Next Article in Journal
Conformational Essentials Responsible for Neurotoxicity of Aβ42 Aggregates Revealed by Antibodies against Oligomeric Aβ42
Next Article in Special Issue
Carbon Quantum Dots Bridged TiO2/CdIn2S4 toward Photocatalytic Upgrading of Polycyclic Aromatic Hydrocarbons to Benzaldehyde
Previous Article in Journal
A Density Functional Theory and Microkinetic Study of Acetylene Partial Oxidation on the Perfect and Defective Cu2O (111) Surface Models
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sensitive, Selective and Reliable Detection of Fe3+ in Lake Water via Carbon Dots-Based Fluorescence Assay

1
School of Materials and Energy, Yunnan University, Kunming 650091, China
2
Yunnan Key Laboratory for Micro/Nano Materials & Technology, Yunnan University, Kunming 650091, China
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(19), 6749; https://doi.org/10.3390/molecules27196749
Submission received: 21 September 2022 / Revised: 6 October 2022 / Accepted: 6 October 2022 / Published: 10 October 2022
(This article belongs to the Special Issue New Trends in Photochemistry in Eurasia)

Abstract

:
In this study, C-dots were facilely synthesized via microwave irradiation using citric acid and ethylenediamine as carbon precursors. The fluorescence emissions of the C-dots could be selectively quenched by Fe3+, and the degree of quenching was linearly related to the concentrations of Fe3+ presented. This phenomenon was utilized to develop a sensitive fluorescence assay for Fe3+ detection with broad linear range (0–250, 250–1200 μmol/L) and low detection limit (1.68 μmol/L). Most importantly, the assay demonstrated high reliability towards samples in deionized water, tap water and lake water, which should find potential applications for Fe3+ monitoring in complicated environments.

Graphical Abstract

1. Introduction

Ferric ions (Fe3+) play essential roles and could participate in many important biological metabolic processes, such as RNA and DNA synthesis, oxygen transport, electron transfer and the formation of heme in the human body [1,2,3]. As such, proper intake of Fe3+ is of great importance for human health; too much Fe3+ consumed over a long period of time could cause harm to the human body. Studies have shown that continuous consumption of unbalanced Fe3+ might be linked to a range of conditions including Parkinson’s disease, Alzheimer’s disease and anemia [4,5,6]. As is well known, most of Fe3+ for the human body mainly comes from food and water; thus, it is of great significance to monitor Fe3+ content in drinking water. As a matter of fact, the European Union requires that the maximum permissible concentration of Fe3+ in drinking water is 3.57 μmol/L.
Compared with the traditional colorimetric method [7], atomic absorption spectrometry [8], and inductively coupled plasma mass spectrometry [9], fluorescence sensing is a relatively new detection method, which has the advantages of high sensitivity, easy operation and good reproducibility [10,11]. As such, many fluorescence sensing platforms have been developed for the sensitive and selective detection of Fe3+ [12], among which carbon dots (C-dots)-based sensing platforms have received enormous attention [10,13,14]. As a new type of carbon-based fluorescent nanomaterials, C-dots have superior properties such as excellent photoluminescence (PL) [15], high biocompatibility [16,17], facile surface functionalization [18,19] as well as low toxicity [20,21] and cost [22], which are ideal for fluorescence sensing assays development [23].
In the past few years, various C-dots-derived sensing platforms have been developed for the detection of Fe3+ (Table S1), which greatly enriched the toolbox for Fe3+ detection and monitoring. Still, these Fe3+ sensing assays generally face some challenges, such as impractical limit of detection (LOD) [24,25,26,27,28,29], in which the LODs of the assays were above the maximum permissible concentration of Fe3+ in drinking water (3.57 μmol/L) as set by the European Union. On the other hand, some assays had excellent LODs; however, their linear ranges were very narrow or biased [30,31], which significantly limited their practical applications. Furthermore, although some of the assays might have relatively good LODs and balanced linear ranges towards Fe3+, their applications in noncontrolled water samples (i.e., natural water) have not been demonstrated [32,33,34,35]. Assays that have excellent LODs, balanced linear ranges and have been demonstrated for applications in noncontrolled water samples are relatively scarce [36,37]. As such, it is still of great necessity and importance to develop C-dots-derived sensing assays capable of the sensitive and selective detection of Fe3+ in natural water samples.
Herein, we report the development of C-dots-based sensing assays for the sensitive and selective detection of Fe3+. Specifically, the C-dots used in this study were facilely prepared from citric acid and ethylenediamine using a microwave synthesizer. PL of aqueous dispersion of the C-dots could be efficiently quenched by Fe3+ solution, and the degree of quenching is linearly related to the concentration of Fe3+ solution. Exploiting this correlation, a facile sensing assay was developed with broad linear ranges (0–250, 250–1200 μmol/L) and a LOD of 1.68 μmol/L; the assay also demonstrated high selectivity towards Fe3+, free from the interference of 14 common ions. Furthermore, the assay was found to be accurate for deionized (DI) water samples, tap water samples and natural lake water samples, indicating the high practicality of this assay.

2. Results and Discussion

2.1. Characterizations of C-Dots

The as-prepared C-dots were brown gels, which were typical for citric acid and ethylenediamine-derived C-dots [38]. The C-dots were fully characterized by TEM, UV-Vis, fluorescence, FTIR, XPS and XRD to investigate their morphological, optical, chemical and structural properties. TEM was applied first to explore the morphological and structural characteristics of the sample. As can be seen, the C-dots were well-dispersed spherical particles with no agglomerates (Figure 1a), the diameters of these particles range from 2.3 to 6.8 nm, with an average diameter of 4.9 nm (Figure 1b). Furthermore, well-resolved lattice fringes with an interplanar spacing of 0.21 nm could be clearly observed in the HRTEM (Figure 1a, inset), which could be attributed to the (100) facet of graphite, indicating the successful carbonization and graphitization of the carbon precursors.
Next, we studied the spectroscopic behaviors of the C-dots with UV-Vis and fluorescence spectroscopy. The aqueous dispersion of the C-dots appeared pale yellow under ambient light and turned bright blue upon being excited by UV light at 365 nm (Figure 1c, inset). The UV-Vis absorption spectrum of C-dots demonstrated two main absorption peaks at 245 and 350 nm in the UV region (Figure 1c), which could be attributed to the n–σ* transition of -NH2 and the n–π* transition of C=O [39,40], respectively. Interestingly, unlike most of the reported C-dots that had excitation-wavelength-dependent emissions, the sample presented strong fluorescence emissions that were independent of the excitation wavelength, with the optimal emission and excitation wavelengths at 453 and 360 nm, respectively (Figure 1d). The absolute fluorescence quantum yield of the C-dots was determined to be 40.87% when the sample was dispersed in DI water with a concentration of 66.4 mg/L.
FTIR was then performed to analyze the surface functional groups of the C-dots (Figure 2a). As can be seen, the peak at 3350 cm−1 could be attributed to the stretching vibration of O-H, and peaks at 2938, 2872 and 950 cm−1 could be ascribed to the stretching vibrations of C-H [27]. The peaks at 1655, 1566, and 1438 cm−1 could be assigned to the bend vibrations from C=O [41], N-H and C-N, respectively, indicating the presence of amide bonds in the samples. Furthermore, the peaks at 1376 and 1041 cm−1 could be attributed to the stretching vibrations of the C-O bond [30,42]. Collectively speaking, the FTIR analysis indicated that there were hydroxyls, amide and alkyl groups on the surface of C-dots, which were expected since the C-dots were prepared from citric acid and ethylenediamine. The results also corresponded well with the analysis from the UV-Vis spectrum in which absorption peaks originated from C-NH2 and C=O bonds were observed.
To further explore the surface chemical states of the C-dots, the XPS spectroscopy of the sample was also investigated. Clearly, the full-scan XPS spectrum of C-dots exhibited three main peaks at 529.8, 396.8, and 284.6 eV, corresponding to O1s, N1s, and C1s, respectively (Figure 2b). Based on the XPS spectrum, the C-dots were mainly composed of carbon, nitrogen and oxygen, with a percentage of 66.16, 19.73 and 14.05%, respectively, indicating that the C-dots have been sufficiently carbonized and nitrogen-doped. As shown, the high-resolution C1s spectrum revealed three main peaks at 284.80, 286.18 and 287.79 eV, which were assigned to C-C/C=C, C-N/C-O and C=O, respectively (Figure 2c). Similarly, the high-resolution spectrum of O1s demonstrated two peaks at 530.94 and 531.99 eV, which could be attributed to C=O and C-O, respectively (Figure 2d). The high-resolution spectrum of N1s had three peaks at 399.16, 400 and 401.19 eV corresponding to pyridinic, pyrrolic and graphitic nitrogen [43], respectively (Figure 2e). In summary, the XPS analysis indicated that there was a rich presence of C=C, C=O, C-O and C-N on the surface of the C-dots, which corresponded well with the analysis from FTIR and UV-Vis spectroscopy.
To further study the structural features of the C-dots, XRD analysis was performed on the C-dots (Figure 2f). As can be seen, the XRD pattern of the C-dots demonstrated a broad peak centered around 24°, which was typical for C-dots and generally indicated the successful carbonization and synthesis of C-dots [44]. Moreover, the peak could be attributed to the (002) facet with an interlayer spacing d of 0.37 nm [45], which is larger than the typical interlayer spacing of 0.34 nm in graphite. The increased d value indicated that there was an increase in the amorphous nature of the C-dots, which could be attributed to the introduction of the oxygen-containing functional groups [46].

2.2. Sensing Assay Development

With the full characterization of C-dots, we set our step to develop a sensitive assay using C-dots as the platform. To our delight, we found that Fe3+ could efficiently quench the PL intensity of C-dots, and the degree of quenching was closely correlated to the concentrations of Fe3+, making it possible to develop a sensitive assay for the detection of Fe3+. Indeed, as the concentrations of Fe3+ gradually increased from 0 to 1200 μmol/L, the emission intensities of C-dots dispersions at 453 nm decreased accordingly (Figure 3a). To elucidate the relationship between the concentrations of Fe3+ and the PL intensities of C-dots, a scheme where F/F0 represented the y-axis and concentrations of Fe3+ stood for the x-axis was plotted (Figure 3b), where F0 and F represented the emission intensities of the blank C-dots dispersions and C-dots dispersions with Fe3+ presented, respectively. As can be seen, the PL intensities of C-dots are linearly related to the concentrations of Fe3+ in two ranges; one is from 0 to 250 μmol/L and the other one is from 250 to 1200 μmol/L (Figure 3b). As demonstrated in the calibration curves, both the two ranges have very good linear relationships between Fe3+ concentrations and the intensities of C-dots dispersions (F/F0). Specifically, the first fitting curve (0–250 μmol/L) has a linear equation of F/F0 = 1.00046–0.00157 [Fe3+] with R2 = 0.99807 (Figure 3c), and the second fitting curve (250–1200 μmol/L) has a linear equation of F/F0= 0.70532–0.00048 [Fe3+] with R2 = 0.99371 (Figure 3d). According to the well-accepted 3σ method, LOD = 3 σ/k, where σ is the standard deviation of the blank samples and k is the slope of the linear calibration plot [47,48,49]. The LOD for the assay based on the calibration curve presented in Figure 3c was determined to be 1.68 μmol/L, which is well below the maximum permissible concentration of Fe3+ in drinking water (3.57 μmol/L), as required by the European Union. In summary, the sensing assay developed in this study has a broad linear range and a very low LOD.

2.3. Selectivity and Reliability of the Assay

With the successful establishment of the sensing assay, we also carried out careful experiments to investigate the selectivity of the assay. As known, many ions often co-occurred in solutions, thus the ability to be free from the interference of other ions is very important for a metal-ion sensing assay. To our delight, the assay developed was free from most of the commonly seen metal ions. As can be seen, common monovalent ions such as Na+, K+ as well as Ag+ hardly alter the PL intensity of C-dots dispersions (Figure 4, green columns), and thus could not interfere with the detection of Fe3+. Similarly, nine commonly seen divalent metal ions, namely, Ba2+, Cd2+, Co2+, Ni2+, Zn2+, Mg2+, Cu2+, Ca2+ and Fe2+, were also tested, which also had no serious interference to the detection of Fe3+ (Figure 4, yellow columns). It is worth noting that, as Fe2+ had no significant interference to Fe3+, the current assay has the potential to be developed as a Fe2+ sensing assay by adding oxidants into the solutions in advance of testing [50]. Lastly, we also tested common trivalent metal ions such as Cr3+ and Al3+; they also had no interference to the detection of Fe3+ (Figure 4, purple columns). Encouraged by these findings, we further evaluated the selectivity of Fe3+ detection in the presence of second cations including Na+, Zn2+ and Cr3+. The results showed that the co-presence of these cations did not have obvious interferences on the sensing of Fe3+ (Figure 4, red columns). In summary, the sensing assay developed in this study had excellent selectivity towards Fe3+ and was free from the interference of most metal ions.
To test the reliability of the sensing assay developed, four spike solutions of Fe3+ in DI water were prepared and their concentrations were determined using the above-established calibrations curves, which were then compared with the actual concentrations. To our delight, the assay developed in this study seemed very reliable towards all the spike solutions tested (Table 1, rows 1–4). Specifically, the recoveries of the four spike samples ranged from 92.85 to 105.50%, which were all within the tolerance range for a sensing assay. Furthermore, extremely high recovery of 99.58% could be achieved when the concentration of Fe3+ solution was set at 50 μmol/L (Table 1, row 1).
As all the studies above were carried out with laboratory-made DI water, to explore the practicality of our assay for the detection of Fe3+ in other water samples, additional experiments were conducted. Specifically, we collected tap and lake water and the presence of Fe3+ in these two samples were investigated. As demonstrated, the fluorescence intensity of C-dots in DI water was identical to that of C-dots in tap and lake water (Figure S1), indicating that the presence of Fe3+ in these two samples was well below the LOD of this assay. Thus, to better evaluate the reliability of this assay in non-laboratory water (i.e., tap and lake water), we made seven spike solutions of Fe3+, three with tap water and four with natural lake water. To our surprise, the assay was also very reliable towards the spike solutions made from the everyday water samples. Specifically, the recoveries of the three tap water samples ranged from 92.67 to 99.01% (Table 1, rows 5–7), while the four lake water samples ranged from 93.52 to 99.45% (Table 1, rows 8–11), all of which had very excellent recovery percentages. Similarly, some extremely high recovery yields were also observed for both tap water samples (99.01%, Table 1, row 5) and lake water samples (99.45%, Table 1, row 11), demonstrating the high practicality of the assay reported in this study.
In conclusion, compared to the literature precedents [24,25,26,27,28,29] listed in Table S1, the assay developed in this study had a relatively good LOD that meets the European Union standards for the detection of drinking water. In addition, the assay also had a wide linear range that was better than most of the reported assays [24,25,26,27,28,29,30,31,32,33,34,36,37], and should be sufficient for most everyday sensing needs. Most importantly, we have demonstrated the potential application of this assay for Fe3+ sensing in natural waters, and the results were better or close to the reported work [36,37].

2.4. Possible PL Quenching Mechanisms

The quenching of a fluorophore by quenchers generally occurred via dynamic quenching (collisional inactivation) or static quenching (static complexation). In a typical static quenching, a ground-state complex is formed through the interaction between C-dots and the quencher. The complex is generally nonfluorescent and could return to the ground state without PL emissions upon absorbing a light [51]. For static quenching, the fluorescence lifetime of C-dots should stay unchanged with or without a quencher, that means τ0/τ = 1. Furthermore, the formation of a non-fluorescent complex generally should result in new peaks in the C-dots absorption spectrum.
To explore the quenching mechanism, Time-correlated single-photon counting (TCSPC) experiments were carried out to investigate the charge transfer and exciton recombination processes of C-dots in the presence and absence of Fe3+ ions. Based on the tests, the fluorescence lifetime of C-dots without Fe3+ was determined to be 11.2 ns, while that in the presence of Fe3+ was determined to be 12 ns (Figure 5a). These results indicated that the presence of Fe3+ did have an influence on the excited states of C-dots; however, the influence was quite minimal. We then carefully evaluated the complexation of C-dots with Fe3+ using UV-Vis absorption spectroscopy (Figure 5b). It was clear that a new absorption peak at 260 nm appeared for C-dots in the presence of Fe3+ (green curve vs. black curve), which might be attributed to the complexation of C-dots with Fe3+, and this is generally considered to be a sign of static quenching [52,53,54]. According to these results, we cautiously propose that the quenching of the fluorescence of C-dots by Fe3+ was mainly via the static quenching path.

3. Materials and Methods

3.1. Reagents and Materials

KCl, NaCl, AgNO3, CuSO4, CaCl2, MgCl2, CoCl2·6H2O, NiCl2, Zn(CH3COO)2, BaCl2, CdCl2, FeCl2, Pb(NO3)2, FeCl3, AlCl3, MnCl2·4H2O and CrCl3·6H2O were purchased from Energy Chemical Reagent Co., Ltd. (Shanghai, China). Citric acid and anhydrous ferric chloride (FeCl3) were obtained from Sahn Chemical Technology Co., Ltd. (Shanghai, China). Ethylene diamine was purchased from Shanghai Aladdin Biochemical Technology Co., Ltd. (Shanghai, China). The deionized water used in all experiments was made from a Master Touch-S laboratory ultrapure water machine (Master Touch, Shanghai, China). Lake water was taken from the Ze Lake at the Chenggong Campus of Yunnan University, Kunming, Yunnan, China. All the reagents were used as received without further purification, unless otherwise noted.

3.2. Synthesis of C-Dots

The C-dots used in this study were prepared as follows: briefly, a mixture of citric acid (4 g) and ethylenediamine (20 mL) was heated in a microwave synthesizer set at 160 °C for 10 min. The resulted reaction mixture was then evaporated at 80 °C for 4 h in a rotary evaporator to remove the excess ethylenediamine, which resulted in the gelatinous brown C-dots.

3.3. Characterizations of C-Dots

The UV-Vis absorption spectrum of C-dots was tested on a UV-Vis spectrophotometer (UV-2600, Shimadzu, Tokyo, Japan), in which the wavelength range was set at 195–1100 nm, the scanning speed was set at medium speed, the sampling interval was set at 1 nm, and all the samplings were repeated twice. The concentration of the C-dots dispersion used for the test was 33.2 mg/L. The fluorescence spectrum of C-dots (8.3 mg/L) was tested by a fluorescence spectrometer (FL, F97 Pro, Shanghai Prism Technology Co., Ltd., Shanghai, China), in which a three-dimensional wavelength scanning mode was adopted. The excitation and emission wavelengths were set at 220–500, and 220–700 nm, respectively; both the excitation and emission widths were set at 10 nm. The scanning speed was set at 1000 nm/min, the scanning interval was set at 1 nm and the gain was set at 650 V. The frequency-domain lifetime was measured by a FLS10000 fluorescence lifetime spectrometer (Edinburgh, UK).
The Fourier transform infrared (FTIR) spectra were measured using a FTIR spectrometer (Nicoletis10, Thermo Scientific, Waltham, MA, USA), in which samples were smeared on KBr before testing. The elemental composition of C-dots was analyzed by a multifunctional X-ray photoelectron spectroscopy (XPS, K-Alpha, Thermo, USA).
The transmission electron microscopy (TEM) of C-dots was measured on a transmission electron microscope (JEM-2100, JEOL, Tokyo, Japan), in which the acceleration voltage was set at 200 kV, and the magnification was 300,000 times. The samples for TEM testing were treated as follows: 50 μL of C-dots dispersion (0.5 g/L) were dispersed in 10 mL of water, and then sonicated for 30 min. Then, a drop of C-dots dispersion was taken and placed on carbon net until the moisture was volatilized before they were applied for the TEM test. The X-ray diffraction (XRD) pattern was obtained on an X-ray diffractometer (DX-2700BH, Haoyuan Instrument Co., Ltd., Dandong, Liaoning, China) with a wavelength (λ) of 0.15406 nm.

3.4. Sensing Assay Development

Detection of Fe3+ in aqueous solutions: Firstly, mother solution/dispersion of Fe3+ and C-dots with a concentration of 3 mmol/L and 500 mg/L, respectively, were prepared. Then, the two solutions/dispersions were mixed accordingly to generate a series of mix dispersions with the concentrations for C-dots as constant (8.3 mg/L), and the concentrations for Fe3+ as gradients (0, 50, 100, 150, 200, 250, 300, 350, 400, 450, 500, 600, 700, 800, 900, 1000, 1100, and 1200 μmol/L). The mix dispersions were allowed to set for 5 min before their PL emissions were measured.
Selectivity of the assay: All the procedures were similar to the above discussions except that the concentrations of the metal ions (Fe3+, Na+, K+, Ag+, Ba2+, Cd2+, Co2+, Ni2+, Zn2+, Mg2+, Cu2+, Ca2+, Fe2+, Cr3+, Al3+) were set as a constant (750 μmol/L).
Reliability of the assay: To test the reliability of the assay, 11 spike solutions of Fe3+ in DI water, tap water and lake water were prepared, and their concentrations were determined based on the calibration curves established in this study. The determined concentrations of Fe3+ were then compared with their actual concentrations to calculate the recovery percentage.

4. Conclusions

A facile synthesis of C-dots via microwave irradiation, with citric acid as a carbon precursor and ethylenediamine as a passivation agent, was developed. The C-dots demonstrated excitation-wavelength-independent emissions at 453 nm, and the emissions could be effectively quenched by Fe3+. Exploiting this phenomenon, a sensitive assay for the detection of Fe3+ in its aqueous solutions was developed, which achieved broad linear ranges (0–250, 250–1200 μmol/L) and excellent LOD (1.68 μmol/L). Furthermore, the assay also demonstrated excellent selectivity towards Fe3+ and was free from interference of most commonly seen metal ions. Most importantly, the assay developed in this study functioned well with high reliability towards both DI water and everyday water (i.e., tap water and lake water), demonstrating its high practicality towards Fe3+ in environmental water samples.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27196749/s1, Table S1: A summary of the representative reports in which C-dots based fluorescence assays were applied for Fe3+ detection; Figure S1: The normalized fluorescence intensities of C-dots dispersions in DI water (green column), tap water (blue column) and lake water (turquoise column), respectively.

Author Contributions

Conceptualization, Z.P. and Z.X.; methodology, Z.X., Y.J. and C.C.; validation, Z.X., C.C. and Y.L.; formal analysis, Z.X., Y.J. and Z.P.; investigation, Z.X., Y.J. and C.C.; resources, Z.P.; data curation, Z.X. and Y.J.; writing–original draft preparation, Z.X.; writing–review & editing, Z.P., Z.X., Y.J., C.C. and Y.L.; supervision, Z.P.; Project administration, Z.P.; Funding acquisition, Z.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Applied Basic Research Program of Yunnan Province under grant number: 2019FB066.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available within the article and supplementary material).

Acknowledgments

The authors thank the Advanced Analysis and Measurement Center of Yunnan University for the sample testing service.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Sample Availability

Samples of the C-dots are available from the authors.

References

  1. Du, F.Y.; Cheng, Z.F.; Tan, W.; Sun, L.S.; Ruan, G.H. Development of sulfur doped carbon quantum dots for highly selective and sensitive fluorescent detection of Fe2+ and Fe3+ ions in oral ferrous gluconate samples. Spectroc. Acta Part A Molec. Biomolec. Spectr. 2020, 226, 117602. [Google Scholar] [CrossRef]
  2. Wang, T.T.; Liu, J.Y.; An, J.D.; Shi, Y.F.; Zhang, Y.Y.; Huo, J.Z.; Huang, Z.G.; Liu, Y.Y.; Ding, B. Hydrothermal synthesis of two-dimensional cadmium(II) micro-porous coordination material based on Bi-functional building block and its application in highly sensitive detection of Fe3+ and Cr2O72−. Spectroc. Acta Part A Molec. Biomolec. Spectr. 2021, 254, 119655. [Google Scholar] [CrossRef]
  3. Zhang, J.F.; Xiang, Q.; Qiu, Q.X.; Zhu, Y.; Zhang, C. Naphthalene/anthracene chromophore-based W/S/Cu cluster-organic frameworks with adjustable Fe3+ sensing properties. J. Solid State Chem. 2021, 298, 122123. [Google Scholar] [CrossRef]
  4. Kienzl, E.; Puchinger, L.; Jellinger, K.; Linert, W.; Stachelberger, H.; Jameson, R.F. The role of transition metals in the pathogenesis of Parkinson’s disease. J. Neurol. Sci. 1995, 134, 69–78. [Google Scholar] [CrossRef]
  5. Viles, J.H. Metal ions and amyloid fiber formation in neurodegenerative diseases. Copper, zinc and iron in Alzheimer’s, Parkinson’s and prion diseases. Coord. Chem. Rev. 2012, 256, 2271–2284. [Google Scholar] [CrossRef]
  6. Saboor, M.; Zehra, A.; Hamali, H.A.; Mobarki, A.A. Revisiting iron metabolism, iron homeostasis and iron deficiency anemia. Clin. Lab. 2021, 67, 660–666. [Google Scholar] [CrossRef]
  7. Liu, B.; Zhuang, J.Y.; Wei, G. Recent advances in the design of colorimetric sensors for environmental monitoring. Environ. Sci. Nano. 2020, 7, 2195–2213. [Google Scholar] [CrossRef]
  8. Ghaedi, M.; Shokrollahi, A.; Niknam, K.; Niknam, E.; Soylak, M. Development of efficient method for preconcentration and determination of copper, nickel, zinc and iron ions in environmental samples by combination of cloud point extraction and flame atomic absorption spectrometry. Cent. Eur. J. Chem. 2009, 7, 148–154. [Google Scholar] [CrossRef]
  9. Laborda, F.; Bolea, E.; Jimenez-Lamana, J. Single particle inductively coupled plasma mass spectrometry: A powerful tool for nanoanalysis. Anal. Chem. 2014, 86, 2270–2278. [Google Scholar] [CrossRef]
  10. Batool, M.; Junaid, H.M.; Tabassum, S.; Kanwal, F.; Abid, K.; Fatima, Z.; Shah, A.T. Metal ion detection by carbon dots-a review. Crit. Rev. Anal. Chem. 2022, 52, 756–767. [Google Scholar] [CrossRef]
  11. Olorunyomi, J.F.; White, J.F.; Gengenbach, T.R.; Caruso, R.A.; Doherty, C.M. Fabrication of a reusable carbon dot/gold nanoparticle/metal-organic framework film for fluorescence detection of lead ions in water. ACS Appl. Mater. Interfaces 2022, 14, 35755–35768. [Google Scholar] [CrossRef] [PubMed]
  12. Li, L.J.; Wang, J.H.; Xu, S.H.; Li, C.X.; Dong, B. Recent progress in fluorescent probes for metal ion detection. Front. Chem. 2022, 10, 875241. [Google Scholar] [CrossRef] [PubMed]
  13. Bilge, S.; Karadurmus, L.; Sinag, A.; Ozkan, S.A. Green synthesis and characterization of carbon-based materials for sensitive detection of heavy metal ions. Trac-Trends Anal. Chem. 2021, 145, 116473. [Google Scholar] [CrossRef]
  14. Dolai, S.; Bhunia, S.K.; Rajendran, S.; UshaVipinachandran, V.; Ray, S.C.; Kluson, P. Tunable fluorescent carbon dots: Synthesis progress, fluorescence origin, selective and sensitive volatile organic compounds detection. Crit. Rev. Solid State Mat. Sci. 2021, 46, 349–370. [Google Scholar] [CrossRef]
  15. Peng, Z.L.; Ji, C.Y.; Zhou, Y.Q.; Zhao, T.S.; Leblanc, R.M. Polyethylene glycol (PEG) derived carbon dots: Preparation and applications. Appl. Mater. Today. 2020, 20, 10067. [Google Scholar] [CrossRef]
  16. Peng, Z.L.; Han, X.; Li, S.H.; Al-Youbi, A.O.; Bashammakh, A.S.; El-Shahawi, M.S.; Leblanc, R.M. Carbon dots: Biomacromolecule interaction, bioimaging and nanomedicine. Coord. Chem. Rev. 2017, 343, 256–277. [Google Scholar] [CrossRef]
  17. Peng, Z.L.; Miyanji, E.H.; Zhou, Y.Q.; Pardo, J.; Hettiarachchi, S.D.; Li, S.H.; Blackwelder, P.L.; Skromne, I.; Leblanc, R.M. Carbon dots: Promising biomaterials for bone-specific imaging and drug delivery. Nanoscale 2017, 9, 17533–17543. [Google Scholar] [CrossRef]
  18. Hoang, V.C.; Dave, K.; Gomes, V.G. Carbon quantum dot-based composites for energy storage and electrocatalysis: Mechanism, applications and future prospects. Nano Energy. 2019, 66, 104093. [Google Scholar] [CrossRef]
  19. Zhou, Y.Q.; Mintz, K.J.; Cheng, L.; Chen, J.Y.; Ferreira, B.; Hettiarachchi, S.D.; Liyanage, P.Y.; Seven, E.S.; Miloserdov, N.; Pandey, R.R.; et al. Direct conjugation of distinct carbon dots as Lego-like building blocks for the assembly of versatile drug nanocarriers. J. Colloid Interface Sci. 2020, 576, 412–425. [Google Scholar] [CrossRef]
  20. Wang, B.Y.; Cai, H.J.; Waterhouse, G.I.N.; Qu, X.L.; Yang, B.; Lu, S.Y. Carbon dots in bioimaging, biosensing and therapeutics: A comprehensive review. Small Sci. 2022, 2, 2200012. [Google Scholar] [CrossRef]
  21. Wu, J.J.; Chen, G.L.; Jia, Y.N.; Ji, C.Y.; Wang, Y.T.; Zhou, Y.Q.; Leblanc, R.M.; Peng, Z.L. Carbon dot composites for bioapplications: A review. J. Mater. Chem. B. 2022, 10, 843–869. [Google Scholar] [CrossRef]
  22. Ji, C.Y.; Han, Q.R.; Zhou, Y.Q.; Wu, J.J.; Shi, W.Q.; Gao, L.P.; Leblanc, R.M.; Peng, Z.L. Phenylenediamine-derived near infrared carbon dots: The kilogram-scale preparation, formation process, photoluminescence tuning mechanism and application as red phosphors. Carbon 2022, 192, 198–208. [Google Scholar] [CrossRef]
  23. Ji, C.Y.; Zhou, Y.Q.; Leblanc, R.M.; Peng, Z.L. Recent developments of carbon dots in biosensing: A review. ACS Sens. 2020, 5, 2724–2741. [Google Scholar] [CrossRef]
  24. Lin, H.T.; Huang, J.; Ding, L.Y. A Recyclable Optical Fiber Sensor Based on Fluorescent Carbon Dots for the Determination of Ferric Ion Concentrations. J. Light. Technol. 2019, 37, 4815–4822. [Google Scholar] [CrossRef]
  25. Chandra, S.; Laha, D.; Pramanik, A.; Chowdhuri, A.R.; Karmakar, P.; Sahu, S.K. Synthesis of highly fluorescent nitrogen and phosphorus doped carbon dots for the detection of Fe3+ ions in cancer cells. Luminescence 2016, 31, 81–87. [Google Scholar] [CrossRef]
  26. Liu, Y.; Liu, Y.; Park, S.J.; Zhang, Y.; Kim, T.; Chae, S.; Park, M.; Kim, H.Y. One-step synthesis of robust nitrogen-doped carbon dots: Acid-evoked fluorescence enhancement and their application in Fe3+ detection. J. Mater. Chem. A 2015, 3, 17747–17754. [Google Scholar] [CrossRef]
  27. Zulfajri, M.; Gedda, G.; Chang, C.J.; Chang, Y.P.; Huang, G.G. Cranberry beans derived carbon dots as a potential fluorescence sensor for selective detection of Fe3+ ions in aqueous solution. ACS Omega 2019, 4, 15382–15392. [Google Scholar] [CrossRef] [Green Version]
  28. Yu, J.; Xu, C.X.; Tian, Z.S.; Lin, Y.; Shi, Z.L. Facilely synthesized N-doped carbon quantum dots with high fluorescent yield for sensing Fe3+. New J. Chem. 2016, 40, 2083–2088. [Google Scholar] [CrossRef]
  29. Ding, H.; Wei, J.S.; Xiong, H.M. Nitrogen and sulfur co-doped carbon dots with strong blue luminescence. Nanoscale 2014, 6, 13817–13823. [Google Scholar] [CrossRef]
  30. Lu, F.; Zhou, Y.H.; Wu, L.H.; Qian, J.; Cao, S.; Deng, Y.F.; Chen, Y. Highly fluorescent nitrogen-doped graphene quantum dots’ synthesis and their applications as Fe(III) ions sensor. Int. J. Opt. 2019, 2019, 8724320. [Google Scholar] [CrossRef]
  31. Wang, R.X.; Wang, X.F.; Sun, Y.M. One-step synthesis of self-doped carbon dots with highly photoluminescence as multifunctional biosensors for detection of iron ions and pH. Sens. Actuator B Chem. 2017, 241, 73–79. [Google Scholar] [CrossRef]
  32. Chang, K.X.; Zhu, Q.J.; Qi, L.Y.; Guo, M.W.; Gao, W.M.; Gao, Q.W. Synthesis and properties of nitrogen-doped carbon quantum dots using lactic acid as carbon source. Materials 2022, 15, 466. [Google Scholar] [CrossRef] [PubMed]
  33. Lu, W.J.; Gong, X.J.; Nan, M.; Liu, Y.; Shuang, S.M.; Dong, C. Comparative study for N and S doped carbon dots: Synthesis, characterization and applications for Fe3+ probe and cellular imaging. Anal. Chim. Acta. 2015, 898, 116–127. [Google Scholar] [CrossRef]
  34. Nair, S.S.P.; Kottam, N.; Kumar, S.G.P. Green synthesized luminescent carbon nanodots for the sensing application of Fe3+ Ions. J. Fluoresc. 2020, 30, 357–363. [Google Scholar]
  35. Hu, Y.Y.; Ji, W.X.; Sun, J.X.; Liu, X.Y.; Zhou, R.; Yan, J.Y.; Zhang, N.N. Simple and eco-friendly synthesis of crude orange-peel-derived carbon nanoparticles for detection of Fe3+ and ascorbic acid. Luminescence 2021, 36, 1385–1394. [Google Scholar] [CrossRef]
  36. Lei, S.Y.; Chang, N.; Zhang, J.M.; Wang, H.T. Dopamine functionalized S, N co-doped carbon dots as a fluorescent sensor for the selective detection of Fe3+ and Fe2+ in water. Anal. Sci. 2021, 37, 851–857. [Google Scholar] [CrossRef]
  37. Fan, Q.; Bao, G.M.; Li, S.H.; Liu, S.Y.; Cai, X.R.; Xia, Y.F.; Li, W.; Wang, X.Y.; Deng, K.; Yuan, H.Q. A dual-channel “on-off-on” fluorescent probe for the detection and discrimination of Fe3+ and Hg2+ in piggery feed and swine wastewater. Anal. Methods 2022, 14, 2318–2328. [Google Scholar] [CrossRef]
  38. Zhou, Y.Q.; ElMetwally, A.E.; Chen, J.Y.; Shi, W.Q.; Cilingir, E.K.; Walters, B.; Mintz, K.J.; Martin, C.; Ferreira, B.; Zhang, W.; et al. Gel-like carbon dots: A high-performance future photocatalyst. J. Colloid Interface Sci. 2021, 599, 519–532. [Google Scholar] [CrossRef]
  39. Soni, N.; Singh, S.; Sharma, S.; Batra, G.; Kaushik, K.; Rao, C.; Verma, N.; Mondal, B.; Yadav, A.; Nandi, C.K. Absorption and emission of light in red emissive carbon nanodots. Chem. Sci. 2021, 12, 3615–3626. [Google Scholar] [CrossRef]
  40. Dang, T.H.T.; Mai, V.T.; Le, Q.T.; Duong, N.H.; Mai, X.D. Post-decorated surface fluorophores enhance the photoluminescence of carbon quantum dots. Chem. Phys. 2019, 527, 110503. [Google Scholar] [CrossRef]
  41. Li, H.; Zhang, Q.; Luo, X.; Yao, Z.; Qu, Y.; Wang, E.; Jiang, B.; Qiu, Z.; Li, C.; Xu, Z. Intrinsic dual emissive carbon dots for ratiometric sensing of acetylcholinesterase fluctuation induced by organophosphorus pesticide intoxication. Sens. Actuator B Chem. 2022, 373, 132590. [Google Scholar] [CrossRef]
  42. Zhu, S.; Meng, Q.; Wang, L.; Zhang, J.; Song, Y.; Jin, H.; Zhang, K.; Sun, H.; Wang, H.; Yang, B. Highly photoluminescent carbon dots for multicolor patterning, sensors, and bioimaging. Angew. Chem. Int. Edit. 2013, 125, 4045–4049. [Google Scholar] [CrossRef]
  43. Zhao, H.G.; Liu, G.J.; You, S.J.; Camargo, F.V.A.; Zavelani-Rossi, M.; Wang, X.H.; Sun, C.C.; Liu, B.; Zhang, Y.M.; Han, G.T.; et al. Gram-scale synthesis of carbon quantum dots with a large Stokes shift for the fabrication of eco-friendly and high-efficiency luminescent solar concentrators. Energy Environ. Sci. 2021, 14, 12. [Google Scholar] [CrossRef]
  44. Jiang, Y.X.; Ji, C.Y.; Wu, J.J.; Han, Q.R.; Cui, C.; Shi, W.Q.; Peng, Z.L. Formation, photoluminescence and in vitro bioimaging of polyethylene glycol-derived carbon dots: The molecular weight effects. Polymer 2022, 243, 124625. [Google Scholar] [CrossRef]
  45. Peng, J.; Gao, W.; Gupta, B.K.; Liu, Z.; Romero-Aburto, R.; Ge, L.H.; Song, L.; Alemany, L.B.; Zhan, X.B.; Gao, G.H.; et al. Graphene quantum dots derived from carbon fibers. Nano Lett. 2012, 12, 844–849. [Google Scholar] [CrossRef]
  46. Dong, X.W.; Su, Y.J.; Geng, H.J.; Li, Z.L.; Yang, C.; Li, X.L.; Zhang, Y.F. Fast one-step synthesis of N-doped carbon dots by pyrolyzing ethanolamine. J. Mater. Chem. C 2014, 2, 7477–7481. [Google Scholar] [CrossRef]
  47. Ji, C.Y.; Zhou, Y.Q.; Shi, W.Q.; Wu, J.J.; Han, Q.R.; Zhao, T.S.; Leblanc, R.M.; Peng, Z.L. Facile and sensitive detection of nitrogen-containing organic bases with near infrared C-dots derived assays. Nanomaterials 2021, 11, 2607. [Google Scholar] [CrossRef] [PubMed]
  48. Mintz, K.; Waidely, E.; Zhou, Y.Q.; Peng, Z.L.; Al-Youbi, A.O.; Bashammakh, A.S.; El-Shahawi, M.S.; Leblanc, R.M. Carbon dots and gold nanoparticles based immunoassay for detection of alpha-L-fucosidase. Anal. Chim. Acta 2018, 1041, 114–121. [Google Scholar] [CrossRef]
  49. Forootan, A.; Sjoback, R.; Bjorkman, J.; Sjogreen, B.; Linz, L.; Kubista, M. Methods to determine limit of detection and limit of quantification in quantitative real-time PCR (qPCR). Biomol. Detect. Quantif 2017, 12, 1–6. [Google Scholar] [CrossRef]
  50. Rattanopas, S.; Piyanuch, P.; Wisansin, K.; Charoenpanich, A.; Sirirak, J.; Phutdhawong, W.; Wanichacheva, N. Indole-based fluorescent sensors for selective sensing of Fe2+ and Fe3+ in aqueous buffer systems and their applications in living cells. J. Photochem. Photobiol. A Chem. 2019, 377, 138–148. [Google Scholar] [CrossRef]
  51. Sekar, A.; Yadav, R.; Basavaraj, N. Fluorescence quenching mechanism and the application of green carbon nanodots in the detection of heavy metal ions: A review. New J. Chem. 2021, 45, 2326–2360. [Google Scholar] [CrossRef]
  52. Rajendran, S.; Ramanaiah, D.V.; Kundu, S.; Bhunia, S.K. Yellow fluorescent carbon dots for selective recognition of As3+ and Fe3+ ions. ACS Appl. Nano Mater. 2021, 4, 10931–10942. [Google Scholar] [CrossRef]
  53. Rajendran, S.; Ben Zichri, S.; Vipinachandran, V.U.; Jelinek, R.; Bhunia, S.K. Triphenylphosphonium-derived bright green fluorescent carbon dots for mitochondrial targeting and rapid selective detection of tetracycline. ChemNanoMat 2021, 7, 545–552. [Google Scholar] [CrossRef]
  54. Preethi, M.; Viswanathan, C.; Ponpandian, N. Fluorescence quenching mechanism of P-doped carbon quantum dots as fluorescent sensor for Cu2+ ions. Colloids Surf. A Physicochem. Eng. Asp. 2022, 653, 129942. [Google Scholar] [CrossRef]
Figure 1. (a) TEM image of C-dots showing the high-resolution lattice fringes (insert); (b) the size distribution histogram; (c) the UV-Vis spectrum of C-dots and the dispersion of C-dots under UV-light excitation (inset, left) and under ambient light (inset, right); (d) the fluorescence emissions of C-dots (8.3 mg/L in DI water) under the excitation of lights with different wavelengths, as indicated in the figure.
Figure 1. (a) TEM image of C-dots showing the high-resolution lattice fringes (insert); (b) the size distribution histogram; (c) the UV-Vis spectrum of C-dots and the dispersion of C-dots under UV-light excitation (inset, left) and under ambient light (inset, right); (d) the fluorescence emissions of C-dots (8.3 mg/L in DI water) under the excitation of lights with different wavelengths, as indicated in the figure.
Molecules 27 06749 g001
Figure 2. (a) The FTIR spectrum of C-dots; (b) full-scale XPS spectrum of C-dots; (c) high-resolution XPS spectra of C 1s; (d) high-resolution XPS spectra of O1s; (e) high-resolution XPS spectra of N 1s; (f) XRD pattern of C-dots.
Figure 2. (a) The FTIR spectrum of C-dots; (b) full-scale XPS spectrum of C-dots; (c) high-resolution XPS spectra of C 1s; (d) high-resolution XPS spectra of O1s; (e) high-resolution XPS spectra of N 1s; (f) XRD pattern of C-dots.
Molecules 27 06749 g002
Figure 3. (a) Fluorescence emission spectra of C-dots in the presence of Fe3+ ions of various concentrations ranging from 0 to 1200 μmol/L; the emissions were excited at the optimal excitation wavelength of 360 nm and recorded at 453 nm; (b) scheme showing the correlation between F/F0 (y-axis) and concentrations of Fe3+ (x-axis) where F0 and F represented the emission intensities of the blank C-dots dispersions and C-dots dispersions with Fe3+ presented, respectively; (c) linear detection range for Fe3+ from 0 to 250 μmol/L; (d) linear detection range for Fe3+ from 250 to 1200 μmol/L).
Figure 3. (a) Fluorescence emission spectra of C-dots in the presence of Fe3+ ions of various concentrations ranging from 0 to 1200 μmol/L; the emissions were excited at the optimal excitation wavelength of 360 nm and recorded at 453 nm; (b) scheme showing the correlation between F/F0 (y-axis) and concentrations of Fe3+ (x-axis) where F0 and F represented the emission intensities of the blank C-dots dispersions and C-dots dispersions with Fe3+ presented, respectively; (c) linear detection range for Fe3+ from 0 to 250 μmol/L; (d) linear detection range for Fe3+ from 250 to 1200 μmol/L).
Molecules 27 06749 g003
Figure 4. Fluorescence emission intensity changes of C-dots dispersions in the presence of 750 μmol/L of different metal ions including Fe3+, Na+, (Fe3+ + Na+), K+, Ag+, Ba2+, Cd2+, Co2+, Ni2+, Zn2+, (Fe3+ + Zn2+), Mg2+, Cu2+, Ca2+, Fe2+, Cr3+, (Fe3+ + Cr3+) and Al3+. F0 represented the emission intensity of the blank C-dots dispersion and F represented the emission intensities of the C-dots dispersions in the presence of the respective metal ions.
Figure 4. Fluorescence emission intensity changes of C-dots dispersions in the presence of 750 μmol/L of different metal ions including Fe3+, Na+, (Fe3+ + Na+), K+, Ag+, Ba2+, Cd2+, Co2+, Ni2+, Zn2+, (Fe3+ + Zn2+), Mg2+, Cu2+, Ca2+, Fe2+, Cr3+, (Fe3+ + Cr3+) and Al3+. F0 represented the emission intensity of the blank C-dots dispersion and F represented the emission intensities of the C-dots dispersions in the presence of the respective metal ions.
Molecules 27 06749 g004
Figure 5. (a) The fluorescence decay curves of C-dots in the absence (black) and presence of Fe3+ ions (red); (b) UV-Vis absorption spectra of C-dots only (black), Fe3+ solution (red), C-dots in the presence of Fe3+ (blue) and the mathematical curve (green) resulted by subtracting the red curve (Fe3+ only) from the blue curve (C-dots + Fe3+), respectively. The negative portion of spectrum in the mathematically calculated green curve was not shown, as it has no physical meaning.
Figure 5. (a) The fluorescence decay curves of C-dots in the absence (black) and presence of Fe3+ ions (red); (b) UV-Vis absorption spectra of C-dots only (black), Fe3+ solution (red), C-dots in the presence of Fe3+ (blue) and the mathematical curve (green) resulted by subtracting the red curve (Fe3+ only) from the blue curve (C-dots + Fe3+), respectively. The negative portion of spectrum in the mathematically calculated green curve was not shown, as it has no physical meaning.
Molecules 27 06749 g005
Table 1. Sensing reliability of the assay developed as demonstrated with Fe3+ spike solutions in DI water, tap water and lake water.
Table 1. Sensing reliability of the assay developed as demonstrated with Fe3+ spike solutions in DI water, tap water and lake water.
SampleFe3+ Added (μmol/L)Fe3+ Found (μmol/L)Recovery (%)
1DI water5049.7999.58
2DI water100105.50105.50
3DI water150139.2792.85
4DI water200196.3298.16
5Tap water100100.9999.01
6Tap water150161.0692.67
7Tap water200208.1795.92
8Lake water5047.9395.86
9Lake water 100 100.9399.07
10Lake water150159.7293.52
11Lake water200198.9099.45
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Xiang, Z.; Jiang, Y.; Cui, C.; Luo, Y.; Peng, Z. Sensitive, Selective and Reliable Detection of Fe3+ in Lake Water via Carbon Dots-Based Fluorescence Assay. Molecules 2022, 27, 6749. https://doi.org/10.3390/molecules27196749

AMA Style

Xiang Z, Jiang Y, Cui C, Luo Y, Peng Z. Sensitive, Selective and Reliable Detection of Fe3+ in Lake Water via Carbon Dots-Based Fluorescence Assay. Molecules. 2022; 27(19):6749. https://doi.org/10.3390/molecules27196749

Chicago/Turabian Style

Xiang, Zhuang, Yuxiang Jiang, Chen Cui, Yuanping Luo, and Zhili Peng. 2022. "Sensitive, Selective and Reliable Detection of Fe3+ in Lake Water via Carbon Dots-Based Fluorescence Assay" Molecules 27, no. 19: 6749. https://doi.org/10.3390/molecules27196749

Article Metrics

Back to TopTop