocomposite was developed for BPA detection. The electrochemical signals of GO grafted on FeVO4 was significantly enhanced by 7.4-fold relative to that of pure FeVO4, attributed to transition metal valence effects and adsorption by H-bond. The optimized system achieved a detection limit of 1.18 μM with a linear range of 0.01–40 μM. Additionally, it displayed outstanding stability, reproducibility and selectivity. These findings demonstrate strong potential for environmental monitoring and clinical diagnostics, particularly for rapid on-site BPA detection in tap water and urine. The sensor of GO/FeVO4 with high conductivity and catalytic activity offers a sensitive, cost-effective solution for BPA detection in environmental and medical clinics.


1. Introduction

As is known, bisphenol A (BPA), a commonly used plastic monomer, is extensively utilized in food packaging and container manufacturing to effectively protect food from external contamination.1,2 However, temperature-dependent BPA migration from plastic products during production poses demonstrated risks to human health, including endocrine disruption and developmental toxicity. According to recent investigations, prolonged exposure to BPA has been epidemiologically associated with various adverse health effects, particularly reproductive system disorders, elevated cancer risks, and endocrine disruption. Hence, developing rapid and sensitive BPA detection methods is critical for environmental protection and public health preservation.3–5

Currently, various analytical methods such as chemiluminescence, mass spectrometry and liquid chromatography, and gas chromatography-mass spectrometry, have been widely employed for the detection of BPA.6 Despite the advantages of specificity and accuracy, high-end chromatographic methods face significant limitation including exorbitant prices, complexity of operation, and poor suitability for real-time monitoring.7–11 Recently, electrochemical sensing technology has gained attention as a potentially effective strategy because of simplified sample preparation, operational ease, high sensitivity and rapid response times.12–16

Ferric vanadate acid (FeVO4), a representative two-dimensional material, exhibits an optimal combination of high sensitivity, excellent selectivity, fast response kinetics, remarkable stability, and unique photoelectric properties. These characteristics enable selective molecular recognition capabilities for precise target analyte detection. Recent research results have demonstrated that FeVO4 nanorods exhibit excellent electrochemical response to BPA, indicating its potential as an efficient sensor material for detecting BPA.17,18 However, the electron transfer capability of FeVO4 is still relatively low, which limits its high sensitivity in electrochemical detection of BPA. Graphene oxide (GO), as an emerging two-dimensional material, possesses lots of advantages including large surface area, high sensitivity, excellent conductivity, drug delivery, and sensing, making it an ideal substance for chemical and biosensors.19–21 Therefore, grafting GO onto the surface of FeVO4 as an electron bridge accelerates the electron transfer of target molecules to the FeVO4 surface, effectively enhancing the detection sensitivity of GO/FeVO4.

Inspirated by these, herein, taking the specific binding ability of FeVO4 to target molecules, which can change the binding state between itself and target molecules, and the transfer capability of GO, we proposed a GO/FeVO4 nanocomposite-based electrochemical sensor for detecting BPA (Scheme 1). The carboxyl and hydroxyl groups on GO sheets could provide binding sites for BPA through H-bonding, and FeVO4 acts as an electron bridge with valence metal effect to accelerate the oxidation of BPA and the rapid transfer of electrons, thereby enhancing the electrochemical sensing signal of BPA. This sensor demonstrates good reproducibility, superior selectivity, stability. These characteristics, combined with its successful application in complex matrices including human urine and tap water, confirm its strong potential for practical environmental and biomedical monitoring.


image file: d5ra03944d-s1.tif
Scheme 1 The GO/FeVO4 electrochemical sensor for BPA detection.

2. Materials and methods

The instruments, other experiments and characterizations are discussed in ESI.

2.1. Materials

BPA was sourced from Sigma Chemical Co. Ltd, while GO was procured from Nanjing Xianfen. Sodium dihydrogen phosphate, disodium hydrogen phosphate, potassium ferrocyanide, potassium ferricyanide, ethylene glycol, hydrochloric acid, ethanol, sodium hydroxide, potassium chloride and the supplier of ferric vanadate acid was Sinopharm Chemicals Reagent Co., Ltd in China. Every reagent used were analytical grade, and they were all used immediately without any additional purification.

2.2. Preparation of GO, FeVO4 and GO/FeVO4

2 mg GO was dissolved in a mixture solution of ethanol and water (Vwater/Vethanol = 1/1) for a 30 min ultrasonic dispersion, and was subsequently diluted into 0.02, 0.04, 0.08, 0.12 and 0.2 mg per mL GO solutions. To synthesize GO/FeVO4, 2 mg FeVO4 was added respectively to the different concentrations of GO solutions and sufficiently dissolve through a 30-minute ultrasonic dispersion to obtain 1%, 2%, 4%, 6% and 10% GO/FeVO4 solutions. For comparison, solutions containing GO or FeVO4 alone was prepared with the same method. Then, 5 μL Nafion was added to the above mixed solution.

2.3. Fabrication of GO/FeVO4 sensor

Before preparation, 0.05 mm alumina powder was applied to polish the surface of glassy carbon electrode (GCE) (3 mm in diameter), and then the GCE surface was thoroughly washed with ultrapure water and dried at room temperature. Next, the pretreated GCE was covered with 10 μL of GO/FeVO4 suspension and dried (Scheme 1). For comparison, the GO or FeVO4 sensor were prepared via the same procedure.

2.4. Apparatus

Using a standard three-electrode arrangement, a CHI 760 electrochemical workstation was used for all electrochemical measurements. The standard setup included a platinum counter electrode, and a saturated calomel electrode (SCE) as the reference electrode, and the working electrode is a GCE modified by GO/FeVO4. The differential pulse voltammetry (DPV) approach was employed to assess BPA, and the parameters used were as follows: 0.05 s pulse width, 50 mV amplitude, 0–1.2 V potential range, and a 0.004 V additional potential. Cyclic voltammetry (CV) was applied to evaluate the electrochemical characterization of different electrodes under the conditions of 0.001 V interval, −0.4 to 0.8 V potential range, and 100 mV s−1 scan rate. Electrochemical impedance spectroscopy (EIS) was carried out with a 5 mV amplitude and the range of frequencies of 1.0–1.0 × 105 Hz in 0.1 M KCl solution, which includes 2.5 mM K3[Fe(CN)6]/K4[Fe(CN)6].

2.5. Pretreatment of real sample

The urine samples were obtained from three healthy volunteers at Department of Clinical Laboratory Medical, The First Affiliated Hospital of Jinan University (Guangzhou, China). Each research participant signed a written informed consent form in accordance with the Declaration of Helsinki. According to the “Ethical Review Method for Life Science and Medical Research Involving Humans” in China, this study does not require further ethical approval from the ethics committee, as its purpose does not involve the exploration of human diseases or any potentially heritable genetic manipulation. Tap water was collected at the laboratory of The First Affiliated Hospital of Jinan University. These real sample was stored at 4 °C and then they were centrifuged at 8000 rpm to obtain the supernatant. The supernatant was further diluted 10-fold with 0.1 M phosphate buffer (pH = 7.0) to prepare a standard solution for subsequent electrochemical testing. DPV technique was adopted for the BPA detection in real sample, and the standard spiking method was applied to evaluate the recovery rates of BPA at concentrations of 2, 4, and 6 μM in real samples.

3. Results and discussion

3.1. Physical characterizations

The morphologies of FeVO4, GO and GO/FeVO4 were observed by transmission electron microscopy (TEM), which was presented as the one-dimensional belt-like shape, respectively (Fig. 1a–c).22 The XRD pattern was used to characterize the chemical makeup of the produced GO, FeVO4, and GO/FeVO4 in Fig. 1d. The prepared GO sheet exhibited a prominent peak at 10.2°, which is in excellent alignment with the values stated in the literature.23,24 The FeVO4 sample was synthesized through a straightforward hydrothermal method (Fig. 1d), displaying some weak diffraction peaks, indicating its low crystallinity. The XRD analysis of the calcined sample was precisely matched to triclinic FeVO4 (PDF #71-1592), consistent with previous reports.25 The result demonstrated GO incorporation did not change the FeVO4 crystal structure.
image file: d5ra03944d-f1.tif
Fig. 1 TEM images of FeVO4 (a), GO (b) and GO/FeVO4 (c); XRD patterns (d) of FeVO4; Fe 2p (e) and V 2p (f) spectra of FeVO4.

Additionally, the surface valence states and elemental composition of FeVO4 nanobelts were investigated using X-ray photoelectron spectroscopy (XPS). Fig. 1e shows the Fe 2p spectrum, and there were six peaks found. Two peaks at 711.51 eV and 724.10 eV show Fe 2p3/2 and Fe 2p1/2, respectively, indicating that Fe2+ is one of the Fe species found on the nanobelts. As seen in (Fig. 1f), the presence of Fe3+ species is confirmed by two further peaks at 714.27 eV and 726.51 eV, and two satellite peaks at 719.69 eV and 733.41 eV, V 2p3/2 and V 2p1/2 are responsible for the peaks at 517.12 eV and 530.15 eV, respectively. According to this discovery, vanadium can be found on the surface of the nanobelt in two valence states: pentavalent (V5+) and tetravalent (V4+). The presence of low-valence Fe/V species on FeVO4 can be attributed to the reaction between FeCl3 and NH4VO3 under high temperature and pressure, which releases ammonia. Acting as a reducing agent, ammonia converts some of the low-valence Fe2+/V4+ species from the high-valence Fe3+/V5+species.26

3.2. Electrochemical characterization

CV analysis was used to compare and analyze the electrochemical performance and electron transport properties of the four suggested sensors (Fig. 2a). The results showed that the GCE electrode displayed the weakest redox peak signal, which can be enhanced by GO and FeVO4, respectively. Notably, the GO/FeVO4 sensor, where GO and FeVO4 were simultaneously added on the GCE electrode, showed the highest redox peak, indicating that the modification by GO can effectively improve the electronic transmission ability of FeVO4. The electron transport on the electrode surfaces of a few sensors was investigated with EIS.27 The electrical impedance of four sensors were represented by the Nyquist diagram (Fig. 2b). The radius is typically used to illustrate the charge transfer resistance, a smaller radius denotes a lower resistance.28,29 The GO sensor has the smallest electrical impedance value, followed by GO/FeVO4, FeVO4 and GCE sensors. The EIS data was further fitted to acquire the equivalent circuit diagram (Fig. 2c). In the Fig. 2d, the Bode plots were also examined to evaluate the capacity of the sample for charge transfer resistance. As the impedance value increases, log|Z| increases in the Bode-magnitude plot.30–34 As a result, the GO/FeVO4 electrode exhibited a faster electron transport capability compared to the GO electrode, with similar trends observed in the Bode-phase plot. Based on these findings, the GO/FeVO4 electrode shows great promise as a superior electrochemical sensing material.
image file: d5ra03944d-f2.tif
Fig. 2 CVs (a), Nyquist plots (b), equivalent circuit (c) and Bode plots (d) of GCE, FeVO4, GO and GO/FeVO4 in a 2.5 mM K3[Fe(CN)6]/K4[Fe(CN)6] solution containing 0.1 M KCl. Equivalent circuit was used for simulating EIS.

3.3. Electrochemical detection of BPA

The electrochemical behavior of 20 μM BPA in 0.1 M phosphate buffer (pH = 7.0) for four various electrodes was reflected by DPV method (Fig. 3a). BPA showed a comparatively consistent oxidation peak potential for the GO/FeVO4 sensor at 0.512 V. Compared with GCE, FeVO4 and GO electrodes, the peak current signal of GO/FeVO4 for BPA is improved 9.0, 7.4 and 2.8 times (Fig. 3b), respectively. This can be ascribed to the remarkable electrocatalytic activity of GO, which accelerates the electron transport of FeVO4 for BPA detection. Additionally, the porous structure of FeVO4 affords the GO nanosheets a large number of attachment sites.35 These results indicate that the GO/FeVO4 sensor has outstanding BPA detection capabilities.
image file: d5ra03944d-f3.tif
Fig. 3 The DPVs (a) and matching histograms (b) for different sensors in a 0.1 M phosphate buffer (pH = 7.0) solution.

3.4. GO/FeVO4 electrode ideal conditions

For the determination of optimal detection conditions of GO/FeVO4 electrode, a series of parameters including content percentage, time, accumulation potential and pH value were investigated with DPV method. Firstly, the content of GO on GO/FeVO4 electrode is one of the important factors for BPA detection. From 1 to 6%, as the GO concentration increased, the peak current gradually rose; however, it sharply decreased when the GO content reached 10%. This drop is attributed to the excessive GO, which obstructed the active sites of FeVO4 (Fig. 4a). Therefore, 6% GO was chosen as the optimal content for BPA detection. In addition, the phosphate buffer at a certain pH level can have a significant impact on BPA detection (Fig. 4b). The peak current of BPA was measured at different pH level range from 5.0–7.5, and the results showed that BPA has the highest current value when pH is 7.0. As a result, a phosphate buffer solution with a pH of 7.0 was selected as a prerequisite for the ensuing detection procedure. Additionally, the accumulation potential and the accumulation time are also critical factors in the electrochemical detection of BPA (Fig. 4c and d). The peak current of BPA was changed at different accumulation potential (−0.2–0.2 V) and accumulation time (100–300 s). The results of the study showed that BPA exhibited the maximum peak current at 250 s (Fig. 4d) and 0 V (Fig. 4c) accumulation potentials.
image file: d5ra03944d-f4.tif
Fig. 4 The impact of various weight ratios of GO/FeVO4 (a), pH value (b), accumulation potential (c), and accumulation time (d) on the 20 μM BPA peak current in 0.1 M phosphate buffer (pH = 7.0) was studied using the GO/FeVO4 sensor.

3.5. Effect of scan rate

In order to uncover the oxidation mechanism at GO/FeVO4 sensor for BPA detection, the impact of scan rate were explored by CV technique. With the increase of scan rate, the value of the BPA oxidation peak current gradually increases (Fig. 5a), and a linear correlation between the square root of the scanning rate, as well as the peak current was found (Fig. 5b), which can be represented by the following equation: Ipa (μA) = (0.6997 ± 0.10)v1/2 (mV1/2 s−1/2) − (2.7462 ± 0.8367), with an R2 value of 0.9056. The outcome shows diffusion regulates the BPA electrochemical response at the GO/FeVO4 sensor surface. Furthermore, as the scan rate raised, the oxidation peak potentials (Epa) moved favorably. The relationship between the Napierian logarithm of Epa and v (ln[thin space (1/6-em)]v) is linear, as shown in Fig. 5c, and can be described by the equation: Epa (V) = (0.0224 ± 0.0017)ln[thin space (1/6-em)]v (mV s−1) + (0.4662 ± 0.0069), with an R2 value of 0.9729.
image file: d5ra03944d-f5.tif
Fig. 5 CVs (a) of 20 μM BPA detected by the GO/FeVO4 electrode in 0.1 M phosphate buffer (pH = 7.0) at different scan rate from 20 to 120 mV s−1. The correlation between peak currents (Ipa) and v1/2 (mV1/2 s−1/2) (b). The correlation of peak potential and Napierian logarithm of scan rate (ln[thin space (1/6-em)]v) (c).

The following formula defines E for diffusion-controlled and irreversible electrode operations:36

E = Eθ + (RT/anF)ln(RTkθ/anF) + (RT/anF)ln[thin space (1/6-em)]v
In this formula, the variables Eθ, R, T, a, n, F, kθ, and v denote the formal redox potential, the gas constant, the absolute temperature, transfer coefficient, the number of electrons involved in the reaction, the Faraday constant, the standard rate constant of the reaction, and the scan rate, respectively. Consequently, the slope in Fig. 5c corresponds to RT/anF. The value of n was determined to be 2.29 (with F = 96[thin space (1/6-em)]480, R = 8.314, and T = 298), assuming that for an irreversible process, a = 0.5. These findings demonstrate that the electrochemical detection of BPA at the GO/FeVO4 sensor surface involves the conversion of two electrons to two protons.37 Scheme 2 suggests a potential response mechanism based on the aforementioned findings. Firstly, the carboxyl and hydroxyl groups on GO sheets may provide binding sites for BPA through hydrogen bonding,38 GO adsorbs onto the surface of BPA, and an oxidation–reduction reaction occurs under electrochemical influence, generating electrons. Then these electrons are rapidly transmitted to the surface of GCE through the highly conductive GO/FeVO4 hybrid network, where FeVO4 acts as an electron bridge with mixed-valence states (Fe2+/V4+ and Fe3+/V5+).39 This efficient electron transfer results in the generation of a well-defined electrochemical signal in the electrochemical workstation, enabling the highly sensitive and selective quantitative and qualitative detection of BPA.


image file: d5ra03944d-s2.tif
Scheme 2 BPA detection reaction mechanism at the GO/FeVO4 sensor.

3.6. Detection of BPA at GO/FeVO4 sensor

The DPV approach was used to measure the various BPA concentrations found by the GO/FeVO4 electrode under optimal experimental circumstances (Fig. 6a and b). From 0.01–40 μM, the oxidation peak currents elevated with the increase of BPA concentration. Furthermore, the following equation illustrates a linear relationship between the corresponding peak current and the BPA concentration: Ipa (μA) = (0.0102 ± 0.00000059)C (μM) + (0.0262 ± 0.000058), R2 = 0.992. The detection limit (LOD = 3sb/S) of BPA is calculated to be 1.18 μM (Table S1), where S is the slope with the standard curve or the sensitivity for the electrochemical sensor platform, and sb is the standard deviation of the blank.40 In comparison to previous studies, the current GO/FeVO4 sensor display the lowest LOD for the electrochemical identification of BPA (Table S2).
image file: d5ra03944d-f6.tif
Fig. 6 DPVs for various BPA concentrations in 0.1 M phosphate buffer (pH = 7.0) at the GO/FeVO4 sensor (a). Linear relationships between the oxidation peak currents and the concentrations of BPA (b).

3.7. Reproducibility and selectivity studies

The stability of the GO/FeVO4 electrode was evaluated through ten consecutive scans by DPV technique. After 10 consecutive tests, the electrochemical signal of BPA still kept 71% of the first cycle, which suggests GO/FeVO4 sensor with excellent stability (Fig. 7a). To determine the reproducibility of the GO/FeVO4 sensor, 5 electrodes of GO/FeVO4 were independently used for BPA detection with a standard deviation (S.D.) of 6.2% (Fig. 7b and Table S3), demonstrating that the GO/FeVO4 sensor has good reproducibility performance for BPA detection. In addition, in Fig. 7c, the BPA oxidation peak current value was recorded after adding various interferers, including NaCl, Glu, KCl, KBr, dopamine (DA), acetaminophen (APAP) and tetrabromobisphenol A (TBBPA) in 0.1 M phosphate buffer (pH = 7.0). The S.D. of BPA in the presence of interfering species were 3.0%, 3.2%, 3.3%, 3.6%, 7.6%, 8.6%, and 2.9%, respectively (Table S4). These results highlight the excellent interference resistance of the GO/FeVO4 sensor, with all S.D. values remaining below 9.0%. The selectivity of this sensor was measured by DPV in 0.1 M phosphate buffer (pH = 7.0), which includes 40 μM DA solution, 30 μM BPA and 50 μM APAP (Fig. 7d).
image file: d5ra03944d-f7.tif
Fig. 7 The stability (a), reproducibility (b), anti-interference (c) and selectivity (d) of the GO/FeVO4 sensor in 0.1 M phosphate buffer (pH = 7.0).

3.8. Detection of real sample

To detect practicability of the designed GO/FeVO4 sensor, actual samples from human urine and tap water were collected to detect the level of BPA. Initially, a linear relationship between peak current and BPA concentrations in 0.1 M phosphate buffer (pH = 7.0) was established (Fig. S1), which can be represented as: Ipa (μA) = (0.01657 ± 0.00205)CBPA (μM) − (0.00523 ± 0.0082), R2 = 0.9553. The standard addition method was used to analyze BPA by DPV method in urine and tap water. Furthermore, the recovery rates of BPA were calculated using the equation mentioned above. According to Table 1, the recovery rates range from approximately 90% to 110%, suggesting that this sensor may be used to accurately assess BPA levels in real samples (Table 1).
Table 1 BPA detection in actual samples
Sample Added (μM) Found (μM) Recovery (%)
Urine 2 2.30 110.1%
4 4.19 104.7%
6 5.92 98.8%
Tap water 2 2.06 103%
4 4 100.2%
6 6.1 101.9%


4. Conclusions

A new ultra-sensitive electrochemical sensor with GO anchored FeVO4 for detecting BPA has been developed. The carboxyl and hydroxyl groups on GO sheets may provide binding sites for BPA through H-bonding, and FeVO4 acts as an electron bridge with valence metal effect to accelerate the oxidation of BPA and the rapid transfer of electrons, thereby enhancing the electrochemical sensing signal of BPA. The GO/FeVO4 sensor displays a wide linear range and a low detection limit, along with outstanding reproducibility, anti-interference ability, selectivity, and stability. In addition, it was successfully used for the detection of real samples in tap water and urine. These findings suggest that the GO/FeVO4 sensor is a potential electrochemical platform for convenient, simple, and highly effective electrochemical detection for tracking endocrine disruptors, which can be extended to the analysis of environmental contaminants.

Data availability

Data will be made available on request.

Author contributions

H. T.: methodology, investigation, formal analysis, writing—original draft. F. X.: formal analysis, data curation, writing—original draft. Z. L.: conceptualization, writing—review and editing. L. Q.: writing—review and editing, supervision. Q. Z.: conceptualization, writing—review and editing, supervision, funding acquisition. M. Z.: supervision. All authors have read and agreed to the published version of the manuscript.

Conflicts of interest

The authors declare no conflicts of interest.

Acknowledgements

This work supported by the Science and Technology Project of Heyuan, China Social Development and Rural Science (No. 2023008) and the Project of Guangdong Administration of Traditional Chinese Medicine (No. 20232177).

References

  1. G. A. Idowu, T. L. David and A. M. Idowu, Mar. Pollut. Bull., 2022, 176, 113444 CrossRef CAS PubMed.
  2. M. K. Fuzak and A. Z. Pollack, Semin. Reprod. Med., 2024, 42, 274–287 CrossRef CAS PubMed.
  3. A. E. Peters, E. A. Ford, S. D. Roman, E. G. Bromfield, B. Nixon, K. G. Pringle and J. M. Sutherland, Hum. Reprod. Update, 2024, 30, 653–691 CrossRef CAS PubMed.
  4. S. Zhou, X. Wang, Y. Huang, Y. Liu, Y. Zheng, P. Chu, L. Zhu and X. Xu, Mar. Pollut. Bull., 2024, 208, 117058 CrossRef CAS PubMed.
  5. J. Liu, L. Zhang, G. Lu, R. Jiang, Z. Yan and Y. Li, Ecotoxicol. Environ. Saf., 2021, 208, 111481 CrossRef CAS PubMed.
  6. W. Sun, Y. Qi, L. Wang, Y. Tan, X. Zhang, J. Wang and Y. Li, Spectrochim. Acta, Part A, 2025, 327, 125318 CrossRef CAS PubMed.
  7. M. Zhang, Y. Sun, Q. Yang and Y. Chen, Surf. Interfaces, 2025, 58, 105851 CrossRef CAS.
  8. Ş. Yılmaz, Surf. Interfaces, 2022, 32, 102171 CrossRef.
  9. J.-Y. Hu, Z. Li, C.-Y. Zhai, J.-F. Wang, L.-X. Zeng and M.-S. Zhu, Rare Met., 2021, 40, 1727–1737 CrossRef CAS.
  10. N. Akbari, B. Basaran, M. Ghazi-Khansari, B. Akbari-Adergani, A. Bakhtiyari, P. Shavali-gilani and P. Sadighara, Microchem. J., 2024, 207, 111724 CrossRef CAS.
  11. K. Alagumalai, R. Shanmugam, S.-M. Chen, T.-W. Chen, A. M. Al-Mohaimeed, W. A. Al-onazi and M. S. Elshikh, Surf. Interfaces, 2021, 23, 101019 CrossRef CAS.
  12. M. L. Báez, A. García, I. Martínez, C. González, M. Gómez and B. Rodríguez, Int. J. Electrochem. Sci., 2024, 19, 100538 CrossRef.
  13. Z. Zhou, J. Han, J. Shao, N. Li, T. Ren, D. Chen and J. Lu, Surf. Interfaces, 2025, 60, 106051 CrossRef CAS.
  14. S. Mo, Z. Wang, D. Ding, Z. Yan, Y. Dai, J. Zhang, H. Liu, T. Liang, J. Tong, Z. Li and X. Wang, Chin. Chem. Lett., 2025, 36, 110190 CrossRef CAS.
  15. Y. Wang, R. Zeng, S. Tian, S. Chen, Z. Bi, D. Tang and D. Knopp, Anal. Chem., 2024, 96, 13663–13671 CrossRef CAS PubMed.
  16. R. Zeng, M. Qiu, Q. Wan, Z. Huang, X. Liu, D. Tang and D. Knopp, Anal. Chem., 2022, 94, 15155–15161 CrossRef CAS PubMed.
  17. M. Javad Javid-Naderi, N. Valizadeh, B. Banimohamad-Shotorbani, M. Shahgolzari, F. Shayegh, R. Maleki-baladi, S. Sargazi and S. Fathi-karkan, Inorg. Chem. Commun., 2023, 157, 111423 CrossRef CAS.
  18. W. Yang, W. Yang, Y. Huang, C. Xu, L. Dong and X. Peng, Chin. Chem. Lett., 2022, 33, 4628–4634 CrossRef CAS.
  19. A. U. Alam and M. J. Deen, Anal. Chem., 2020, 92, 5532–5539 CrossRef CAS PubMed.
  20. G. Cai, Z. Yu, R. Ren and D. Tang, ACS Sens., 2018, 3, 632–639 CrossRef CAS PubMed.
  21. R. Zeng, Z. Luo, L. Zhang and D. Tang, Anal. Chem., 2018, 90, 12299–12306 CrossRef CAS PubMed.
  22. J. Hu, C. Zhai, L. Zeng, Y. Du and M. Zhu, Catal. Sci. Technol., 2018, 8, 3562–3571 RSC.
  23. D. Ghosh, S. Mehra, S. Sahay, P. K. Singh and S. K. Maji, Int. J. Biol. Macromol., 2017, 100, 37–54 CrossRef CAS PubMed.
  24. P. Viprya, D. Kumar and S. Kowshik, Proc. Eng., 2023, 59, 84 Search PubMed.
  25. C. Wang, C. Jin, T. Wang, Z. Liang, F. Xie and Y. Chang, Mater. Lett., 2025, 378, 137637 CrossRef CAS.
  26. B. Patella, M. Buscetta, S. Di Vincenzo, M. Ferraro, G. Aiello, C. Sunseri, E. Pace, R. Inguanta and C. Cipollina, Sens. Actuators, B, 2021, 327, 128901 CrossRef CAS.
  27. P. Devi, C. Sharma, P. Kumar, M. Kumar, B. K. Bansod, M. K. Nayak and M. L. Singla, J. Hazard. Mater., 2017, 322, 85–94 CrossRef CAS PubMed.
  28. J. Chen, P. Tong, L. Huang, Z. Yu and D. Tang, Electrochim. Acta, 2019, 319, 375–381 CrossRef CAS.
  29. Z. Qiu, D. Tang, J. Shu, G. Chen and D. Tang, Biosens. Bioelectron., 2016, 75, 108–115 CrossRef CAS PubMed.
  30. S. Ahmad, M. H. Ayoub, A. M. Khan, A. Waseem, M. Yasir, M. S. Khan, T. M. Bajwa and A. J. Shaikh, Colloids Surf., A, 2022, 647, 129057 CrossRef CAS.
  31. L. Meng, R. Guo, X. Sun, F. Li, J. Peng, C. Chen, T. Li and J. Deng, Ceram. Int., 2020, 46, 21360–21366 CrossRef CAS.
  32. M. A. Awad, A. A. Hendi, S. Natarajan, K. M. O. Ortashi, S. S. A. Alsaif, R. A. Alnamlah, A. Rasheed and H. Althobaiti, J. King Saud Univ., Sci., 2023, 35, 102857 CrossRef.
  33. Y. Ou, R. Zhu, J. Peng, J. Gao, F. Zhang, M. Hojamberdiev and G. Zhu, Appl. Nanosci., 2021, 13, 465–474 CrossRef.
  34. P. Kamedulski, M. Skorupska, P. Binkowski, W. Arendarska, A. Ilnicka and J. P. Lukaszewicz, Sci. Rep., 2021, 11, 22054 CrossRef CAS PubMed.
  35. W. Wang, Y. Zhang, X. Huang and Y. Bi, J. Mater. Chem. A, 2019, 7, 10949–10953 RSC.
  36. N. Boulanger, A. S. Kuzenkova, A. Iakunkov, A. Nordenström, A. Y. Romanchuk, A. L. Trigub, P. V. Zasimov, M. Prodana, M. Enachescu, S. Bauters, L. Amidani, K. O. Kvashnina, S. N. Kalmykov and A. V. Talyzin, Adv. Mater. Interfaces, 2022, 9, 2200510 CrossRef CAS.
  37. Y. Tian, Z. Yu, L. Cao, X. L. Zhang, C. Sun and D.-W. Wang, J. Energy Chem., 2021, 55, 323–344 CrossRef CAS.
  38. T. S. Silva, M. F. R. Bihain, A. K. dos Santos Pereira and D. H. Pereira, J. Nanopart. Res., 2025, 27, 1–14 CrossRef.
  39. J. Zhang, W. Zhao, Z. Li, G. Lu and M. Zhu, Chem. Eng. J., 2021, 403, 126384 CrossRef CAS.
  40. Y. Li, F. Zhang, Z. Li, L. Sun, Z. Wang, P. Li, Y. Sun, J. Ren, Y. Wang and M. Cribb, Atmos. Res., 2017, 188, 80–89 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ra03944d

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.