Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
Next Article in Journal
Early Identification of Gait Asymmetry Using a Dual-Channel Hybrid Deep Learning Model Based on a Wearable Sensor
Previous Article in Journal
Strength Properties and Damage Evolution Mechanism of Single-Flawed Brazilian Discs: An Experimental Study and Particle Flow Simulation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ultrasensitive Determination of L-Cysteine with g-C3N4@CdS-Based Photoelectrochemical Platform

Marine Engineering College, Dalian Maritime University, Linghai Road 1, Dalian 116026, China
*
Author to whom correspondence should be addressed.
Symmetry 2023, 15(4), 896; https://doi.org/10.3390/sym15040896
Submission received: 23 March 2023 / Revised: 1 April 2023 / Accepted: 7 April 2023 / Published: 11 April 2023
(This article belongs to the Section Chemistry: Symmetry/Asymmetry)

Abstract

:
L-cysteine, a component of the symmetric L-cystine, is essential in numerous biological activities. Thus, detecting cysteine rapidly, selectively, and sensitively is of tremendous interest. Herein, g-C3N4@CdS composites were employed as sensing elements in a photoelectrochemical platform for L-cysteine sensing. In this system, g-C3N4@CdS composites provided much better optoelectronic function than bare CdS materials owing to their high photon-to-current conversion efficiency and excellent anti-photocorrosion properties. The innovative photoelectrochemical sensor has a wide determination range of 5 to 190 µM, a very low detection limit of 1.56 µM, a fast response time, and good long-term stability (ca. 1 month). Without applying any separation procedures, a low concentration of CySH was successfully detected in human urine samples, which is compatible with the results of chemiluminescence.

1. Introduction

L-cysteine (CySH) is an important amino acid owing to its crucial role in biological systems [1,2]. It is known to be an active site in the catalytic function of certain enzymes known as cysteine proteases, as well as in many other peptides and proteins [3,4]. Furthermore, the couple L-cystine/CySH is commonly used as a model for the role of the disulfide bond and thiol group in proteins in a variety of biological media [5]. Given the wide range of physiological and pathophysiological effects [6,7], several strategies for quantifying this critical amino acid have been developed, including cyclic voltammetry [8,9,10,11,12], chemiluminescence [13,14], photoelectrochemistry [15,16,17], fluorimetry [18,19], capillary zone electrophoresis [20], and liquid chromatography [21]. Among these, photoelectrochemical analysis has received a great deal of attention for CySH sensing because of its inherent advantages like simplicity, easy miniaturization, high sensitivity, and low cost.
In the case of the photoelectrochemical detection process, light is used to excite photoactive materials on the electrode, and the photocurrent is used as the detection signal. Thus, the development of photoactive materials with good optoelectronic properties is critical to the creation of an excellent photoelectrochemical sensor. Generally, inorganic semiconductors nanoparticles (such as TiO2, ZnO, Cu2O, CdS, ZnS, and Bi2MoO6) have been regarded as attractive candidates for photoelectric applications due to their size-tunable optical and electronic properties, as well as their efficient multiple charge carrier generations [22,23,24,25,26,27,28,29]. Among these inorganic semiconductors, CdS, a non-centrosymmetric hexagonal zinc ore crystal structure, has been used to design the photoelectrochemical platform for CySH sensing because it can easily absorb CySH from solution via the formation of Cd-S bonds [30,31,32]. For instance, Long et al. designed an ultra-sensitive photoelectrochemical sensor with a low detection limit of 0.1 µM for CySH employing methyl viologen-coated CdS quantum dots as photoactive materials [33]. Even though ultra-detection has been achieved by using CdS as photoactive materials, the low photon-to-current conversion efficiency and photocorrosion of CdS severely limit future improvements in photoelectrochemical properties for photoelectrochemical sensors. Therefore, there is still a significant amount of interest in finding low-cost, highly efficient, stable, and sensitive photoactive materials for photoelectrochemical sensors.
Recently, a polymeric photocatalyst known as graphitic carbon nitride (g-C3N4) has received a lot of attention due to its abundance, high thermal and chemical stability, and visible light response [34,35,36,37]. On the basis of the excellent properties mentioned above, it has been reported that g-C3N4 materials improve performance for photocatalysis [38,39,40,41] and biosensing [42,43,44]. Herein, combining the excellent photoelectrochemical properties of g-C3N4 nanosheets and CdS, g-C3N4@CdS composites were synthesized in a simple way. To our knowledge, there are few reports on CySH sensing based on g-C3N4@CdS composites. Using these photoactive materials, a novel photoelectrochemical platform was developed and used to assay CySH in human urine. It was discovered that such a photoelectrochemical sensor has a fast response, high sensitivity, high stability, and excellent selectivity for CySH.

2. Materials and Methods

All chemicals were obtained from commercial sources and utilized without further purification. Thioacetamide, polyvinylpyrrolidone (PVP, K-30), melamine, and chemiluminescence reagent were obtained from Sigma-Aldrich. CdCl2·2.5H2O (99%) was purchased from Alfa. The PBS buffer was made from sodium phosphate (NaH2PO4/Na2HPO4, 81:19 (molar ratio)) and sodium chloride dissolved in deionized water at final concentrations of 10 mmol L-1 (pH: 7.4).
XRD patterns of the samples were carried out with a D/MAX 2500V/PC X-ray diffraction (Cu Kα radiation, λ = 0.15406 nm) at 40 kV and 30 mA. The scanning range of 2θ was 10–80°. XPS was performed with an ESCALAB-MK II 250 photoelectron spectrometer using Al as the X-ray source for excitation. TEM measurements were carried out with a TECNAI G2 high-resolution transmission electron microscope and a 200 kV accelerating voltage. Fluorescence emission spectra were recorded on a Hitachi F-4600 fluorescence spectrophotometer at an excitation wavelength of 325 nm. All electrochemical experiments were performed with a CHI660A Electrochemical Workstation (CHI). ITO or modified ITO was used as the working electrode, a platinum wire was used as the auxiliary electrode, and an Ag/AgCl electrode (3 M KCl) was used as the reference electrode. In this work, E(RHE) = E(Ag/AgCl) + 0.198 V + 0.059 pH. The PBS solution was used as a supporting electrolyte. Before the experiment, the PBS solution was bubbled with N2 for 15 min. LED light with different wavelengths (365 nm, 425 nm, 470 nm, 545 nm, and 645 nm, Beijing Perfectlight Technology) was used as the source of the photoelectrochemical sensor. Electrochemical impedance spectroscopy (EIS) measurements were carried out using a Solartron 1255 B Frequency Response Analyzer (Solartron Inc.UK) in a 0.1 M PBS aqueous solution.
g-C3N4 was first synthesized with thermal polycondensation of melanin in argon at 550 °C for 2 h at a heating rate of 2.5 °C min−1. The obtained bulk g-C3N4 was rinsed with water and ethanol three times and dried at 70 °C. Then, 50 mg of bulk g-C3N4 was dispersed in 50 mL of water and ultrasonicated for 2 h. The ultrathin g-C3N4 (utg-C3N4) was dispersed in the supernatant, which was centrifuged at 3500 r·min−1 for 20 min, and the supernatant was collected and prepared for use. In order to prepare the g-C3N4@CdS nanocomposites, 1.3 mL of NH2OH and 0.2 g of PVP were dispersed into a 25 mL g-C3N4 dispersion solution. After vigorous stirring, 0.057 g CdCl2·2.5H2O and 0.1 g thioacetamide were added. The mixed solution was reacted for 4 h at 80 °C under vigorous stirring. Finally, the obtained product was centrifuged at 6500 rpm and then dried at 60 °C in a vacuum drier. As a control experiment, the bare GR and CdS were also synthesized under the same process without adding g-C3N4 and CdCl2·2.5H2O, respectively.
The ITO electrode was washed with 1 M NaOH and then 30% H2O2. Then, the obtained electrode was further cleaned with acetone and twice-distilled water three times. Finally, the obtained electrode was dried at room temperature. To create a g-C3N4@CdS modified ITO electrode, 100 L of the g-C3N4@CdS suspension solution (5 mg mL−1) was cast onto the ITO electrode and dried at room temperature. Similar steps were used to prepare the CdS-modified ITO electrode.
The working electrode was illuminated from the back (an illumination power on the working electrode of 73.89 mW/cm2), effectively preventing interference from the color sample. Each sample was detected three times, and the average value was calculated. The photoelectrocurrent was obtained using the following procedure: I = Ismpale − Iblank (Ismpale, the photoelectrochemical current of the sample; Iblank, the photoelectrochemical current without the sample).

3. Results and Discussion

We first characterized the XRD patterns of the as-prepared photoactive materials. As shown in Figure 1a, two diffraction peaks were observed on the XRD pattern of g-C3N4 nanosheets. The stacking of conjugated double bonds was represented by one prominent diffraction peak at 27.4°. This peak was indexed as the (002) peak for graphitic materials and closely matched the interlayer d-spacing of 0.336 nm for g-C3N4. The other faint diffraction peak at 13.0° corresponded to a 0.672 nm interplanar separation, which was indexed as (100) in JCPDS 87-1526. These two diffraction peaks correlate well with the g-C3N4 values given in the literature [37,40,44]. Four discernible diffraction peaks were observed on the XRD spectra of CdS nanoparticles, which were indexed as (100), (101), (110), and (112) in JCPDS 89-0440 [45]. The g-C3N4@CdS composite samples presented diffraction peaks for both g-C3N4 and CdS, which demonstrated that g-C3N4@CdS was successfully synthesized.
The morphology of the g-C3N4@CdS nanocomposites was investigated by TEM. As shown in Figure 1b and Figure S1, CdS nanoparticles with diameters of about 20 nm were found to be uniformly dispersed on the surface of g-C3N4 nanosheets. The efficient optoelectronic properties of the g-C3N4@CdS nanocomposites could be guaranteed by the good distribution and high coverage of CdS on the g-C3N4 nanosheets. XPS measurements were further carried out to investigate the chemical composition of g-C3N4@CdS nanocomposites. High-resolution spectra of C1s (Figure 1c) at 285.5 eV and N1s at 398.5 eV were assigned to the sp2 C=N bond in the s-triazine ring [23,40]. Peaks in the C1s zone at 288.3 eV and 284.6 eV are attributed to electrons originating from an sp2 C atom attached to a -NH2 group and an aromatic carbon atom [35]. The contributions in the N1s zone at 399.5 and 401.2 eV are ascribed to N atoms that are bound to three C atoms; these N atoms are present in the heptazine ring and as a bridging atom, respectively [37]. After the dispersion of CdS nanoparticles onto the surface of g-C3N4 nanosheets, the peaks of Cd3d and S2p could be observed on the XPS spectra of g-C3N4@CdS. As shown in Figure 1c, two peaks at 405.2 eV and 411.5 eV were observed on the high-resolution spectrum of Cd3d, which could be assigned to the Cd2+ ions of the CdS. An s2p peak at 161.5 eV was also observed, as expected for the sulfide in CdS [45]. All the above results demonstrate that g-C3N4@CdS nanocomposites were successfully synthesized.
The as-prepared g-C3N4@CdS photoactive materials were further applied to design a photoelectrochemical platform for CySH sensing. As shown in Figure 2a, a photocurrent of 353 nA was observed on a bare CdS modified ITO electrode (CdS/ITO) in 0.1 M PBS under 545 nm light irradiation, whereas a photocurrent of 449 nA was observed on a g-C3N4@CdS modified ITO electrode (g-C3N4@CdS/ITO), indicating that the addition of g-C3N4 improves the photocurrent conversion efficiency of CdS. To explore the above reason, electrochemical impedance spectroscopy was conducted at 0.0 V under 545 nm light illumination. As shown in Figure 2b, the charge transfer resistance of g-C3N4@CdS/ITO was 189.7 kΩ, which is much smaller than that of bare materials (517.2 kΩ). This smaller arc radius implies a much higher charge transfer efficiency [23,25]. These results were supported by the data of photoluminescence spectra (Figure S2). Furthermore, after adding 50 µM CySH to the electrolyte solution, the photocurrent reached 994 nA on g-C3N4@CdS/ITO, nearly 2.1 times that of CdS/ITO. The increase in the photocurrent was attributed to the oxidization of CySH by the photogenerated holes, effectively avoiding electron–hole recombination. The photoelectrochemical process of g-C3N4@CdS/ITO for CySH oxidation is proposed in Scheme 1. Upon irradiation with light, CdS was excited and underwent charge separation to yield electrons (e-) and holes (h+). The photoelectrons can arrive at g-C3N4 through electron tunneling and subsequently to the ITO electrode [46,47,48]. Finally, the photocurrent was produced. When CySH was introduced, the h+ of the CdS could be refilled by electrons from the CySH, ensuring that the occupied holes were available for the following excitation. The photocurrent could be greatly improved by this process. It was noted that the intensity of the photocurrent was consistently enhanced by increasing concentrations of CySH, thus promising the quantitative determination of CySH.
The irradiation wavelength for the g-C3N4@CdS-based photoelectrochemical platform was optimized. As shown in Figure 2c, under 645 nm irradiation, a slight photocurrent of 59 nA was observed on the g-C3N4@CdS/ITO with an applied potential of 0.0 V in 50 μM CySH electrolyte solution. When the irradiation wavelength was decreased to 545 nm, the photocurrent significantly increased to 971 nA under the same conditions. Furthermore, we found that the photocurrent increased sharply as the exciting wavelength below 470 nm. Although the photocurrent responses exhibited a preferential sensitivity at short wavelengths, this causes additional issues of serious interference due to coexisting species during actual sample examinations (e.g., glucose and ascorbic acid), which should be avoidable at weaker wavelength irradiations [23]. Therefore, 545 nm was chosen for the detection of CySH, which could simultaneously ensure the excellent sensitivity and stability of this photoelectrochemical sensor.
We further investigated the photoelectrochemical properties of the as-prepared photoelectrochemical sensor under different working potentials. As shown in Figure 2c, the photocurrent increment increased significantly as the applied potential rose from −0.2 to 0.0 V and tended to reach a maximum at 0.0 V, while from 0.0 to 0.3 V, essentially no discernible change was observed under 545 nm illumination in 50 μM CySH solution. The photocurrent increased again as the applied voltage increased to 0.4 V. Nevertheless, the photocurrent at 0.0 V was 70.1% of the photocurrent at 0.5 V, showing adequate sensitivity for the photoelectrochemical detection of CySH. Additionally, a voltage of 0.0 V was selected for the photoelectrochemical sensing of CySH in the subsequent tests to make it convenient to develop the two-electrode system for a photoelectrochemical CySH instrument.
The photocurrent of the photoelectrochemical sensor was investigated with increasing concentrations of CySH to validate its performance in the determination of CySH. As shown in Figure 3a, the photocurrent increased as the concentration of CySH increased. Good linear ranges are from 5 to 190 μM (R2 = 0.997) for the detection of CySH with a lower detection limit of 1.56 μM. The ultra-low detection limit was superior to some CySH sensors reported in the literature [49,50,51,52,53,54]. The photoelectrochemical sensor has outstanding features, such as a wide response range, a low detection limit, and a rapid response, which ensures that real samples are detected. The reproducibility of the g-C3N4@CdS/ITO was investigated. Six electrodes were tested at the same concentration of CySH (50 μM), and the results showed that g-C3N4@CdS/ITO had good reproducibility with a relative 2.9%. Additionally, since stability is crucial for sensors, the operation stability as well as the long-term stability of such a photochemical sensor were studied. The photocurrent response remained at 90.3% of the value of the initial reaction after four weeks. Additionally, even after more than 1400 s of scanning, the photocurrent of CySH maintained 96.6% of its starting value (Figure 3c), which meets the requirements for CySH sensing. Yet, the photocurrent could decrease further after more than 1400 s test due to photocorrosion, a similar phenomenon was observed in the literature [55]. All these results demonstrated that g-C3N4@CdS would perform exceptionally well in the quantitative analysis of CySH.
The selectivity of the photoelectrochemical sensor was confirmed by evaluating its photoresponse toward various probable interference species found in the detection solution containing 50 μmol L−1 CySH. In total, 1000 times Na+, K+, Ca2+, Mg2+, Zn2+, Ni2+, Cl, NO3, ClO4, CO32−, SO42−, and PO43−; 500 times L-proline, ethanol, L-histidine, L-glycine, methanol, L-threonine, fructose, and glucose; and 100 times uric acid were texted. As shown in Figure 4, the results revealed that these species did not cause significant interference with the detection of CySH on the as-prepared g-C3N4@CdS-based photoelectrochemical platform.
To confirm the feasibility, the proposed method was applied to determine CySH in human urine. Three human serum samples were detected with the g-C3N4@CdS-based photoelectrochemical platform. The corresponding CySH detection results are shown in Table 1. All the results were found to be consistent with the chemiluminescence results. The performance of practical investigations demonstrated that the current photoelectrochemical platform is excellent for detecting CySH.

4. Conclusions

In conclusion, a novel photoelectrochemical platform for the detection of CySH was developed by employing g-C3N4@CdS composites as photoactive materials. The photoelectrochemical sensor exhibits a linear response of 5–190 µM, a lower detection limit of 1.56 μM, a fast response time, and good long-term stability (ca. 1 month). CySH was effectively tested at low concentrations in human urine samples, which was consistent with the chemiluminescence results. Compared with previous reports, this photoelectrochemical sensor has significant advantages, such as good anti-interference qualities and a high level of stability and reproducibility, and it was also demonstrated to be a practical and cost-effective method for CySH detection.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/sym15040896/s1, Figure S1: TEM image of bare g-C3N4; Figure S2: Photoluminescence (PL) spectra of pure gC3N4 and g-C3N4@ CdS samples.

Author Contributions

Conceptualization, W.M.; data curation, H.Z. and S.Q.; investigation, G.Z. and K.Z.; methodology, H.Z. and G.Z.; resources, H.W.; validation, S.Q. and K.Z.; supervision, W.M.; writing—original draft preparation, H.Z. and W.M.; writing—review and editing, W.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the China Postdoctoral Science Foundation (grant number 2021M700651) and the Fundamental Research Funds for the Central Universities (grant numbers 3132022216 and 3132022217).

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Das, A.; Ash, D.; Fouda, A.Y.; Sudhahar, V.; Kim, Y.-M.; Hou, Y.; Hudson, F.Z.; Stansfield, B.K.; Caldwell, R.B.; McMenamin, M.; et al. Cysteine oxidation of copper transporter CTR1 drives VEGFR2 signalling and angiogenesis. Nat. Cell Biol. 2022, 24, 35–50. [Google Scholar] [CrossRef] [PubMed]
  2. Meng, J.; Fu, L.; Liu, K.; Tian, C.; Wu, Z.; Jung, Y.; Ferreira, R.B.; Carroll, K.S.; Blackwell, T.K.; Yang, J. Global profiling of distinct cysteine redox forms reveals wide-ranging redox regulation in C. elegans. Nat. Commun. 2021, 12, 1415. [Google Scholar] [CrossRef] [PubMed]
  3. Anonymous. A creatine kinase inhibitor targeting a redox-regulated cysteine. Nat. Chem. Biol. 2023. online ahead of print. [Google Scholar] [CrossRef]
  4. Paulsen, C.E.; Carroll, K.S. Cysteine-Mediated Redox Signaling: Chemistry, Biology, and Tools for Discovery. Chem. Rev. 2013, 113, 4633–4679. [Google Scholar] [CrossRef] [PubMed]
  5. Spataru, N.; Sarada, B.V.; Popa, E.; Tryk, D.A.; Fujishima, A. Voltammetric determination of L-cysteine at conductive diamond electrodes. Anal. Chem. 2001, 73, 514–519. [Google Scholar] [CrossRef]
  6. El-Bindary, M.A.; El-Desouky, M.G.; El-Bindary, A.A. Metal–organic frameworks encapsulated with an anticancer compound as drug delivery system: Synthesis, characterization, antioxidant, anticancer, antibacterial, and molecular docking investigation. Appl. Organomet. Chem. 2022, 36, e6660. [Google Scholar] [CrossRef]
  7. Altalhi, T.A.; Ibrahim, M.M.; Mersal, G.A.M.; Mahmoud, M.H.H.; Kumeria, T.; El-Desouky, M.G.; El-Bindary, A.A.; El-Bindary, M.A. Adsorption of doxorubicin hydrochloride onto thermally treated green adsorbent: Equilibrium, kinetic and thermodynamic studies. J. Mol. Struct. 2022, 1263, 133160. [Google Scholar] [CrossRef]
  8. Huang, J.; Tao, F.; Li, F.; Cai, Z.; Zhang, Y.; Fan, C.; Pei, L. Controllable synthesis of BiPr composite oxide nanowires electrocatalyst for sensitive L-cysteine sensing properties. Nanotechnology 2022, 33, 345704. [Google Scholar] [CrossRef]
  9. Matsunaga, T.; Kondo, T.; Shitanda, I.; Hoshi, Y.; Itagaki, M.; Tojo, T.; Yuasa, M. Sensitive electrochemical detection of L-Cysteine at a screen-printed diamond electrode. Carbon 2021, 173, 395–402. [Google Scholar] [CrossRef]
  10. Luo, Y.; Zhang, L.; Liu, W.; Yu, Y.; Tian, Y. A Single Biosensor for Evaluating the Levels of Copper Ion and L-Cysteine in a Live Rat Brain with Alzheimer’s Disease. Angew. Chem. Int. Ed. 2015, 54, 14053–14056. [Google Scholar] [CrossRef]
  11. Yu, Y.; Lee, S.J.; Theerthagiri, J.; Fonseca, S.; Pinto, L.M.C.; Maia, G.; Choi, M.Y. Reconciling of experimental and theoretical insights on the electroactive behavior of C/Ni nanoparticles with AuPt alloys for hydrogen evolution efficiency and Non-enzymatic sensor. Chem. Eng. J. 2022, 435, 134790. [Google Scholar] [CrossRef]
  12. Theerthagiri, J.; Lee, S.J.; Karuppasamy, K.; Park, J.; Yu, Y.; Kumari, M.L.A.; Chandrasekaran, S.; Kim, H.-S.; Choi, M.Y. Fabrication strategies and surface tuning of hierarchical gold nanostructures for electrochemical detection and removal of toxic pollutants. J. Hazard. Mater. 2021, 420, 126648. [Google Scholar] [CrossRef] [PubMed]
  13. Wu, Q.; Wang, P.; Yang, X.; Wei, M.; Zhang, M. Turn Off-On Electrochemiluminescence Sensor Based on MnO2/Carboxylated Graphitic Carbon Nitride Nanocomposite for Ultrasensitive L-Cysteine Detection. J. Electrochem. Soc. 2019, 166, B994–B999. [Google Scholar] [CrossRef]
  14. Zhu, R.; Zhang, Y.; Fang, X.; Cui, X.; Wang, J.; Yue, C.; Fang, W.; Zhao, H.; Li, Z. In situ sulfur-doped graphitic carbon nitride nanosheets with enhanced electrogenerated chemiluminescence used for sensitive and selective sensing of l-cysteine. J. Mater. Chem. B 2019, 7, 2320–2329. [Google Scholar] [CrossRef]
  15. Meng, H.; Chen, M.; Mo, F.; Guo, J.; Liu, P.; Fu, Y. Construction of self-enhanced photoelectrochemical platform for l-cysteine detection via electron donor-acceptor type coumarin 545 aggregates. Chem. Commun. 2021, 57, 11557–11560. [Google Scholar] [CrossRef]
  16. Peng, J.; Huang, Q.; Liu, Y.; Huang, Y.; Zhang, C.; Xiang, G. Photoelectrochemical Detection of L-cysteine with a Covalently Grafted ZnTAPc-Gr-based Probe. Electroanalysis 2020, 32, 1237–1242. [Google Scholar] [CrossRef]
  17. Yu, S.-Y.; Gao, Y.; Chen, F.-Z.; Fan, G.-C.; Han, D.-M.; Wang, C.; Zhao, W.-W. Fast electrochemical deposition of CuO/Cu2O heterojunction photoelectrode: Preparation and application for rapid cathodic photoelectrochemical detection of L-cysteine. Sens. Actuators B-Chem. 2019, 290, 312–317. [Google Scholar] [CrossRef]
  18. Cifteci, A.; Celik, S.E.; Apak, R. Gold-Nanoparticle Based Turn-on Fluorometric Sensor for Quantification of Sulfhydryl and Disulfide Forms of Biothiols: Measurement of Thiol/Disulfide Homeostasis. Anal. Lett. 2022, 55, 648–664. [Google Scholar] [CrossRef]
  19. Niu, L.Y.; Chen, Y.Z.; Zheng, H.R.; Wu, L.Z.; Tung, C.H.; Yang, Q.Z. Design strategies of fluorescent probes for selective detection among biothiols. Chem. Soc. Rev. 2015, 44, 6143–6160. [Google Scholar] [CrossRef]
  20. Ta, H.Y.; Collin, F.; Perquis, L.; Poinson, V.; Ong-Meang, V.; Couderc, F. Twenty years of amino acid determination using capillary electrophoresis: A review. Anal. Chim. Acta 2021, 1174, 338233. [Google Scholar] [CrossRef]
  21. Li, Z.; Liu, M.; Chen, C.; Pan, Y.; Cui, X.; Sun, J.; Zhao, F.; Cao, Y. Simultaneous determination of serum homocysteine, cysteine, and methionine in patients with schizophrenia by liquid chromatography-tandem mass spectrometry. Biomed. Chromatogr. 2022, 36, e5366. [Google Scholar] [CrossRef]
  22. Ma, W.; Han, D.; Gan, S.; Zhang, N.; Liu, S.; Wu, T.; Zhang, Q.; Dong, X.; Niu, L. Rapid and specific sensing of gallic acid with a photoelectrochemical platform based on polyaniline-reduced graphene oxide-TiO2. Chem. Commun. 2013, 49, 7842–7844. [Google Scholar] [CrossRef] [PubMed]
  23. Ma, W.; Han, D.; Zhou, M.; Sun, H.; Wang, L.; Dong, X.; Niu, L. Ultrathin g-C3N4/TiO2 composites as photoelectrochemical elements for the real-time evaluation of global antioxidant capacity. Chem. Sci. 2014, 5, 3946–3951. [Google Scholar] [CrossRef]
  24. Ma, W.; Wang, L.; Zhang, N.; Han, D.; Dong, X.; Niu, L. Biomolecule-Free, Selective Detection of o-Diphenol and Its Derivatives with WS2/TiO2-Based Photoelectrochemical Platform. Anal. Chem. 2015, 87, 4844–4850. [Google Scholar] [CrossRef] [PubMed]
  25. Wang, L.; Han, D.; Ni, S.; Ma, W.; Wang, W.; Niu, L. Photoelectrochemical device based on Mo-doped BiVO4 enables smart analysis of the global antioxidant capacity in food. Chem. Sci. 2015, 6, 6632–6638. [Google Scholar] [CrossRef] [Green Version]
  26. Wang, L.; Ma, W.; Gan, S.; Han, D.; Zhang, Q.; Niu, L. Engineered Photoelectrochemical Platform for Rational Global Antioxidant Capacity Evaluation Based on Ultrasensitive Sulfonated Graphene-TiO2 Nanohybrid. Anal. Chem. 2014, 86, 10171–10178. [Google Scholar] [CrossRef]
  27. Wang, H.-Y.; Xu, Y.-T.; Wang, B.; Yu, S.-Y.; Shi, X.-M.; Zhao, W.-W.; Jiang, D.; Chen, H.-Y.; Xu, J.-J. A Photoelectrochemical Nanoreactor for Single-Cell Sampling and Near Zero-Background Faradaic Detection of Intracellular microRNA. Angew. Chem. Int. Ed. 2022, 134, e202212752. [Google Scholar]
  28. Zhao, W.-W.; Xu, J.-J.; Chen, H.-Y. Photoelectrochemical DNA Biosensors. Chem. Rev. 2014, 114, 7421–7441. [Google Scholar] [CrossRef]
  29. Zhao, W.-W.; Xu, J.-J.; Chen, H.-Y. Photoelectrochemical bioanalysis: The state of the art. Chem. Soc. Rev. 2015, 44, 729–741. [Google Scholar] [CrossRef]
  30. Barroso, J.; Diez-Buitrago, B.; Saa, L.; Moller, M.; Briz, N.; Pavlov, V. Specific bioanalytical optical and photoelectrochemical assays for detection of methanol in alcoholic beverages. Biosens. Bioelectron. 2018, 101, 116–122. [Google Scholar] [CrossRef] [Green Version]
  31. Shen, Q.; Shi, X.; Fan, M.; Han, L.; Wang, L.; Fan, Q. Highly sensitive photoelectrochemical cysteine sensor based on reduced graphene oxide/CdS:Mn nanocomposites. J. Electroanal. Chem. 2015, 759, 61–66. [Google Scholar] [CrossRef]
  32. Wang, Q.; Ruan, Y.-F.; Zhao, W.-W.; Lin, P.; Xu, J.-J.; Chen, H.-Y. Semiconducting Organic-Inorganic Nanodots Heterojunctions: Platforms for General Photoelectrochemical Bioanalysis Application. Anal. Chem. 2018, 90, 3759–3765. [Google Scholar] [CrossRef] [PubMed]
  33. Long, Y.-T.; Kong, C.; Li, D.-W.; Li, Y.; Chowdhury, S.; Tian, H. Ultrasensitive Determination of Cysteine Based on the Photocurrent of Nafion-Functionalized CdS-MV Quantum Dots on an ITO Electrode. Small 2011, 7, 1624–1628. [Google Scholar] [CrossRef] [PubMed]
  34. Lin, L.; Yu, Z.; Wang, X. Crystalline Carbon Nitride Semiconductors for Photocatalytic Water Splitting. Angew. Chem. Int. Ed. 2019, 58, 6164–6175. [Google Scholar] [CrossRef]
  35. Wang, Y.; Wang, X.; Antonietti, M. Polymeric Graphitic Carbon Nitride as a Heterogeneous Organocatalyst: From Photochemistry to Multipurpose Catalysis to Sustainable Chemistry. Angew. Chem. Int. Ed. 2012, 51, 68–89. [Google Scholar] [CrossRef]
  36. Zhang, J.; Chen, Y.; Wang, X. Two-dimensional covalent carbon nitride nanosheets: Synthesis, functionalization, and applications. Energy Environ. Sci. 2015, 8, 3092–3108. [Google Scholar] [CrossRef]
  37. Zheng, Y.; Lin, L.; Wang, B.; Wang, X. Graphitic Carbon Nitride Polymers toward Sustainable Photoredox Catalysis. Angew. Chem. Int. Ed. 2015, 54, 12868–12884. [Google Scholar] [CrossRef]
  38. Cao, S.; Low, J.; Yu, J.; Jaroniec, M. Polymeric Photocatalysts Based on Graphitic Carbon Nitride. Adv. Mater. 2015, 27, 2150–2176. [Google Scholar] [CrossRef]
  39. Fu, J.; Yu, J.; Jiang, C.; Cheng, B. g-C3N4-Based Heterostructured Photocatalysts. Adv. Energy Mater. 2018, 8, 1701503. [Google Scholar] [CrossRef]
  40. Ong, W.-J.; Tan, L.-L.; Ng, Y.H.; Yong, S.-T.; Chai, S.-P. Graphitic Carbon Nitride (g-C3N4)-Based Photocatalysts for Artificial Photosynthesis and Environmental Remediation: Are We a Step Closer To Achieving Sustainability? Chem. Rev. 2016, 116, 7159–7329. [Google Scholar] [CrossRef]
  41. Yu, H.; Shi, R.; Zhao, Y.; Bian, T.; Zhao, Y.; Zhou, C.; Waterhouse, G.I.N.; Wu, L.-Z.; Tung, C.-H.; Zhang, T. Alkali-Assisted Synthesis of Nitrogen Deficient Graphitic Carbon Nitride with Tunable Band Structures for Efficient Visible-Light-Driven Hydrogen Evolution. Adv. Mater. 2017, 29, 1605148. [Google Scholar] [CrossRef] [PubMed]
  42. Chen, L.; Zeng, X.; Si, P.; Chen, Y.; Chi, Y.; Kim, D.-H.; Chen, G. Gold Nanoparticle-Graphite-Like C3N4 Nanosheet Nanohybrids Used for Electrochemiluminescent Immunosensor. Anal. Chem. 2014, 86, 4188–4195. [Google Scholar] [CrossRef] [PubMed]
  43. Chen, Y.; Tan, C.; Zhang, H.; Wang, L. Two-dimensional graphene analogues for biomedical applications. Chem. Soc. Rev. 2015, 44, 2681–2701. [Google Scholar] [CrossRef] [PubMed]
  44. Zhou, Z.; Zhang, Y.; Shen, Y.; Liu, S.; Zhang, Y. Molecular engineering of polymeric carbon nitride: Advancing applications from photocatalysis to biosensing and more. Chem. Soc. Rev. 2018, 47, 2298–2321. [Google Scholar] [CrossRef]
  45. Chen, J.; Wu, X.-J.; Yin, L.; Li, B.; Hong, X.; Fan, Z.; Chen, B.; Xue, C.; Zhang, H. One-pot Synthesis of CdS Nanocrystals Hybridized with Single-Layer Transition-Metal Dichalcogenide Nanosheets for Efficient Photocatalytic Hydrogen Evolution. Angew. Chem. Int. Ed. 2015, 54, 1210–1214. [Google Scholar] [CrossRef]
  46. Chen, P.; Liu, F.; Ding, H.; Chen, S.; Chen, L.; Li, Y.-J.; Au, C.-T.; Yin, S.-F. Porous double-shell CdS@C3N4 octahedron derived by in situ supramolecular self-assembly for enhanced photocatalytic activity. Appl. Catal. B-Environ. 2019, 252, 33–40. [Google Scholar] [CrossRef]
  47. Dai, X.; Xie, M.; Meng, S.; Fu, X.; Chen, S. Coupled systems for selective oxidation of aromatic alcohols to aldehydes and reduction of nitrobenzene into aniline using CdS/g-C3N4 photocatalyst under visible light irradiation. Appl. Catal. B-Environ. 2014, 158, 382–390. [Google Scholar] [CrossRef]
  48. Jo, W.-K.; Selvam, N.C.S. Z-scheme CdS/g-C3N4 composites with RGO as an electron mediator for efficient photocatalytic H-2 production and pollutant degradation. Chem. Eng. J. 2017, 317, 913–924. [Google Scholar] [CrossRef]
  49. Selvarajan, S.; Alluri, N.R.; Chandrasekhar, A.; Kim, S.-J. Direct detection of cysteine using functionalized BaTiO3 nanoparticles film based self-powered biosensor. Biosens. Bioelectron. 2017, 91, 203–210. [Google Scholar] [CrossRef]
  50. Khamcharoen, W.; Henry, C.S.; Siangproh, W. A novel L-cysteine sensor using in-situ electropolymerization of L-cysteine: Potential to simple and selective detection. Talanta 2022, 237, 122983. [Google Scholar] [CrossRef]
  51. Atacan, K. CuFe2O4/reduced graphene oxide nanocomposite decorated with gold nanoparticles as a new electrochemical sensor material for L-cysteine detection. J. Alloys Compd. 2019, 791, 391–401. [Google Scholar] [CrossRef]
  52. Bonacin, J.A.; Dos Santos, P.L.; Katic, V.; Foster, C.W.; Banks, C.E. Use of Screen-printed Electrodes Modified by Prussian Blue and Analogues in Sensing of Cysteine. Electroanalysis 2018, 30, 170–179. [Google Scholar] [CrossRef]
  53. Koczorowski, T.; Rebis, T.; Szczolko, W.; Antecka, P.; Teubert, A.; Milczarek, G.; Goslinski, T. Reduced graphene oxide/iron(II) porphyrazine hybrids on glassy carbon electrode for amperometric detection of NADH and L-cysteine. J. Electroanal. Chem. 2019, 848, 113322. [Google Scholar] [CrossRef]
  54. Castro e Silva, C.d.C.; Breitkreitz, M.C.; Santhiago, M.; Correa, C.C.; Kubota, L.T. Construction of a new functional platform by grafting poly(4-vinylpyridine) in multi-walled carbon nanotubes for complexing copper ions aiming the amperometric detection of L-cysteine. Electrochim. Acta 2012, 71, 150–158. [Google Scholar] [CrossRef]
  55. Chen, Y.; Zhong, W.; Chen, F.; Wang, P.; Fan, J.; Yu, H. Photoinduced self-stability mechanism of CdS photocatalyst: The dependence of photocorrosion and H2-evolution performance. J. Mater. Sci. Technol. 2022, 121, 19–27. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns of g-C3N4, CdS, and g-C3N4@CdS nanocomposites. (b) TEM image of g-C3N4@CdS nanocomposites. (c) High-resolution XPS spectra of C 1s, N1s, Cd3d, and S2p regions of g-C3N4@CdS nanocomposites.
Figure 1. (a) XRD patterns of g-C3N4, CdS, and g-C3N4@CdS nanocomposites. (b) TEM image of g-C3N4@CdS nanocomposites. (c) High-resolution XPS spectra of C 1s, N1s, Cd3d, and S2p regions of g-C3N4@CdS nanocomposites.
Symmetry 15 00896 g001
Scheme 1. Schematic illustration of the photoelectrochemical process for CySH sensing at g-C3N4@CdS modified ITO electrode.
Scheme 1. Schematic illustration of the photoelectrochemical process for CySH sensing at g-C3N4@CdS modified ITO electrode.
Symmetry 15 00896 sch001
Figure 2. (a) Photocurrent responses of CdS (a, a0) and g-C3N4@CdS/ITO electrode without (a,b) and with (a0, b0) 50 µM CySH. (b) EIS plots of CdS/ITO and g-C3N4@CdS/ITO at 0.0 V under 545 nm irradiation. (c) Photocurrent responses of g-C3N4@CdS/ITO at different potentials. (d) Photocurrent responses of g-C3N4@CdS/ITO at different wavelengths.
Figure 2. (a) Photocurrent responses of CdS (a, a0) and g-C3N4@CdS/ITO electrode without (a,b) and with (a0, b0) 50 µM CySH. (b) EIS plots of CdS/ITO and g-C3N4@CdS/ITO at 0.0 V under 545 nm irradiation. (c) Photocurrent responses of g-C3N4@CdS/ITO at different potentials. (d) Photocurrent responses of g-C3N4@CdS/ITO at different wavelengths.
Symmetry 15 00896 g002
Figure 3. (a) Photocurrent responses of g-C3N4@CdS/ITO with different concentrations of CySH. (b) The linear calibration curve of the g-C3N4@CdS/ITO for CySH. (c) The stability of 50 μM CySH on g-C3N4@CdS/ITO.
Figure 3. (a) Photocurrent responses of g-C3N4@CdS/ITO with different concentrations of CySH. (b) The linear calibration curve of the g-C3N4@CdS/ITO for CySH. (c) The stability of 50 μM CySH on g-C3N4@CdS/ITO.
Symmetry 15 00896 g003
Figure 4. Photocurrent response of g-C3N4@CdS/ITO upon addition of 1000 times Na+, K+, Ca2+, Mg2+, Zn2+, Ni2+, Cl, NO3, ClO4, CO32−, SO42−, and PO43−; 500 times L-proline, L-glycine, L-histidine, ethanol, methanol, L-threonine, fructose, and glucose; and 100 times uric acid in 0.1 M PBS (pH = 7.4) containing 50 μM CySH at 0 V under 545 nm light excitation.
Figure 4. Photocurrent response of g-C3N4@CdS/ITO upon addition of 1000 times Na+, K+, Ca2+, Mg2+, Zn2+, Ni2+, Cl, NO3, ClO4, CO32−, SO42−, and PO43−; 500 times L-proline, L-glycine, L-histidine, ethanol, methanol, L-threonine, fructose, and glucose; and 100 times uric acid in 0.1 M PBS (pH = 7.4) containing 50 μM CySH at 0 V under 545 nm light excitation.
Symmetry 15 00896 g004
Table 1. Results of analysis of CySH in real samples (n = 3).
Table 1. Results of analysis of CySH in real samples (n = 3).
SamplePhotoelectrochemical Sensor (μM)Chemiluminescence (μM)
Urine 15.6 ± 0.56.0 ± 0.3
Urine 212.7 ± 1.911.5 ± 0.6
Urine 37.3 ± 0.86.9 ± 0.2
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

Zhang, H.; Qi, S.; Wang, H.; Zhang, G.; Zhu, K.; Ma, W. Ultrasensitive Determination of L-Cysteine with g-C3N4@CdS-Based Photoelectrochemical Platform. Symmetry 2023, 15, 896. https://doi.org/10.3390/sym15040896

AMA Style

Zhang H, Qi S, Wang H, Zhang G, Zhu K, Ma W. Ultrasensitive Determination of L-Cysteine with g-C3N4@CdS-Based Photoelectrochemical Platform. Symmetry. 2023; 15(4):896. https://doi.org/10.3390/sym15040896

Chicago/Turabian Style

Zhang, Hefeng, Shengliang Qi, Haidong Wang, Guanghui Zhang, Kaixin Zhu, and Weiguang Ma. 2023. "Ultrasensitive Determination of L-Cysteine with g-C3N4@CdS-Based Photoelectrochemical Platform" Symmetry 15, no. 4: 896. https://doi.org/10.3390/sym15040896

APA Style

Zhang, H., Qi, S., Wang, H., Zhang, G., Zhu, K., & Ma, W. (2023). Ultrasensitive Determination of L-Cysteine with g-C3N4@CdS-Based Photoelectrochemical Platform. Symmetry, 15(4), 896. https://doi.org/10.3390/sym15040896

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