Next Article in Journal
Pysanka-Inspired Electrode Modification with Aptamer Encapsulation in ZIF-8 for Urine Creatinine Electrochemical Biosensing
Previous Article in Journal
A Red-Emitting Fluorescence Probe for Rapid Detecting Exogenous and Endogenous Peroxynitrite in Living Cells with High Sensitivity and Selectivity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation and Application of a Fast, Naked-Eye, Highly Selective, and Highly Sensitive Fluorescent Probe of Schiff Base for Detection of Cu2+

1
Department of Pharmacy, Kangda College of Nanjing Medical University, Lianyungang 222000, China
2
School of Pharmacy, China Pharmaceutical University, Nanjing 211198, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Chemosensors 2023, 11(11), 556; https://doi.org/10.3390/chemosensors11110556
Submission received: 7 October 2023 / Revised: 31 October 2023 / Accepted: 2 November 2023 / Published: 6 November 2023
(This article belongs to the Section Applied Chemical Sensors)

Abstract

:
A fluorescent probe, N′-((3-methyl-5-oxo-1-phenyl-4, 5-dihydro-1H-pyrazol-4-yl) methylene)-2-oxo-2H-chromene-3-carbohydrazide (MPMC), was synthesized and characterized. Characterizations of the synthetic MPMC were conducted via proton nuclear magnetic resonance (1HNMR) spectroscopy and carbon-13 nuclear magnetic resonance spectroscopy (13C NMR). The fluorescence emission behaviors of the MPMC probe towards diverse metal ions were detected, and the probe exhibited high sensitivity and selectivity towards Cu2+ over other metal ions via the quenching of its fluorescence. Furthermore, the existence of other metal actions made no apparent difference to the fluorescence intensity of the MPMC-Cu2+ system; that is, MPMC displayed a good anti-interference ability. Job’s plot of the MPMC and copper ions indicated that the detection limit was 10.23 nM (R2 = 0.9612) for the assayed actions, with a stoichiometric ratio of 1:1 for MPMC and Cu2+. Additionally, the color of the MPMC probe solution changed from nearly colorless to yellow in the presence of Cu2+ in visible light, and the color change could be observed by the naked eye. Similarly, the color resolved from bright yellow into blue in ultraviolet light. Moreover, reusability studies indicated that the MPMC probe was reusable. The pH effect of the MPMC probe on Cu2+ had a broad range of pH detection, i.e., from 4.0 to 11.0. The response time of the MPMC probe for determining Cu2+ was within 1 min. The recognition of Cu2+ via MPMC performed on pre-treated paper under sunlight and UV light both had a distinct colour change. Thus, the solid-state method for detecting Cu2+ with the naked eye was both economical and convenient.

1. Introduction

Copper (II) ion is one of the essential transition metal ions that play a crucial role in many key physiological processes in living organisms, including electron transport oxidoreductases, the production of hemocyanin, blue copper protein, cytochrome C oxidase, lactase, ascorbate oxidase, superoxide dismutase, skin pigmentation, and connective tissue repair [1,2,3]. However, excessive amounts of Cu2+ can lead to a variety of diseases, such as induction of cell death, Parkinson’s, Alzheimer’s, and prion diseases [4,5,6]. In addition, the widespread use of Cu2+ in the electronic and electrical industries makes it an environmental pollutant. Especially in recent years, the operations of ore mining along with metals extraction in nature are growing, which brings about severe heavy metals pollution (copper, cadmium, etc.) in the atmosphere, water resources, and soil. These heavy metals are incapable of being degraded by organisms; instead, they can be enriched thousands of times via the biomagnification of food chain; then, they intrude into the human body, making them a serious health hazard [7,8]. Consequently, the development of a highly sensitive and selective Cu2+ assay for the efficient detection of Cu2+ in aqueous solutions or biological systems is of great importance for biology, clinical medicine, chemistry, and environmental sciences.
Nowadays, various detection technologies have been developed to test Cu2+, such as chemiluminescence [9], electro-chemistry [10], colorimetry [11], atomic absorption spectrometry [12], and inductively coupled plasma mass spectrometry [13]. However, these detection strategies have the drawbacks of low sensitivity and cumbersome operation. In recent years, fluorescent molecular probes have been rapidly developing in molecular recognition because of their excellent recognition properties, high selectivity, and sensitivity, being accessible in site detection and real-time imaging. The “host–guest” pattern is a fundamental structural model of fluorescent probes, including, chiefly, fluorophore and receptor, connected by covalent bonding through a spacer. Recognition units that possess different recognition functions serve to grasp and recognize test samples. Meanwhile, for the difference of molecular structure before and after grasping, fluorophore’s fluorescence property displays different fluorescence signals (wavelength or intensity variation), so as to achieve the purpose of recognition and detection [14,15,16,17]. Actually, most fluorescent probes that identify Cu2+ are restricted in their applications for the poor selectivity, short wavelength emission, interference from autofluorescence, the need for a high organic phase detection environment, and the presence of many possible interfering agents.
Coumarin (2H-1-benzopyran-2-one), a fluorophore used extensively, possesses high fluorescence quantum yields, good photostability, large stokes shifts, and easy structural modifications. Coumarin derivatives have been found in various plants, which are widely used in aqueous environmental monitoring, antibacterial and antitumor medications, and others [18]. Currently, coumarin-based fluorescent probes consist essentially of Schiff bases [19,20,21,22] and biological thiols [23,24]. In particular, a few studies on the coumarin-based pyrazolone fluorescent probes have been reported [23,24,25].
In the field of analytical chemistry, as a conjugated double bond, Schiff bases are recommended for the detection and identification of metal ions. Additionally, Schiff bases are applied to the quantitative analysis of certain ions by means of chromatography, fluorescence analysis, and photometric analysis. Meanwhile, for the space-turning effect of -C=N- in the molecular structure, Schiff bases and their complexes tend to coordinate easily. In addition, lone-pair electrons of the integral structure are more abundant, such that the coordination sites also become more abundant, which indicates that Schiff bases have good, coordinated properties. At this stage, Schiff bases make coordination with metallic cations easier, especially the heavy metal ions. Thus, Schiff bases are good ligands for metal ions [19,20,21,22].
Thus, a fluorescent probe based on coumarin and pyrazole Schiff base, N′-((3-methyl-5-oxo-1-phenyl-4, 5-dihydro-1H-pyrazol-4-yl) methylene)-2-oxo-2H-chromene-3-carbohy-drazide (MPMC), was designed and synthesized in this paper [26]. The coumarin moiety belonging to MPMC served as the fluorophore; and the acylhydrazone structure belonging to MPMC served as the recognition receptor and bursting part. The results demonstrate that the MPMC probe was selective and sensitive for the detection of Cu2+.

2. Experiments

2.1. Reagents and Chemicals

Phenylhydrazine (molecular weight: 108.14; relative density: 1.099 g/mL at 25 °C; and CAS number: 100-63-0), and 2-hydroxybenzaldehyde (molecular weight: 122.12; relative density: 1.146 g/mL at 25 °C; and CAS number: 90-02-8) were both analytical reagents, purchased from Macklin Inc. (Shanghai, China). Piperidine (molecular weight: 85.15; relative density: 0.862 g/mL at 20 °C; and CAS number: 110-89-4) was of analytical grade and purchased from TCI chemicals (Shanghai, China). Ethyl acetoacetate (molecular weight: 130.14; relative density: 1.029 g/mL at 20 °C; and CAS number: 141-97-9), an analytical reagent, was obtained from Aladdin Inc. (Shanghai, China). A total of 85% hydrazine hydrate (molecular weight: 50.06; relative density: 1.03 g/mL at 25 °C; and CAS number: 7803-57-8), was of analytical grade and obtained from Nanjing Reagent (Nanjing, China). Phosphoric trichloride (molecular weight: 153.33; relative density: 1.645 g/mL at 25 °C; and CAS number: 10025-87-3), analytical grade, was purchased from Kaiwei Chemical (Shanghai, China). Diethyl malonate (molecular weight: 160.17; relative density: 1.055 g/mL at 25 °C; and CAS number: 105-53-3), anhydrous ethanol (molecular weight: 46.07, relative density: 0.789 g/mL at 20 °C; and CAS number: 64-17-5), 36% acetic acid (molecular weight: 60.05; relative density: 1.05 g/mL at 25 °C; and CAS number: 64-19-7), hydrochloric acid (molecular weight: 36.46; relative density: 1.17 g/mL at 20 °C; and CAS number: 7647-01-0), N,N-dimethylformamide (molecular weight: 73.09; relative density: 0.948 g/mL at 20 °C; and CAS number: 68-12-2), sodium hydroxide (molecular weight: 40.01; relative density: 2.12 g/mL at 20 °C; and CAS number: 1310-73-2), cupric chloride anhydrous (molecular weight: 134.45; relative density: 3.386 g/mL at 25 °C; and CAS number: 7447-39-4), zinc chloride (molecular weight: 136.30; relative density: 1.01 g/mL at 20 °C; and CAS number: 7646-85-7), manganese(II) chloride (molecular weight: 125.84; relative density: 2.98 g/mL at 25 °C; and CAS number: 7773-01-5), nickel(II) sulfate (molecular weight: 154.76; relative density: 3.68 g/mL at 25 °C; and CAS number: 7786-81-4), iron(III) chloride (molecular weight: 162.20; relative density: 2.84 g/mL at 25 °C; and CAS number: 7705-08-0), ferrous chloride (molecular weight: 126.75; relative density: 3.16 g/mL at 25 °C; and CAS number: 7758-94-3), calcium chloride (molecular weight: 110.98; relative density: 1.086 g/mL at 25 °C; and CAS number: 10043-52-4), magnesium sulfate (molecular weight: 120.37; relative density: 1.07 g/mL at 25 °C; and CAS number: 7487-88-9), lead chloride (molecular weight: 278.11; relative density: 5.85 g/mL at 25 °C; and CAS number: 7758-95-4), sodium chloride (molecular weight: 58.44; relative density: 2.165 g/mL at 25 °C; and CAS number: 7647-14-5), potassium chloride (molecular weight: 74.55; relative density: 1.98 g/mL at 25 °C; and CAS number: 7447-40-7), silver nitrate (molecular weight: 169.87; relative density: 4.35 g/mL at 25 °C; and CAS number: 7761-88-8), and barium chloride (molecular weight: 208.23; relative density: 3.856 g/mL at 25 °C; and CAS number: 10361-37-2) were all of analytical grade, and obtained from Sinopharm (Shanghai, China). A stock solution of MPMC was prepared in EtOH at a concentration of 1.0 × 10−3 mol/L. The stock perchlorate solutions (including the perchlorate of Cu2+, Pb2+, Fe3+, Fe2+, Mg2+, Ni2+, Cd2+, Co2+, Mn2+, Ca2+, Ba2+, Na+, K+, Zn2+ and Ag+) were freshly prepared in deionized water at a concentration of 1.0 × 10−3 mol/L.
1H and 13C NMR spectra were recorded on a Bruker AVANCE III 400 MHz NMR spectrometer in DMSO-d6 with TMS as an internal standard. Mass spectrum was recorded on the Thermo Q-Exactive mass spectrometer. The melting point was measured on the XRC-1 melting point instrument. The ultraviolet absorption was recorded on Cary50. The fluorescence test was recorded on the RF-6000 luminescence spectrophotometer. ESI-MS data were recorded with Mariner System 5304 mass spectrometer. Elemental analyses (C, H, and N) were gathered on a CHN-O-Rapid instrument within 0.4% of the theoretical values.

2.2. Synthesis of MPMC

The synthetic route of the target MPMC probe is shown in Scheme 1. Edaravone aldehyde derivative (compound 2) and coumarin-3-carbohydrazide (compound 4) were synthesized in reference to the methods reported in the literature [27,28,29,30].

2.2.1. Synthesis of Edaravone

A total of 13.0 g and 0.1 mol of acetyl ethyl acetate was first mixed with 5 mL of 60% ethanol solution while stirring. Next, a solution consisting of phenylhydrazine (10.8 g, 0.1 mol) and absolute ethanol was added dropwise at 45 °C in approximately 30 min. After dripping, the reaction proceeded at 45 °C for 20 min. Thereafter, the reaction solution was cooled to room temperature, supplemented with 1 mL of concentrated hydrochloric acid, reacting at 45 °C for another 2 h. The following is the addition of 10% sodium hydroxide solution, dropwise, to adjust the pH to be 7.0. Then, the solution was added by 20 mL of water and stirred at room temperature for 1 h, followed by filtering to obtain light yellow crystals. Finally, crystals were washed with cold absolute ethanol and heated to reflux in ethyl acetate and absolute ethanol solution (with a volume ratio of 2:1) to obtain white and pure edaravone (compound 1). The synthetic route of edaravone is shown in Scheme 2.

2.2.2. Synthesis of Edaravone Aldehyde Derivative

Edaravone (8.70 g, 0.05 mol) was dissolved in 1 mL of N,N-dimethylformamide under stirring; then, 10.1 mL of phosphorus oxychloride was introduced slowly in drops by an external ice bath. The reaction mixture was then heated to reflux for 1.5 h. After cooling, it was introduced into 200 mL of ice water, then oscillated, stood still, and filtered; thus, crude edaravone aldehyde derivative was obtained. Subsequently, the crude product was treated by absolute ethanol washing, heated to reflux in ethanol, followed by drying to gain the pure edaravone aldehyde derivative (compound 2). The synthetic route of edaravone aldehyde derivative is shown in Scheme 3.

2.2.3. Synthesis of Ethyl Coumarin-3-Carboxylate

Salicylaldehyde (4.90 g, 0.04 mol), diethyl malonate (7.25 g, 0.045 mol), and 20 mL of absolute ethanol were successively added and thoroughly mixed, and then 0.5 mL of hexahydropyridine and two drops of acetic acid were dripped. In succession, the reaction mixture was then heated to reflux for 2 h while stirring. After being cooled to dilution crystallization, filtering was adopted to obtain the crude product, which was then washed with ethanol several times and dried to obtain the pure ethyl coumarin-3-carboxylate (compound 3). The synthetic route of ethyl coumarin-3-carboxylate is shown in Scheme 4.

2.2.4. Synthesis of Coumarin-3-Carbohydrazide

In the presence of coumarin-3-carboxylic acid ethyl ester (2.00 g, 0.01 mol) and 50 mL of anhydrous ethanol, five drops of hydrazine hydrate (0.01 mol) were then added. The reaction mixture was heated to reflux for 10 h under agitation. Then, crystals of coumarin-3-carbohydrazide were formed while being cold. After filtering, yellow crystals were obtained. The yellow crystals were washed by absolute ethanol and dissolved with reflux in absolute ethanol. Finally, bright yellow crystals, which were the pure coumarin-3-carbohydrazide (compound 4), were produced via the dryness. The synthetic route of coumarin-3-carbohydrazide is shown in Scheme 5.

2.2.5. Synthesis of MPMC Probe

Compound 2 (1.0 g, 4.95 mmol) and compound 4 (1.0 g, 4.95 mmol) were dissolved in 40 mL of ethanol. With the addition of catalyst acetic acid, the reaction proceeded at reacting temperature of 78 °C, heating to reflux for 2 h. After restoration at room temperature, a light-yellow precipitate was generated and was obtained via filtering, washing with ethanol several times. Finally, the precipitate was dried to give the MPMC probe (0.902 g, yield 46.96%). m.p. > 280 °C. 1H NMR and 13C NMR data of MPMC are shown as follows: 1H NMR (400 MHz, DMSO-d6) δ 11.14 (s, 1H), 9.01 (s, 1H), 8.77 (s, 1H), 7.93 (d, J = 7.6 Hz, 1H), 7.76 (d, J = 8.0 Hz, 1H), 7.73–7.68 (m, 1H), 7.45 (d, J = 8.6 Hz, 1H), 7.44–7.37 (m, 2H), 7.00 (s, 1H), 6.97 (d, J = 7.4 Hz, 1H), 4.30 (q, J = 7.0 Hz, 2H), 1.32 (t, J = 7.1 Hz, 3H). 13C NMR (100 MHz, DMSO-d6) δ 163.26, 163.07, 159.11, 156.48, 155.01, 149.19, 134.97, 133.72, 131.28, 130.76, 125.33, 120.08, 118.66, 118.29, 118.14, 117.00, 116.63, 61.71, 14.54. MS (ESI): 389.38 (M + H) +. Anal.Calcd for C21H16N4O4: C, 64.94; H, 4.15; N, 14.43. Found: C, 64.92; H, 4.13; N, 14.41.

2.3. General Procedure for the Spectrum Measurement

Stock solutions of MPMC (1 mM) were prepared in ethanol. The metal ions stock solutions (1 mM) were prepared with the nitrate or chloride salts (Cu2+, Pb2+, Fe3+, Fe2+, Mg2+, Ni2+, Cd2+, Co2+, Mn2+, Ca2+, Ba2+, Na+, K+, Zn2+ and Ag+) in deionized water. Absorption and emission spectra were obtained at room temperature using PBS solution (pH 7.5) in MPMC (10 μM) with different concentrations of analyzer. Fluorescence spectra were recorded at an excitation wavelength of 358 nm. The total concentration of Cu2+ and MPMC was kept constantly (2.0 mM). The fluorescence intensity of MPMC was then recorded by varying the molar ratio of MPMC to Cu2+. The selectivity of MPMC towards Cu2+ was tested by comparing other metal ions. MPMC (1 mM) was treated with Cu2+ (1 mM) and other metal ions (10 mM) for 10 min and the fluorescence intensity of the mixtures was recorded. For reproducibility testing, Cu2+ (1.0 mM) was incubated with MPMC aqueous solution (1.0 mM), and the fluorescence of MPMC was quenched. The fluorescence intensity of MPMC (1.0 mM) and MPMC-Cu2+ ensemble (1.0 mM) was determined in a series of buffers (pH 1.0 to 14.0).

3. Results and Discussion

3.1. Fluorescence Spectra of the MPMC Probe for Selectivity and Anti-Interference Detection

The fluorescence emission behaviors of the MPMC probe towards diverse metal ions were detected in ethanol (EtOH)/H2O (v/v = 1/1) solution, respectively. Fluorescence intensities of diverse metal ions added to MPMC were detected as the wavelength variation. The fluorescence emission spectrums are illustrated in Figure 1a. As shown in Figure 1a, when excitation wavelength 358 nm was given to the MPMC probe, a fluorescence emission band at 548 nm generated. Then fluorescence quenching followed along with the insertion of Cu2+ (1.0 equiv), with the emission minimum being at 548 nm. Conversely, the fluorescence emission spectrums of MPMC were affected little by introducing other metal ions, in addition to a weak quenching of nickel ions.
Furthermore, in view of potential interference of other metal ions in practical applications, sorts of metal ions were added successively into the MPMC-Cu2+ system, and the influences on fluorescence selectivity were investigated.
A total of 100 μM of competitive metal ions (Pb2+, Fe3+, Fe2+, Mg2+, Ni2+, Cd2+, Co2+, Mn2+, Ca2+, Ba2+, Na+, K+, Zn2+, Ag+), along with 10 μM MPMC in EtOH-H2O (v/v = 1/1), were prepared and determined to access the fluorescence intensities. As Figure 1b illustrates, a fluorescence emission band at 548 nm was generated again. Thereafter, with the addition of 10 μM Cu2+, the fluorescence quenching took place. The existence of other metal actions made no difference on the fluorescence intensity apparently. In general, the MPMC probe has the characteristic of selectively recognizing Cu2+, displaying a good anti-interference ability.

3.2. Titration Experiment of the MPMC Probe to Cu2+

Colorimetric experiments were conducted to study the specificity of the MPMC probe towards Cu2+. As Figure 2a shows, while adding Cu2+ to MPMC dissolved in EtOH/H2O (v/v = 1/1), the color of solution passed from colorless to yellow in seconds. The result shows that the MPMC probe can realize the colorimetric detection of Cu2+ with a detection limit 10 μM. In the same way, the fluorescence color turned from colorless to yellow under 365 nm ultraviolet (UV) light, shown in Figure 2b. The applied results clearly demonstrate that the MPMC probe could have an application in the qualitative and quantitative detection of Cu2+ in the form of color change and spectrum signals multiplication.
As the concentration variation of Cu2+, fluorescence titration experiments were performed to explore the sensitivity of the MPMC probe to Cu2+. As illustrated in Figure 3, with the Cu2+ concentration rising, the fluorescence intensity of the MPMC probe reflected a continued decrease. The linear equation for copper (II) ions is Y = −728.21X + 1388.2, and the correlated coefficient is 0.9612. Based on the formula L = 3σ/Ka, the minimum detectable concentration of the probe for copper (II) ions was calculated as 10.23 nM [31]. Compared with other fluorescent probes for the detection of Cu2+ listed in Table 1, the MPMC probe prepared in this paper has a lower detection limit.

3.3. Study of EDTA Effect of MPMC Probe on Cu2+

To fully investigate the response between MPMC probe and Cu2+, the procedures involved the addition of ethylene diamine tetraacetate acid (EDTA) to MPMC-Cu2+ complex, and the detection of fluorescence intensity was performed. As shown in Figure 4, the fluorescence intensity was recovered with the additional amount of EDTA added into the MPMC-Cu2+ complex. It suggests that the complexation ability of EDTA to Cu2+ is stronger than that of the MPMC probe. EDTA seized the Cu2+ bound to the MPMC probe, leading to the recovery of fluorescence. Namely, the fluorescence recognition between MPMC probe and Cu2+ is reversible [41]. Then, the fluorescence quenching still occurred after the sustainable addition of Cu2+. Similarly, the fluorescence intensity recovered again after the addition of EDTA. The results indicate that the MPMC probe is reusable.

3.4. Study of pH Effect of MPMC Probe to Cu2+

The MPMC probe is required not only for highly sensitive and selective performance, but also for its good sensing ability at different acidities in practical applications. The sensing ability was detected by adjusting the pH of EtOH-PBS (phosphate buffer solution) from 1.0 to 14.0. As Figure 5 illustrates, with a range 4.0–11.0 of pH variation, the fluorescence intensity of the MPMC probe was maintained constantly. It then dropped dramatically in a range 5.0–8.0 pH while adding Cu2+ to the MPMC solution. At a low pH range, the fluorescence intensity exhibited no variance; it was probably caused by hydrolysis of Cu2+ under acidic conditions, which inhibited the formation of the MPMC-Cu2+ complex30. Thus, the prepared MPMC probe acted as a fluorescent pH sensor, suitable for a broad range of pH detection (4.0 to 11.0), especially from 5.0 to 8.0.

3.5. Study of the Response Time of MPMC Probe to Cu2+

The response time of the MPMC probe in determining copper (II) ions was obtained. The variation of the fluorescence intensity at 548 nm with the reaction time of the MPMC probe along with Cu2+ is shown in Figure 6. After the addition of copper (II) ions, the fluorescence intensity of the MPMC probe declined markedly within 1 min then achieved an equilibrium after 2 min.

3.6. Contact Mode Detection between MPMC Probe and Cu2+

To explore the binding ratio of the MPMC probe to Cu2+, different ratios of the MPMC probe to Cu2+ (1:9, 2:8, 3:7, 4:6, 5:5, 6:4, 7:3, 8:2, 9:1) were prepared. As the molar fraction changed, the fluorescence characteristics were analyzed, and Job’s plot curve is illustrated in Figure 7. With the molar fraction varying from 1:9 to 5:5, the fluorescence intensity decreased first, while the fluorescence intensity was almost unchanged in the range from 5:5 to 9:1. The molar fraction corresponding to the minimum fluorescence intensity of the MPMC probe appeared at 0.5, indicating that MPMC-Cu2+ complex was formed by 1:1. The possible binding mechanism of MPMC to Cu2+-induced fluorescence changes is shown in Figure 8.

3.7. The Test Paper for Cu2+ Ions

To study the multifunction applications of the solid-state MPMC probe for its high efficiency and simplicity, the recognition of Cu2+ by MPMC was performed on pre-treated paper. Filter papers were steeped in MPMC probe dissolved in an ethanol-saturated solution (1.0 × 10−3 mol/L) for a few seconds to obtain test strips. Afterward, solid-state experiments were carried out by drying test strips in air and treating with an aqueous solution of Cu2+ (1.0 × 10−3 mol/L). Under sunlight, test strips before and after solid-state experiments presented white and yellow colors, respectively, as seen in Figure 9. In the same way, the color became light yellow and blue under 365 nm UV light, respectively [42]. Thus, the solid-state method for detecting Cu2+ with the naked eye was both economical and convenient.

4. Conclusions

A fluorescent probe with highly selective and sensitive performance based on coumarin and pyrazole Schiff base has been synthesized. The fluorescence emission behaviors of the MPMC probe towards diverse metal ions were detected, and the probe exhibited high sensitivity and selectivity towards Cu2+ over other metal ions via the quenching of its fluorescence. Furthermore, the existence of other metal actions made no apparent difference to the fluorescence intensity of the MPMC-Cu2+ system; that is, MPMC displayed a good anti-interference ability. Job’s plot of MPMC and copper ions indicated that the detection limit was 10.23 nM (R2 = 0.9612) for the assayed actions, with a stoichiometric ratio of 1:1 for MPMC and Cu2+. Additionally, the color of the MPMC probe solution changed from nearly colorless to yellow in the presence of Cu2+ in visible light, and the color change could be observed by the naked eye. Similarly, the color resolved from bright yellow into blue in ultraviolet light. Moreover, the MPMC probe was reusable. The pH effect of the MPMC probe on Cu2+ had a broad range of pH detection (4.0 to 11.0, especially from 5.0 to 8.0). The response time of the MPMC probe for determining Cu2+ was within 1 min. The recognition of Cu2+ by MPMC performed on pre-treated paper under sunlight and UV light both had a distinct colour change. In conclusion, the MPMC probe could be applied to recognize Cu2+ in the environment; moreover, applications in biological systems may be achieved in future.

Author Contributions

Conceptualization, H.-L.L., J.L., P.-Y.C. and X.W.; Data curation, H.-L.L., J.L., P.-Y.C., X.W., M.W., K.W., Y.L. and Y.-Y.C.; Formal Analysis, P.-Y.C., S.C., M.W., K.W., Y.L. and Y.-Y.C.; Funding acquisition, H.-L.L., J.L., S.C. and M.W.; Investigation, J.L., P.-Y.C., S.C., M.W., K.W., Y.L. and Y.-Y.C.; Methodology, H.-L.L. and P.-Y.C.; Project administration, H.-L.L., J.L., P.-Y.C. and X.W.; Resources, H.-L.L., J.L. and X.W.; Supervision, H.-L.L. and X.W.; Validation, J.L., P.-Y.C., S.C., M.W., K.W., Y.L. and Y.-Y.C.; Visualization, H.-L.L., J.L., P.-Y.C., X.W., S.C., M.W., K.W., Y.L. and Y.-Y.C.; Writing—original draft, H.-L.L., J.L., P.-Y.C. and X.W.; Writing—review and editing, H.-L.L., J.L., P.-Y.C., X.W., S.C., M.W., K.W., Y.L. and Y.-Y.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Scientific Research Personnel Training Program of Kangda College of Nanjing Medical University (No. KD2022KYRC007, Kangda College of Nanjing Medical University), the Science and Technology Funds of Nanjing Medical University (No. NMUB20210365, Nanjing Medical University), the Social Development Funds of Lianyungang (No. SF2107, Lianyungang Science and Technology Bureau), the Science and Technology Funds of Kanada College of Nanjing Medical University (No. KD2021KYJJZD009 and KD2022KYJJZD002, Kanada College of Nanjing Medical University), and the Education Research Project of Kangda College of Nanjing Medical University (KD2021JYYJYB010, Kanada College of Nanjing Medical University).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author. The data are not publicly available due to restrictions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Peers, G.; Price, N.M. Copper-containing plastocyanin used for electron transport by an oceanic diatom. Nature 2006, 441, 341–344. [Google Scholar] [CrossRef] [PubMed]
  2. Uauy, R.; Olivares, M.; Gonzalez, M. Essentiality of copper in humans. Am. J. Clin. Nutr. 1998, 67, 952S–959S. [Google Scholar] [CrossRef] [PubMed]
  3. Linder, M.C.; Hazegh-Azam, M. Copper biochemistry and molecular biology. Am. J. Clin. Nutr. 1996, 63, 797S–811S. [Google Scholar] [CrossRef] [PubMed]
  4. Gaggelli, E.; Kozlowski, H.; Valensin, D.; Valensin, G. Copper homeostasis and neurodegenerative disorders (Alzheimer’s, prion, and Parkinson’s diseases and amyotrophic lateral sclerosis). Chem. Rev. 2006, 106, 1995–2044. [Google Scholar] [CrossRef] [PubMed]
  5. Strausak, D.; Mercer, J.F.B.; Dieter, H.H.; Stremmel, W.; Multhaup, G. Copper in disorders with neurological symptoms: Alzheimer’s, Menkes, and Wilson diseases. Brain Res. Bull. 2001, 55, 175–185. [Google Scholar] [CrossRef]
  6. Tsvetkov, P.; Coy, S.; Petrova, B.; Dreishpoon, M.; Verma, A.; Abdusamad, M.; Rossen, J.; Joesch-Cohen, L.; Humeidi, R.; Spangler, R.D.; et al. Copper induces cell death by targeting lipoylated TCA cycle proteins. Science 2022, 375, 1254–1261. [Google Scholar] [CrossRef]
  7. Sun, J.Z.; Xu, X.; Yu, G.H.; Li, W.N.; Shi, J.S. Coumarin-based tripodal chemosensor for selective detection of Cu(II) ion and resultant complex as anion probe through a Cu(II) displacement approach. Tetrahedron 2018, 74, 987–991. [Google Scholar] [CrossRef]
  8. Meng, X.J.; Cao, D.L.; Hu, Z.Y.; Han, X.H.; Li, Z.C.; Liang, D.; Ma, W.B. A coumarin based highly selective fluorescent chemosensor for sequential recognition of Cu2+ and PPi. Tetrahedron Lett. 2018, 59, 4299–4304. [Google Scholar] [CrossRef]
  9. Amjadi, M.; Abolghasemi-Fakhri, Z. Gold nanostar-enhanced chemiluminescence probe for highly sensitive detection of Cu(II) ions. Sens. Actuators B Chem. 2018, 257, 629–634. [Google Scholar] [CrossRef]
  10. Yu, Y.Y.; Yu, C.; Yin, T.X.; Ou, S.S.; Sun, X.Y.; Wen, X.R.; Zhang, L.; Tang, D.Q.; Yin, X.X. Functionalized poly (ionic liquid) as the support to construct a ratiometric electrochemical biosensor for the selective determination of copper ions in AD rats. Biosens. Bioelectron. 2017, 87, 278–284. [Google Scholar] [CrossRef]
  11. Al-Nidawi, M.; Alshana, U. Reversed-phase switchable-hydrophilicity solvent liquid-liquid microextraction of copper prior to its determination by smartphone digital image colorimetry. J. Food Compos. Anal. 2021, 104, 104140. [Google Scholar] [CrossRef]
  12. Gamela, R.R.; Costa, V.C.; Pereira, E.R. Multivariate Optimization of Ultrasound-Assisted Extraction Procedure for the Determination of Ca, Fe, K, Mg, Mn, P, and Zn in Pepper Samples by ICP OES. Food Anal. Method. 2020, 13, 69–77. [Google Scholar] [CrossRef]
  13. Bonin, M.; Lariviere, D.; Povinec, P.P. Detection of radium at the attogram per gram level in copper by inductively coupled plasma mass spectrometry after cation-exchange chromatography. Anal. Methods 2020, 12, 2272–2278. [Google Scholar] [CrossRef]
  14. Zhou, Z.L.; Tang, H.H.; Chen, S.Y.; Huang, Y.H.; Zhu, X.H.; Li, H.T.; Zhang, Y.Y.; Yao, S.Z. A turn-on red-emitting fluorescent probe for determination of copper(II) ions in food samples and living zebrafish. Food Chem. 2021, 343, 128513. [Google Scholar] [CrossRef]
  15. Zhang, Y.D.; Cai, Y.H.; He, Y.H.; Lin, Q.L.; Ren, J.L.; Cao, D.S.; Zhang, L. A label-free fluorescent peptide probe for sensitive and selective determination of copper and sulfide ions in aqueous systems. RSC Adv. 2021, 11, 7426–7435. [Google Scholar] [CrossRef]
  16. Zeng, X.D.; Gao, S.; Jiang, C.; Duan, Q.X.; Ma, M.S.; Liu, Z.G.; Chen, J. Rhodol-derived turn-on fluorescent probe for copper ions with high selectivity and sensitivity. Luminescence 2021, 36, 1761–1766. [Google Scholar] [CrossRef]
  17. Liu, J.Q.; Liu, Z.C.; Wang, W.K.; Tian, Y. Real-time Tracking and Sensing of Cu+ and Cu2+ with a Single SERS Probe in the Live Brain: Toward Understanding Why Copper Ions Were Increased upon Ischemia. Angew. Chem. Int. Ed. 2021, 60, 21351–21359. [Google Scholar] [CrossRef]
  18. Sun, X.Y.; Liu, T.; Sun, J.; Wang, X.J. Synthesis and application of coumarin fluorescence probes. RSC Adv. 2020, 10, 10826–10847. [Google Scholar] [CrossRef]
  19. Wang, K.N.; Sun, P.Z.; Chao, X.J.; Cao, D.X.; Mao, Z.W.; Liu, Z.Q. A coumarin Schiff’s base two-photon fluorescent probe for hypochlorite in living cells and zebrafish. RSC Adv. 2018, 8, 6904–6909. [Google Scholar] [CrossRef]
  20. He, G.J.; Hua, X.B.; Yang, N.; Li, L.L.; Xu, J.H.; Yang, L.L.; Wang, Q.Z.; Ji, L.G. Synthesis and application of a “turn on” fluorescent probe for glutathione based on a copper complex of coumarin hydrazide Schiff base derivative. Bioorg. Chem. 2019, 91, 103176. [Google Scholar] [CrossRef]
  21. Yan, L.Q.; Hu, C.J.; Li, J.P. A fluorescence turn-on probe for rapid monitoring of hypochlorite based on coumarin Schiff base. Anal. Bioanal. Chem. 2018, 410, 7457–7464. [Google Scholar] [CrossRef]
  22. Wang, Y.; Hao, X.H.; Liang, L.X.; Gao, L.Y.; Ren, X.M.; Wu, Y.G.; Zhao, H.C. A coumarin-containing Schiff base fluorescent probe with AIE effect for the copper(ii) ion. RSC Adv. 2020, 10, 6109–6113. [Google Scholar] [CrossRef]
  23. Mani, K.S.; Rajamanikandan, R.; Ravikumar, G.; Pandiyan, B.V.; Kolandaivel, P.; Ilanchelian, M.; Rajendran, S.P. Highly Sensitive Coumarin-Pyrazolone Probe for the Detection of Cr3+ and the Application in Living Cells. ACS Omega 2018, 3, 17212–17219. [Google Scholar] [CrossRef]
  24. Babur, B.; Seferoglu, N.; Seferoglu, Z. A coumarin-pyrazolone based fluorescent probe for selective colorimetric and fluorimetric fluoride detection: Synthesis, spectroscopic properties and DFT calculations. J. Mol. Struct. 2018, 1161, 218–225. [Google Scholar] [CrossRef]
  25. Babur, B.; Seferoglu, N.; Ocal, M.; Sonugur, G.; Akbulut, H.; Seferoglu, Z. A novel fluorescence turn-on coumarin-pyrazolone based monomethine probe for biothiol detection. Tetrahedron 2016, 72, 4498–4502. [Google Scholar] [CrossRef]
  26. Honglei, L.; Pengyu, C.; Juan, L.; Sai, C.; Meng, W. Coumarin Pyrazolone Fluorescent Probe as Well as Preparation Method and Application Thereof. CN202210882392.3[P]. 11 November 2022. Available online: https://d.wanfangdata.com.cn/patent/ChJQYXRlbnROZXdTMjAyMzA5MDESEENOMjAyMjEwODgyMzkyLjMaCGxzemJpOXd5 (accessed on 1 November 2023).
  27. Vijay, K.; Nandi, C.; Samant, S.D. Synthesis of a dihydroquinoline based merocyanine as a ‘naked eye’ and ‘fluorogenic’ sensor for hydrazine hydrate in aqueous medium and hydrazine gas. RSC Adv. 2014, 4, 30712–30717. [Google Scholar] [CrossRef]
  28. Surati, K.R.; Thaker, B.T. Synthesis, spectroscopic and thermal investigation of Schiff-base complexes of Cu(II) derived from heterocyclic beta-diketone with various primary amines. J. Coord. Chem. 2006, 59, 1191–1202. [Google Scholar] [CrossRef]
  29. Naik, M.D.; Bodke, Y.D.; Kumar, V.M.; Revanasiddappa, B.C. An efficient one-pot synthesis of coumarin-amino acid derivatives as potential anti-inflammatory and antioxidant agents. Synth. Commun. 2020, 50, 1210–1216. [Google Scholar] [CrossRef]
  30. Taha, M.; Shah, S.A.A.; Afifi, M.; Imran, S.; Sultan, S.; Rahim, F.; Khan, K.M. Synthesis, alpha-glucosidase inhibition and molecular docking study of coumarin based derivatives. Bioorg. Chem. 2018, 77, 586–592. [Google Scholar] [CrossRef]
  31. Xia, X.C.; Fu, Y.; Tang, H.; Li, Y.; Wang, F.Y.; Ren, J. A self-assembled micellar nanoprobe for specific recognition of hydrazine in vitro and in vivo. Sens. Actuators B Chem. 2018, 272, 479–484. [Google Scholar] [CrossRef]
  32. Yin, J.; Wang, Z.; Zhao, F.; Yang, H.; Li, M.; Yang, Y. A novel dual functional pyrene-based turn-on fluorescent probe for hypochlorite and copper (II) ion detection and bioimaging applications. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2020, 239, 118470. [Google Scholar] [CrossRef] [PubMed]
  33. Wang, Y.Y.; Ai, Y.; Zhang, Y.L.; Ren, Y.Y.; Wang, J.L.; Yao, F.Y.; Li, W.Y.; Zhou, Y.; Sun, Y.N.; Liu, J.L.; et al. A new probe with high selectivity and sensitivity for detecting copper ions in traditional Chinese medicine and water sample. Inorg. Chem. Commun. 2021, 128, 108563. [Google Scholar] [CrossRef]
  34. Wang, M.M.; Wang, C.; Wang, M.; Sun, T.M.; Huang, Y.; Tang, Y.F.; Ju, J.F.; Shen, L.J.; Hu, Y.Y.; Zhu, J.L. A Dual-Functional “On-Off-On” Relay Fluorescent Probe for the Highly Sensitive Detection of Copper(II) and Phosphate Ions. ChemistrySelect 2020, 5, 1331–1334. [Google Scholar] [CrossRef]
  35. Anand, T.; Sivaraman, G.; Chellappa, D. Quinazoline copper(II) ensemble as turn-on fluorescence sensor for cysteine and chemodosimeter for NO. J. Photochem. Photobiol. A 2014, 281, 47–52. [Google Scholar] [CrossRef]
  36. Arslan, F.N.; Geyik, G.A.; Koran, K.; Ozen, F.; Aydin, D.; Elmas, S.N.K.; Gorgulu, A.O.; Yilmaz, I. Fluorescence “Turn On-Off” Sensing of Copper (II) Ions Utilizing Coumarin-Based Chemosensor: Experimental Study, Theoretical Calculation, Mineral and Drinking Water Analysis. J. Fluoresc. 2020, 30, 317–327. [Google Scholar] [CrossRef] [PubMed]
  37. Kim, G.; Choi, D.; Kim, C. A Benzothiazole-Based Fluorescence Turn-on Sensor for Copper(II). J. Fluoresc. 2021, 31, 1203–1209. [Google Scholar] [CrossRef] [PubMed]
  38. Silpcharu, K.; Soonthonhut, S.; Sukwattanasinitt, M.; Rashatasakhon, P. Fluorescent Sensor for Copper(II) and Cyanide Ions via the Complexation-Decomplexation Mechanism with Di(bissulfonamido)spirobifluorene. ACS Omega 2021, 6, 16696–16703. [Google Scholar] [CrossRef]
  39. Tamil Selvan, G.; Varadaraju, C.; Tamil Selvan, R.; Enoch, I.; Mosae Selvakumar, P. On/Off Fluorescent Chemosensor for Selective Detection of Divalent Iron and Copper Ions: Molecular Logic Operation and Protein Binding. ACS Omega 2018, 3, 7985–7992. [Google Scholar] [CrossRef]
  40. Rupa, S.A.; Patwary, M.A.M.; Ghann, W.E.; Abdullahi, A.; Uddin, A.K.M.R.; Mahmud, M.M.; Haque, M.A.; Uddin, J.; Kazi, M. Synthesis of a novel hydrazone-based compound applied as a fluorescence turn-on chemosensor for iron(iii) and a colorimetric sensor for copper(ii) with antimicrobial, DFT and molecular docking studies. RSC Adv. 2023, 13, 23819–23828. [Google Scholar] [CrossRef]
  41. Zhang, Y.P.; Teng, Q.; Yang, Y.S.; Guo, H.C.; Xue, J.J. A novel coumarin-based pyrazoline fluorescent probe for detection of Fe3+ and its application in cells. Inorg. Chim. Acta 2021, 525, 120469. [Google Scholar] [CrossRef]
  42. Zong, L.Y.; Wang, C.; Song, Y.C.; Hu, J.; Li, Q.Q.; Li, Z. A fluorescent and colorimetric probe based on naphthalene diimide and its high sensitivity towards copper ions when used as test strips. RSC Adv. 2019, 9, 12675–12680. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Reagents and conditions: (i) ethyl acetoacetate, 60% ethanol solution, 45 °C, 2 h; (ii) N,N-dimethylformamide, POCl3, 100 °C, 15 h; (iii) diethyl malonate, piperidine, ethanol, 25 °C, overnight; (iv) ethanol, 85% hydrazine hydrate, 20 h; (v) ethanol, CH3COOH, reflux, 2 h.
Scheme 1. Reagents and conditions: (i) ethyl acetoacetate, 60% ethanol solution, 45 °C, 2 h; (ii) N,N-dimethylformamide, POCl3, 100 °C, 15 h; (iii) diethyl malonate, piperidine, ethanol, 25 °C, overnight; (iv) ethanol, 85% hydrazine hydrate, 20 h; (v) ethanol, CH3COOH, reflux, 2 h.
Chemosensors 11 00556 sch001
Scheme 2. Synthesis of edaravone.
Scheme 2. Synthesis of edaravone.
Chemosensors 11 00556 sch002
Scheme 3. Synthesis of edaravone aldehyde derivative.
Scheme 3. Synthesis of edaravone aldehyde derivative.
Chemosensors 11 00556 sch003
Scheme 4. Synthesis of ethyl coumarin-3-carboxylate.
Scheme 4. Synthesis of ethyl coumarin-3-carboxylate.
Chemosensors 11 00556 sch004
Scheme 5. Synthesis of coumarin-3-carbohydrazide.
Scheme 5. Synthesis of coumarin-3-carbohydrazide.
Chemosensors 11 00556 sch005
Figure 1. (a) Fluorescence emission spectra of the MPMC probe (10 μM) without or with metal ions (including Cu2+, Pb2+, Fe3+, Fe2+, Mg2+, Ni2+, Cd2+, Co2+, Mn2+, Ca2+, Ba2+, Na+, K+, Zn2+, Ag+) (10 μM) in ethanol with excitation wave length of 358 nm. (b) Fluorescence response of MPMC (10 μM) to Cu2+ in the presence of various metal ions (including Pb2+, Fe3+, Fe2+, Mg2+, Ni2+, Cd2+, Co2+, Mn2+, Ca2+, Ba2+, Na+, K+, Zn2+, Ag+) (100 μM) in EtOH/H2O (v/v = 1/1) (λex = 358 nm, λem = 548 nm).
Figure 1. (a) Fluorescence emission spectra of the MPMC probe (10 μM) without or with metal ions (including Cu2+, Pb2+, Fe3+, Fe2+, Mg2+, Ni2+, Cd2+, Co2+, Mn2+, Ca2+, Ba2+, Na+, K+, Zn2+, Ag+) (10 μM) in ethanol with excitation wave length of 358 nm. (b) Fluorescence response of MPMC (10 μM) to Cu2+ in the presence of various metal ions (including Pb2+, Fe3+, Fe2+, Mg2+, Ni2+, Cd2+, Co2+, Mn2+, Ca2+, Ba2+, Na+, K+, Zn2+, Ag+) (100 μM) in EtOH/H2O (v/v = 1/1) (λex = 358 nm, λem = 548 nm).
Chemosensors 11 00556 g001
Figure 2. (a) Colorimetric performance of sensor MPMC (1 mM) upon addition of different metal ions (including Cu2+, Pb2+, Fe3+, Fe2+, Mg2+, Ni2+, Cd2+, Co2+, Mn2+, Ca2+, Ba2+, Na+, K+, Ag+) (1 mM) in EtOH/H2O (v/v = 1/1) solution; (b) color change induced upon addition of Cu2+ under 365 nm UV lamp.
Figure 2. (a) Colorimetric performance of sensor MPMC (1 mM) upon addition of different metal ions (including Cu2+, Pb2+, Fe3+, Fe2+, Mg2+, Ni2+, Cd2+, Co2+, Mn2+, Ca2+, Ba2+, Na+, K+, Ag+) (1 mM) in EtOH/H2O (v/v = 1/1) solution; (b) color change induced upon addition of Cu2+ under 365 nm UV lamp.
Chemosensors 11 00556 g002
Figure 3. (a) The fluorescence spectra of MPMC (10 μM) with the increasing concentration of Cu2+ ion (0.01−10.0 equiv.) in EtOH/H2O (v/v = 1/1). (b) The changes of fluorescence signal with different concentrations of copper ions. (c) The linear fit between MPMC and Cu2+.
Figure 3. (a) The fluorescence spectra of MPMC (10 μM) with the increasing concentration of Cu2+ ion (0.01−10.0 equiv.) in EtOH/H2O (v/v = 1/1). (b) The changes of fluorescence signal with different concentrations of copper ions. (c) The linear fit between MPMC and Cu2+.
Chemosensors 11 00556 g003
Figure 4. (a) The fluorescence spectra of MPMC in the presence of Cu2+ and EDTA in EtOH/H2O (v/v = 1/1) (λex = 358 nm, λem = 548 nm). (b) Changes in emission spectra of MPMC in the presence of Cu2+ and EDTA in EtOH/H2O (v/v = 1/1) (λex = 358 nm, λem = 548 nm).
Figure 4. (a) The fluorescence spectra of MPMC in the presence of Cu2+ and EDTA in EtOH/H2O (v/v = 1/1) (λex = 358 nm, λem = 548 nm). (b) Changes in emission spectra of MPMC in the presence of Cu2+ and EDTA in EtOH/H2O (v/v = 1/1) (λex = 358 nm, λem = 548 nm).
Chemosensors 11 00556 g004
Figure 5. Fluorescence intensity changes of MPMC probe (black line) and MPMC-Cu2+ complexes (red line) under different pH values in phosphate buffer system (λex = 358 nm, λem = 548 nm).
Figure 5. Fluorescence intensity changes of MPMC probe (black line) and MPMC-Cu2+ complexes (red line) under different pH values in phosphate buffer system (λex = 358 nm, λem = 548 nm).
Chemosensors 11 00556 g005
Figure 6. Time-dependent changes of MPMC (10 μM) (black line) with the addition of Cu2+ (10 μM) (red line) in EtOH/H2O (v/v = 1/1) (λex = 358 nm, λem = 548 nm).
Figure 6. Time-dependent changes of MPMC (10 μM) (black line) with the addition of Cu2+ (10 μM) (red line) in EtOH/H2O (v/v = 1/1) (λex = 358 nm, λem = 548 nm).
Chemosensors 11 00556 g006
Figure 7. Job’s plot of MPMC and copper ions ([MPMC] + [Cu2+] = 20 μM) in EtOH/H2O (v/v = 1/1) by fluorescence spectra, where the fluorescence intensity at 548 nm was plotted against the mole fraction of [[Cu2+]/[[MPMC] + [Cu2+]]].
Figure 7. Job’s plot of MPMC and copper ions ([MPMC] + [Cu2+] = 20 μM) in EtOH/H2O (v/v = 1/1) by fluorescence spectra, where the fluorescence intensity at 548 nm was plotted against the mole fraction of [[Cu2+]/[[MPMC] + [Cu2+]]].
Chemosensors 11 00556 g007
Figure 8. The proposed sensing mechanism of MPMC to Cu2+ in the system.
Figure 8. The proposed sensing mechanism of MPMC to Cu2+ in the system.
Chemosensors 11 00556 g008
Figure 9. Photographs showing the color changes of the MPMC probe (1.0 mM) before and after addition of Cu2+ (1.0 mM) under (a) sunlight and (b) 365 nm UV light.
Figure 9. Photographs showing the color changes of the MPMC probe (1.0 mM) before and after addition of Cu2+ (1.0 mM) under (a) sunlight and (b) 365 nm UV light.
Chemosensors 11 00556 g009
Table 1. Performance of MPMC compared with available Cu2+ probes.
Table 1. Performance of MPMC compared with available Cu2+ probes.
CompoundMolecular FormulaSolventRepeatabilityDetection LimitReference
Chemosensors 11 00556 i001C18H19N5O3CH3CN:HEPES (3:2, v/v)No0.12 μM[20]
Chemosensors 11 00556 i002C37H27NO3THF/H2ONo0.36 μM[22]
Chemosensors 11 00556 i003C22H14N4OEtOH-HEPESNo157 nM[32]
Chemosensors 11 00556 i004C23H16N4O3Tris-HCl buffer solutionYes84 nM[33]
Chemosensors 11 00556 i005C23H16N2O5EtOH-HEPES
(80:20, v/v)
No6.14 × 10−7 M[34]
Chemosensors 11 00556 i006C19H13N3O2SDMSO:H2O
(1:9, v/v)
No20.0 nM[35]
Chemosensors 11 00556 i007C16H9F3O4CH3CN/HEPES (95/5, v/v)No24.5 nM[36]
Chemosensors 11 00556 i008C19H15N3OSDMSONo3.3 μM[37]
Chemosensors 11 00556 i009C45H36N4O8S450% CH3CN and 20 mM HEPESNo390 nM[38]
Chemosensors 11 00556 i010C20H22N2O2Water/DMF
(9.9/0.1)
No-[39]
Chemosensors 11 00556 i011C13H11BrN2O2SDMSONo2.35 μM[40]
Chemosensors 11 00556 i012C21H16N4O4EtOH/H2O
(1:1, v/v)
Yes10.23 nMThis work
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

Liu, J.; Cheng, P.-Y.; Chen, S.; Wang, M.; Wei, K.; Li, Y.; Cao, Y.-Y.; Wang, X.; Li, H.-L. Preparation and Application of a Fast, Naked-Eye, Highly Selective, and Highly Sensitive Fluorescent Probe of Schiff Base for Detection of Cu2+. Chemosensors 2023, 11, 556. https://doi.org/10.3390/chemosensors11110556

AMA Style

Liu J, Cheng P-Y, Chen S, Wang M, Wei K, Li Y, Cao Y-Y, Wang X, Li H-L. Preparation and Application of a Fast, Naked-Eye, Highly Selective, and Highly Sensitive Fluorescent Probe of Schiff Base for Detection of Cu2+. Chemosensors. 2023; 11(11):556. https://doi.org/10.3390/chemosensors11110556

Chicago/Turabian Style

Liu, Juan, Peng-Yu Cheng, Sai Chen, Meng Wang, Kai Wei, Yuan Li, Yao-Yao Cao, Xing Wang, and Hong-Lei Li. 2023. "Preparation and Application of a Fast, Naked-Eye, Highly Selective, and Highly Sensitive Fluorescent Probe of Schiff Base for Detection of Cu2+" Chemosensors 11, no. 11: 556. https://doi.org/10.3390/chemosensors11110556

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