Next Article in Journal
Reactivity of the Ethenium Cation (C2H5+) with Ethyne (C2H2): A Combined Experimental and Theoretical Study
Previous Article in Journal
A Versatile Broadband Attached Proton Test Experiment for Routine 13C Nuclear Magnetic Resonance Spectroscopy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Naphthoquinoline-Dione-Based Cu2+ Sensing Probe with Visible Color Change and Fluorescence Quenching in an Aqueous Organic Solution

Department of Bionano Engineering, Hanyang University, Ansan 15588, Republic of Korea
*
Authors to whom correspondence should be addressed.
Molecules 2024, 29(4), 808; https://doi.org/10.3390/molecules29040808
Submission received: 28 December 2023 / Revised: 30 January 2024 / Accepted: 5 February 2024 / Published: 9 February 2024
(This article belongs to the Section Bioorganic Chemistry)

Abstract

:
Copper metal ions (Cu2+) are widely used in various industries, and their salts are used as supplementary components in agriculture and medicine. As this metal ion is associated with various health issues, it is necessary to detect and monitor it in environmental and biological samples. In the present report, we synthesized a naphthoquinoline-dione-based probe 1 containing three ester groups to investigate its ability to detect metal ions in an aqueous solution. Among various metal ions, probe 1 showed a vivid color change from yellow to colorless in the presence of Cu2+, as observed by the naked eye. The ratiometric method using the absorbance ratio (A413/A476) resulted in a limit of detection (LOD) of 1 µM for Cu2+. In addition, the intense yellow-green fluorescence was quenched upon the addition of Cu2+, resulting in a calculated LOD of 5 nM. Thus, probe 1 has the potential for dual response toward Cu2+ detection through color change and fluorescence quenching. 1H-NMR investigation and density functional theory (DFT) calculations indicate 1:1 binding of the metal ion to the small cavity of the probe comprising four functional groups: the carbonyl group of the amide (O), the amino group (N), and two t-butyl ester groups (O). When adsorbed onto various solid surfaces, such as cotton, silica, and filter paper, the probe showed effective detection of Cu2+ via fluorescence quenching. Probe 1 was also useful for Cu2+ sensing in environmental samples (sea and drain water) and biological samples (live HeLa cells).

Graphical Abstract

1. Introduction

Potential adaptability, easy synthesis and structural modification, and outstanding photo-physical properties of small molecular probes led to their wide applications in various fields, including medicinal chemistry, bio-imaging, drug delivery, solar cells, and OLEDs [1,2,3,4]. Recently, significant attention has been paid to molecular probes that exhibit a vibrant color change in the presence of specific analytes [5,6,7,8]. Metal ion sensing in aqueous systems is particularly demanding due to the critical roles of various metal ions in physiology. However, at their individual optimal concentrations, these metal ions perform diverse functions, including pH regulation, osmotic balance, metabolism, and signaling transduction [9,10]. Their abnormal concentrations could lead to damage of organs that can eventually show detrimental effects on the human body, such as physical disorders and acute/chronic diseases [11,12,13]. Heavy metal ions, such as Hg2+/CH3Hg+ and Pb2+, are toxic to humans [14,15]. Thus, selective detection of individual metal ions is indispensable, and monitoring the quantity/concentration of these metal ions in environmental sources is required and should be regulated, especially in drinking water (e.g., tap/lake water) [16,17,18,19]. Cu2+ is the third most common heavy metal ion in the human body and functions to sustain bones, blood vessels, the immune system, nerves, and metabolism, and it serves a critical function in iron absorption in the human body [20]. However, excessive Cu2+ exposure, even for a limited time, can lead to serious health concerns, such as liver/kidney issues, gastrointestinal (GI) dysfunction, and many serious neurodegenerative diseases (e.g., Menkes disease) [21,22,23,24,25]. It is also widely used in construction, transport, and electronics, and many of its salts are being used as supplementary components in agriculture and medicine [26,27]. This metal ion seeps into groundwater and deteriorates the quality of drinking water, posing a threat to human health. The U.S. Environmental Protection Agency (EPA) [28] currently sets the limit of Cu2+ concentration in drinking water at 1.3 ppm (~20 µM).
Among various analytical techniques to detect metal ions in biological and environmental samples [29], fluorescence spectroscopy possesses many advantages over others in terms of affordability, quick response, convenience, easy operation, on-site detection/monitoring, and naked eye recognition [30,31,32,33,34,35]. Thus, many researchers have made efforts to create selective, reliable, convenient, and efficient sensors with a variety of fluorescent platforms, such as fluorescein, rhodamine B, triphenylamine, quinolone, pyrene, anthraquinone, chalcone, benzothiazole/benzimidazole amine, coumarin-conjugated di-(2-picolyl)amine, peptide, and Schiff base compounds [36,37,38,39,40,41,42,43]. Many of these probes utilized ethylene diamine-like structures for strong and selective Cu2+ binding among different metal ions, leading to high selectivity toward Cu2+ detection. For instance, S. Anbu et al. reported a benzimidazole-based bis-Schiff base probe as a fluorogenic chemosensor. The Schiff base-benzimidazole conjugate mimics ethylene diamine as a metal ion-binding group. This probe detected Cu2+ and Zn2+ in an aqueous solution via fluorescence ‘turn-on’ and ‘turn-off’ mechanisms, respectively, with limits of detection (LODs) of 24 and 2 nM [36]. K. Guo and J. Li used an imine-linked triphenylamine–benzothiazole conjugate as a metal ion-sensing probe. The probe containing an ethylenediamine-like binding site for metal ions was able to detect two metal ions (Cu2+ and Fe3+) in CH3CN, with LODs of 0.3 and 3 μM, respectively [37]. The probe showed a color change from colorless to red in the presence of Cu2+, while the interaction of the probe with Fe3+ resulted in the ‘switch on’ fluorescence emission. In an aqueous CH3CN solution, this benzothiazole-based probe was selective for Zn2+, with an LOD of 3 μM. D. Aydin et al. reported a phenolphthalein-based Schiff base probe for selective Cu2+ detection. The metal binding site of the probe is structurally similar to ethylenediamine but comprises a different spacer (propylene) and atom (oxygen) from the latter. Upon the addition of the metal ion, the probe displayed a color change from yellow to colorless in an HEPES:CH3CN (1:1, pH = 7) solution. In addition, the fluorescence emission of the probe turned off in the presence of Cu2+, giving an LOD of 84 nM [38]. G. Singh et al. reported a pyrene-bearing organosilane-based nanoparticle for the selective detection of Cu2+. This pyrene-decorated nanoparticle also contains an ethylenediamine-like binding site and showed a sensitive Cu2+ detection (LOD = 6 µM) in methanol [43]. Inspired by the structural feature of such ethylenediamine-like binding sites present in these previously reported probes, we prepared a new probe (probe 1) containing three ester groups based on the naphthoquinoline-dione platform in the current study. The resulting three ester group probe selectively detected Cu2+ by luminescence/fluorescence and absorption/color change. The probe 1 solution changed from bright yellow to colorless upon Cu2+ binding, demonstrating a low calculated LOD of 5 nM. Cu2+ sensing was also realized using this probe adsorbed onto solid surfaces (cotton, filter paper, silica gel, and TLC plate). In addition, Cu2+ added in environmental samples (sea/drain water) or bio-samples (HeLa cells) was successfully detected by the probe.

2. Results and Discussion

2.1. Synthesis and Characterization of Probe 1

Probe 1 was synthesized from compound 3 via a two-step reaction (Scheme 1). The amino group bearing the naphthoquinoline-dione derivative (compound 3) was reacted with t-butyl bromoacetate in DMSO using K2CO3 (base) and TBAHSO4 (catalyst) to produce compound 2 (50%). The formation of compound 2 was verified by its 1H-NMR spectrum, where the amidic NH of compound 3 was lost at 12.45 ppm, and two new singlets of NCH2 and t-butyl groups appeared in the aliphatic region at δ 5.13 and 1.49 ppm, respectively (Figure 1A). A further reaction of compound 2 with t-butyl bromoacetate gave probe 1 a moderate yield (50%). The 1H NMR spectrum of the probe showed a loss of the NH2 signal at δ 5.89 ppm with the simultaneous appearance of four doublets at δ 3.95, 4.93, 5.16, and 5.63 ppm, corresponding to two CH2 groups attached to the amino group. These spectral features match the chemical structure of probe 1 containing three t-butyl acetate groups (Figure 1A). The 2D NMR (1H-1H NOESY and COSY) spectra verify the chemical structure of probe 1 (Figure 1B–E and Figures S1–S4). The 2D NMR (1H-1H NOESY) spectrum displayed a strong correlation between the three t-butyl protons, their neighboring protons (NαCH2), and the nearest aromatic protons (Figures S1 and S2). The CH2 attached to the amidic N correlates with the aromatic proton (Hg), while two rigid CH2 groups attached to the amino N strongly correlate with the nearby aromatic proton (Ha) (Figure 1B,C). Further correlations were observed between Hg (doublet) and Hf (triplet), as well as between Hf (triplet) and He (doublet) (Figure 1B,C). The 2D NMR (1H-1H COSY) spectrum of the probe revealed correlations between two neighboring protons (Figure 1D,E and Figures S3–S5). The 13C NMR and HRMS spectra further confirmed the identity of the probe (see electronic Supplementary Materials).

2.2. Spectroscopic Characterization and Sensing Ability of Probe 1

To explore the probe potential in metal ion sensing, we performed fluorescence studies of probe 1 in acetonitrile (ACN)). In the range of excitation wavelengths from 390 to 500 nm, the probe exhibited maximum emission or the highest quantum yield (ø = 0.45) at 470 nm (Figure S6). Thus, this excitation wavelength was chosen for further fluorescence studies of the probe. Different trivalent (Cr3+, Fe3+, Ga3+, Al3+, and Ru3+), divalent (Mg2+, Ca2+, Ba2+, Fe2+, Co2+, Zn2+, Cd2+, Pb2+, Cu2+, Ni2+, and Hg2+), and monovalent metal ions (Ag+, Cs+, Na+, and K+) were added individually to the ACN solution containing probe 1 at 5 µM. Among these metal ions, only Cu2+ was able to selectively quench the green luminescence of probe 1 (Figure S7A,B and Figure 2A). Of note, compound 3, a starting material for the preparation of probe 1, was previously found to be Hg2+ selective [44]. Interestingly enough, the structural modification from compound 3 to probe 1 changed the probe selectivity toward Cu2+ from Hg2+. In the titration, upon the sequential addition of Cu2+ in ACN, the fluorescence intensity of probe 1 (5 µM, ACN) was reduced regularly and became saturated at about 60 µM [Cu2+] (Figure 2B and Figure S7C). This titration data indicate 1:1 stoichiometry between probe 1 and the metal ion with a high association constant (Ka = 3.98 × 104 M−1) (Figure S7C). The fluorescence emission intensity at 528 nm (I528) linearly decreased with increasing [Cu2+], demonstrating an LOD of 0.5 µM for this sensing system (inset of Figure 2B). In addition, no significant interference was observed from other metal ions (Figure S7D).
The UV–Vis spectrum of probe 1 (50 µM, ACN) exhibited a strong band at 476 nm, likely originating from the n → π* transition of the naphthoquinoline-dione system. Among the various metal ions studied, the addition of only Cu2+ resulted in a significant change in the UV–Vis absorption spectrum. The absorbance maximum was blue shifted from 476 to 413 nm, along with the appearance of a shoulder at 431 nm (Figure 2C). Upon the addition of Cu2+, the probe 1 solution changed from yellow-green to pale green in daylight, as detected by the naked eye and reflected in the absorption spectra obtained under the same conditions (Figure 2C,D). Over the course of UV–Vis titration carried out in ACN, the absorption band of probe 1 at 476 nm decreased with an increasing amount of Cu2+. At the same time, a new band of the probe at 413 nm appeared, with an isosbestic point observed at 445 nm Figure 2E,F). The absorption spectrum of the probe failed to result in further changes when more than 10 equiv. of Cu2+ was added to the probe solution. Non-linear regression analysis of this titration data suggests the formation of a 1:1 complex between the probe and the metal ion (Ka = 3.41 × 103 M−1) (Figure 2F). The linear relationship between A476 and [Cu2+] observed at the initial stage of the titration was used to obtain an LOD of 1 µM for the sensing system (Figure S8A). A ratiometric approach using the ratio of two absorption values (A413/A476) allowed us to detect Cu2+ over a wide range of concentrations (1 µM to 1 mM), with an observed LOD of 1 µM (Figure S8B). The probe 1 solutions containing different metal ions showed no noticeable alteration in absorption spectra upon the addition of Cu2+, indicating that these metal ions did not interfere with the Cu2+ detection of the probe (Figure S9). In addition, the reversibility of probe 1 in Cu2+ binding was demonstrated by the use of EDTA (disodium ethylene-diamine tetra-acetic acid) as a metal ion chelator. The green luminescence observed for the free probe was fully recovered after adding EDTA to the [probe 1-Cu2+] solution (Figure S10).
To evaluate the solvent effects on the probe’s ability to sense Cu2+, we prepared probe 1 solutions at 5 µM using various solvents and added 20 equiv. of Cu2+ to each (Figure 3 and Figure S11). The yellow-greenish luminescence was significantly quenched (60 to 98%) for the probe dissolved in ACN, THF, acetone, DCM, or CHCl3. Upon the addition of Cu2+, the probe dissolved in EtOH, MeOH, tBuOH, or AW11 (ACN/water (1:1)) resulted in moderate luminescence quenching, whereas little change in luminescence intensity was observed for the probe in pure water, dioxane, or toluene (Figure 3C). To explore the potential utility in real sample applications, we investigated the Cu2+-induced changes in photoluminescence and fluorescence of the probe using a binary solvent system of ACN (A) and water (W) (Figure 4 and Figure S12). With water percentage increasing up to 60% in ACN, the fluorescence emission peak of probe 1 gradually red shifted from 528 nm (green) to 567 nm (yellow), with slight decreases in emission intensity (Figure 4A and Figure S12A). When 80% water in ACN was used as a medium, we found a 60% decrease in the fluorescence intensity of probe 1 compared to that in pure ACN. This could be due to the occurrence of substantial probe aggregation, as expected from the presence of three bulky t-butyl groups in the probe. Upon the addition of Cu2+, significant fluorescence quenching was observed for probe 1 dissolving in the different binary solvents (Figure 4B and Figure S12B). This quenching effect on probe fluorescence became less obvious when the water content was greater than 80% in ACN (Figure 4C). Thus, Cu2+ can be effectively detected via fluorescence quenching when the water content is less than 80% in ACN. We chose 20% aqueous ACN as a binary solvent system for further studies of the probe.
Upon excitation at 470 nm, the probe 1 solution (5 µM, HEPES (pH 7.4): ACN = (1:4)) exhibited an intense yellow fluorescence emission at 550 nm, along with a good quantum yield (Φ = 0.40) (Figure S13). Among the various metal ions, the probe responded only to Cu2+; the yellow luminescence and fluorescence emission of probe 1 were quenched (Figure 5A). Titration with Cu2+ showed a gradual decrease in fluorescence emission intensity of the probe at 550 nm (I550) with increasing [Cu2+] (Figure 5B and Figure S14). The titration data (I550 vs. [Cu2+]) are consistent with a 1:1 complex formation between probe 1 and Cu2+ with an association constant (Ka = 1.91 × 104 M−1). This association constant corresponds to nearly half of that observed for the [probe 1-Cu2+] complex in ACN (Figure S14), indicating weaker binding of the metal ion to the probe in aqueous ACN solutions compared to pure ACN. The probe detected Cu2+ at concentrations as low as 0.5 µM, while the calculated LOD was 5 nM (inset of Figure 5B and Figure S15). This LOD is comparable to or better than those of many recently reported Cu2+-sensing probes (Table S1). We further applied the probe for metal ion sensing in seawater and drain water from a small lake, which showed the detection of Cu2+ as low as 0.5 µM. A slight decrease was found in probe sensitivity for the application (Figure S16).

2.3. Mechanism of Cu2+ Interaction with Probe 1

To explore the molecular interaction between probe 1 and Cu2+, we conducted 1H NMR titration of the probe with Cu2+ in DMSO-d6. With a gradual increase in Cu2+ concentration from 0 to 1 equiv., most of the probe signals (e.g., N(CH2) signals) were up-field shifted, along with notable broadening of the aromatic proton peaks (Ha, Hf, and He). These changes are likely due to the paramagnetic character of Cu2+ combined with probe aggregation upon binding to the metal ion (Figure 6 and Figure S17). This result indicates that the tertiary N of naphthoquinoline-dione and the carbonyl (-CO) oxygen of the ester group coordinate in Cu2+ binding. Further addition of Cu2+ up to 2 equiv. failed to produce a further change in the 1H NMR signals, ensuring the 1:1 complex formation between the probe and Cu2+.
To further understand the molecular interactions of Cu2+ with probe 1, density functional theory (DFT) calculations were performed at the B3LYP/6-31G* level (Figure 7 and Figures S18–S20). Such calculations provide information about the energy levels and electronic distributions of the frontier molecular orbitals (FMOs) (i.e., HOMO [highest occupied molecular orbital] and LUMO [lowest unoccupied molecular orbital]). As the 1:1 complex between probe 1 and the metal ion was demonstrated by various experiments (fluorescence, UV–Vis, and NMR studies), we used this stoichiometry as a model for the calculations. The calculations indicate that four atoms (i.e., amine N, two O’s of two different ester groups, and carbonyl O of the amide in probe 1) coordinate with Cu2+ to form the tetrahedral complex (Figure 7A and Figures S18–S20). Upon complexation, the ΔEH/L (HOMO and LUMO energy gap) was increased from 3.16 to 3.22 eV (Figure 7B), qualitatively explaining the blue shifts of the absorption and fluorescence spectra observed for probe 1 upon interaction with Cu2+. A minor deviation from the experimental result may be due to the fact that we carried out the calculations under the vacuum condition. The HOMO electron density of probe 1 mainly was distributed around the tertiary amino N and the amide group-containing ring of naphthoquinoline-dione, while the LUMO electron density was localized on the naphthoquinoline-dione ring. The electron densities calculated for the aHOMO and bHOMO of probe 1 complexed with Cu2+ were both mainly located around the carbonyl functional group of the naphthoquinoline-dione ring. However, the LUMO electron density of the probe moved to the metal binding site upon complexation with the metal ion (Figure 7B). This electron transfer is most likely responsible for the effective quenching of the green/yellow fluorescence emission of probe 1 upon Cu2+ binding (Scheme 2).

2.4. Practical Application

To examine the practical utility, we first investigated the ability of the probe to detect Cu2+ in a solid state. For this purpose, various solid surfaces, such as a TLC plate, cotton, and Whatman filter paper, were used for probe adsorption. The probe adsorbed on these solid supports yielded bright yellow luminescence under 365 nm illumination (Figure 8). The bright yellow ‘Cu2+’ written on the TLC plate using the probe 1 solution (5 µM, ACN) was quenched following the application of Cu2+ spray (20 µM) (Figure 8A). Similar results were obtained for probe 1 on cotton buds and silica gels (Figure 8B,C). Yellow luminescence of the spots on the filter paper and TLC plate gradually weakened with increasing amounts of Cu2+. The filter paper and TLC plate-supported probe 1 showed the ability to detect 0.5 µM and 1 µM Cu2+, respectively (Figure 8D,E).
Probe 1 was further applied to detect Cu2+ in live HeLa cells. For this experiment, HeLa cells were incubated with probe 1 over a range of concentrations (0–20 µM) for 30 min. More than 80% of the cells treated with the probe survived under the conditions, indicating that the probe is only slightly toxic to the cells (Figure S21). Under irradiation at 488 nm, the cells without the probe displayed negligible emission, but incubation of the cells with probe 1 (5 µM) produced an intense green emission in the cytoplasm (Figure 9). This optical change confirms that the probe penetrated the plasma membranes but not the double-layered nuclear membranes. When the HeLa cells containing probe 1 were treated with 20 µM Cu2+, the green emission of the cells was completely quenched. This quenching likely originates from the complexation of the probe with metal ions in the cytoplasm. The bright-field images indicate that these cells were viable throughout the experiment. Hence, intracellular Cu2+ can be detected in a non-toxic manner using probe 1, which is crucial for bio-applications.

3. Materials and Methods

3.1. General Chemicals and Materials

All reagents and solvents used in the experiments, including t-butyl bromoacetate, metal perchlorates, tetrabutylammonium bisulfate (TBAHSO4), dimethyl sulfoxide (DMSO), THF, acetonitrile (ACN), and potassium carbonate (K2CO3), were purchased from Aldrich (St. Louis, MO, USA) (analytical grade) and used without further purification. The compound 1-amino-3H-naphtho[1,2,3-de]quinoline-2,7-dione (compound 3) was synthesized according to a procedure reported previously [44]. All the tested metal ions were used in the perchlorate forms (M(ClO4)x), except Ru3+ and Au3+, which were used in the form of chloride salts. Stock solutions of the metal ions were prepared in ACN (1 mM) or water (10 mM) depending on the sensing medium (ACN or HEPES (pH 7.4: ACN (1:4)) used for metal ion detection. All the fluorescence spectra were recorded on a ChronosBH fluorescence lifetime spectrometer. UV–Vis (ultraviolet–visible) spectra were recorded on a Shimadzu UV-2600PC spectrophotometers with a quartz cuvette (Shimadzu, Kyoto, Japan). The cell holder was maintained at 25 °C. The 1H and 13C nuclear magnetic resonance (NMR) spectra of all new compounds were recorded on a Bruker 400 MHz spectrophotometer using CDCl3 or DMSO-d6 as a solvent and tetramethylsilane (TMS) as an internal standard. The NMR data are reported as chemical shifts in ppm (parts per million) (δ), multiplicities (s = singlet, d = doublet, m = multiplet), coupling constants (Hz), integration, and interpretation. All spectrophotometric titration curves were fitted with gnu plot software 5.0.

3.2. Photo-Physical Studies: Parameters and Conditions

All metal ion-induced color changes, UV–Vis, and fluorescence spectra were obtained in ACN (CH3CN) or a binary solvent (HEPES (pH 7.4): ACN (1:4)). All absorption and fluorescence scans were saved as ACSII files and further processed in ExcelTM to produce the graphs shown. A stock solution of probe 1 was prepared at 1 mM in ACN and was used for photo-physical studies following appropriate dilution with ACN or a binary solvent (HEPES (pH 7.4): ACN (1:4)). The UV–Vis and fluorescence studies were performed using the probe at 5 and 50 µM, respectively. The titration experiments were carried out by adding Cu2+ to the probe 1 solution until the absorption/fluorescence spectra of the solutions reached saturation. The association constants between the probe and Cu2+ were determined by fitting the absorption and fluorescence spectral data obtained from the titration experiment of probe 1 with Cu2+. The titration data were fitted with the global analysis program gnu plot software.

3.3. Theoretical Calculations

Energy-optimized conformations of probe 1 and its Cu2+ complex were obtained via gradient-correlated DFT (density functional theory) calculations using Becke’s three-parameter exchange functional [45] and the Lee–Yang–Parr (B3LYP) exchange–correlation function [46] with 6-31G* basis sets for C, H, N, and O. B3LYP/6-31G* calculations were employed for excited-state optimization. For simplicity, all calculations were carried out in a vacuum to avoid the solvent effect. All stationary points were verified as the minima via calculations from Hessian and harmonic frequency analyses [47,48].

3.4. Synthesis of Probe 1

3.4.1. Tert-butyl 2-(1-amino-2,7-dioxo-2,7-dihydro-3H-cnaphtho[1,2,3-de]quinolin-3-yl)acetate (2)

A solution of 1-amino-3H-naphtho[1,2,3-de]quinoline-2,7-dione (compound 3, 1.13 g, 5 mmol) in DMSO (10 mL) containing K2CO3 (1.38 g, 10 mmol) and TBAHSO4 (phase transfer catalyst; 2%) was stirred at room temperature for 20 min. Next, tert-butyl bromoacetate (0.975 mg, 5.0 mmol) was added to the reaction mixture, and the resulting solution was stirred at room temperature for 48 h. The progress of the reaction was monitored by TLC; upon reaction completion, a brine solution was added. A solid material collected by filtration was identified as unreacted compound 3. The filtrate was concentrated and purified by column chromatography to yield the product (2) (50%) as a yellow solid. 1H NMR (400 MHz, CDCl3): δ 1.49 (s, 9H, 3 × CH3), 5.13 (s, 2H, CH2), 5.89 (s, 2H, NH2), 7.35 (d, J = 8.0 Hz, 1H, ArH), 7.51–7.57 (m, 2H, 2 × ArH), 7.76 (t, J = 8.0 Hz, 1H, ArH), 8.33 (d, J = 8.0 Hz, 1H, ArH), 8.51 (d, J = 8.0 Hz, 1H, ArH), and 8.54 (d, J = 8.0 Hz, 1H, ArH); 13C NMR (100 MHz, CDCl3): δ 28.24, 46.02, 83.40, 107.65, 117.91, 120.93, 123.85, 125.96, 126.00, 128.11, 128.91, 131.85, 132.62, 133.54, 135.57, 136.77, 159.03, 166.82, and 182.69. HRMS (M + H+) = 377.1499; C22H21N2O4 (theoretical (M + H+) = 377.1501).

3.4.2. Probe 1

Compound 2 (190 mg, 0.5 mmol) was stirred vigorously in THF (20 mL) at room temperature, followed by the addition of K2CO3 (1.38 g, 10 mmol) and TBAHSO4 (phase transfer catalyst; 2%) for one hour under an N2 atmosphere. Then, t-butylbromoacetate (1.95 g, 10 mmol) was added dropwise to the reaction mixture, followed by stirring at room temperature for 72 h. TLC was applied to monitor the progress of the reaction; upon reaction completion, a brine solution was added. This solution was extracted with an equal portion of DCM (x2) to extract the organic compounds completely. The solution was dried with anhydrous Na2SO4, concentrated, and subjected to column chromatography to afford the pure product (1) (50%) as a red solid. 1H NMR (CDCl3, 400 MHz): δ 1.32 (s, 9H, 3 × CH3), 1.38 (s, 9H, 3 × CH3), 1.45 (s, 9H, 3 × CH3), 3.95 (d, J = 18.0 Hz, 1H (NCH), one H of rigid two NCH′H), 4.93 (d, J = 17.2 Hz, 1H (NCH′), one H of rigid two NCH′H), 5.15 (s, 1H of NCH2, the other H of this CH2 could exchange with D due to its H-bonding with –C=O (oxygen)), 5.16 (d, J = 17.2 Hz, 1H (NCH′), one H of rigid two NCH′H), 5.63 (d, J = 18.0 Hz, 1H (NCH), one H of rigid two NCH′H), 7.29 (d, J = 8.0 Hz, 1H ArH), 7.53 (t, J = 8.0 Hz, 1H ArH), 7.55–7.60 (m, 2H, 2 × ArH), 8.32 (d, J = 8.0 Hz, 1H ArH), and 8.37–8.39 (m, 2H, 2 × ArH); 1H NMR (DMSO-d6, 400 MHz): δ 1.25 (s, 9H, 3 × CH3), 1.32 (s, 9H, 3 × CH3), 1.43 (s, 9H, 3 × CH3), 4.22 (d, J = 18.0 Hz, 1H (NCH), one H of rigid two NCH′H), 5.12 (q, J = 17.6 Hz, 2H, 2 × (NCH′)), 5.28 (d, J = 18.0 Hz, 1H (NCH′), one H of rigid two NCH′H), 5.67 (s, 1H of NCH2, the other H of this CH2 could exchange with D due to its H-bonding with –C=O (oxygen)), 7.64 (t, J = 8.0 Hz, 1H ArH), 7.71–7.75 (m, 2H, 3 × ArH), 8.18 (d, J = 8.0 Hz, 1H ArH), and 8.25–8.27 (m, 2H, 2 × ArH); 13C NMR (CDCl3, 100 MHz): δ 28.03, 28.19, 45.60, 56.48, 68.17, 82.11, 82.95, 83.21, 111.17, 117.55, 118.89, 123.57, 126.43, 127.75, 127.93, 128.27, 128.62, 129.32, 131.40, 134.26, 135.03, 157.67, 166.99, 168.93, 169.22, and 182.65. HRMS (M + H+) = 605.2861; C34H40N2O8 (theoretical (M + H+) = 605.2857).

3.5. Live Cell Imaging

HeLa cells were obtained from the KCLB (Korean Cell Line Bank, Seoul, Republic of Korea) and grown in MEM (Minimum Essential Media) supplemented with 10% fetal bovine serum (FBS), 100 µg/mL streptomycin, and 100 IU/mL penicillin. The cells were maintained in a humidified incubator at 37 °C at 5% CO2. A total of 3.8 × 104 cells were seeded onto a 13 mm glass bottom of a 35 mm confocal dish and grown for one day/24 h (until 60–70% confluence). The probe stock solution was prepared by dissolving probe 1 in DMSO. Experiments were performed in triplicate in FBS and antibiotic-free medium as follows. HeLa cells were incubated with probe 1 (5 µM) at 37 °C at 5% CO2 for half an hour. The final DMSO concentration was 0.5% in the sample. After washing the incubated sample twice with 1 x phosphate-buffered saline (PBS) (pH = 7.4), the cells were further incubated with Cu2+ (20 µM, H2O) for half an hour under the same conditions and were washed three times with 1 × PBS. A K1-Fluo confocal laser scanning microscope with a 40× oil immersion objective lens was used to obtain confocal microscopy images. Bright-field images after treatment with probe 1 and Cu2+ were obtained to evaluate the cell viability.

4. Conclusions

We prepared a new naphthoquinoline-dione-based probe 1 containing three ester groups starting from compound 3 with no ester groups. While compound 3 was selective for Hg2+ sensing, the new probe 1 exhibited high selectivity for Cu2+ in ACN, as well as in a 20% aqueous ACN solution. The probe displayed a color change from yellow to colorless upon selective interaction with Cu2+. A ratiometric approach using two absorption values (A413/A476) allowed the detection of Cu2+ at as low as 1.0 µM. In addition, the yellow fluorescence emission of the probe in the aqueous solution (HEPES (pH 7.4): ACN = 1:4) was selectively quenched in the presence of Cu2+ among various metal ions. Hence, probe 1 provides a dual sensing method for Cu2+ detection (absorption and fluorescence emission) and is chromo-fluorogenic. Fluorescence titration of the probe with Cu2+ produced a low calculated LOD of 5 nM. Fluorescence, UV–visible, and 1H-NMR titration studies, in addition to DFT calculations, supported the 1:1 complex formation between Cu2+ and probe 1. This probe showed the ability to detect Cu2+ in a real sample (sea/drain water) with an observed LOD of 0.5 µM, which signifies its potential for on-site monitoring of the metal ion. In addition, the probe was successfully used for Cu2+ detection after adsorption onto solid supports, such as cotton, filter paper, a TLC plate, and silica gels. Importantly, the probe was able to penetrate cell membranes and detect Cu2+ in the cytoplasm of HeLa cells. Therefore, this study provides insight into the development of new chromo-fluorogenic heteroaromatic sensors for the detection of different analytes.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29040808/s1.

Author Contributions

Conceptualization, A.K. and P.S.C.; methodology, A.K.; software, A.K.; validation, A.K.; formal analysis, A.K.; investigation, A.K.; resources, P.S.C.; data curation, A.K.; writing—original draft preparation, A.K.; writing—review and editing, A.K. and P.S.C.; visualization, A.K.; supervision, P.S.C.; project administration, A.K.; funding acquisition, P.S.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Research Foundation of Korea (NRF) (2018R1A6A1A03024231 to P.S.C.).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in article and Supplementary Materials.

Conflicts of Interest

There are no conflicts of interest to declare.

References

  1. Feng, X.; Wang, X.; Redshaw, C.; Tang, B.Z. Aggregation behaviour of pyrene-based luminescent materials, from molecular design and optical properties to application. Chem. Soc. Rev. 2023, 52, 6715–6753. [Google Scholar] [CrossRef]
  2. Chanda, K.; Balamurali, M.M. Light emitting probes—Approaches for interdisciplinary applications. Chem. Soc. Rev. 2021, 50, 3706–3719. [Google Scholar] [CrossRef] [PubMed]
  3. Yang, M.; Zhang, M.; Jia, M. Optical sensor arrays for the detection and discrimination of natural products. Nat. Prod. Rep. 2023, 40, 628–645. [Google Scholar] [CrossRef]
  4. Park, S.-H.; Kwon, N.; Lee, J.-H.; Yoon, J.; Shin, I. Synthetic ratiometric fluorescent probes for detection of ions. Chem. Soc. Rev. 2020, 49, 143–179. [Google Scholar] [CrossRef]
  5. Chua, M.H.; Hui, B.Y.K.; Chin, K.L.O.; Zhu, Q.; Liu, X.; Xu, J. Recent advances in aggregation-induced emission (AIE)-based chemosensors for the detection of organic small molecules. Mater. Chem. Front. 2023, 7, 5561–5660. [Google Scholar] [CrossRef]
  6. Chyan, W.; Zhang, D.Y.; Lippard, S.J.; Radford, R.J. Reaction-based fluorescent sensor for investigating mobile Zn2+ in mitochondria of healthy versus cancerous prostate cells. Proc. Natl. Acad. Sci. USA 2014, 111, 143–148. [Google Scholar] [CrossRef] [PubMed]
  7. Wu, D.; Chen, L.; Lee, W.; Ko, G.; Yin, J.; Yoon, J. Recent progress in the development of organic dye based near-infrared fluorescence probes for metal ions. Coord. Chem. Rev. 2018, 354, 74–97. [Google Scholar] [CrossRef]
  8. Peng, S.; He, Q.; Vargas-Zúñiga, G.I.; Qin, L.; Hwang, I.; Kim, S.K.; Heo, N.J.; Lee, C.-H.; Dutta, R.; Sessler, J.L. Strapped calix[4]pyrroles: From syntheses to applications. Chem. Soc. Rev. 2020, 49, 865–907. [Google Scholar] [CrossRef]
  9. Berg, J.M.; Shi, Y. The galvanization of biology: A growing appreciation for the roles of zinc. Science 1996, 271, 1081–1085. [Google Scholar] [CrossRef]
  10. Shimizu, T.; Lengalova, A.; Martinek, V.; Martinkova, M. Heme: Emergent roles of heme in signal transduction, functional regulation and as catalytic centres. Chem. Soc. Rev. 2019, 48, 5624–5657. [Google Scholar] [CrossRef]
  11. Sarpong-Kumankomah, S.; Gibson, M.A.; Gailer, J. Organ damage by toxic metals is critically determined by the bloodstream. Coord. Chem. Rev. 2018, 374, 376–386. [Google Scholar] [CrossRef]
  12. Goodman, J.E.; Prueitt, R.L.; Dodge, D.G.; Thakali, S. Carcinogenicity assessment of water-soluble nickel compounds. Crit. Rev. Toxicol. 2009, 39, 365–417. [Google Scholar] [CrossRef] [PubMed]
  13. Carter, K.P.; Young, A.M.; Palmer, A.E. Fluorescent Sensors for Measuring Metal Ions in Living Systems. Chem. Rev. 2014, 114, 4564–4601. [Google Scholar] [CrossRef] [PubMed]
  14. Tchounwou, P.B.; Yedjou, C.G.; Patlolla, A.K.; Sutton, D.J. Heavy Metals Toxicity and the Environment. Mol. Clin. Environ. Toxicol. 2014, 2012, 133–164. [Google Scholar]
  15. Clarkson, T.W.; Magos, L. The Toxicology of Mercury and Its Chemical Compounds. Crit. Rev. Toxicol. 2006, 36, 609–662. [Google Scholar] [CrossRef] [PubMed]
  16. Kumar, A.; Chae, P.S.; Kumar, S. A dual-responsive anthrapyridone-triazole-based probe for selective detection of Ni2+ and Cu2+: A mimetic system for molecular logic gates based on color change. Dye. Pigment. 2020, 174, 108092. [Google Scholar] [CrossRef]
  17. Lippard, S.J.; Berg, J.M. Principles of Bioinorganic Chemistry; University Science Books: Mill Valley, CA, USA, 1994. [Google Scholar]
  18. Kumar, A.; Kumar, S.; Chae, P.S. A Chromo-Fluorogenic Naphthoquinolinedione-Based Probe for Dual Detection of Cu2+ and Its Use for Various Water Samples. Molecules 2022, 27, 785. [Google Scholar] [CrossRef] [PubMed]
  19. Kumar, A.; Vanita, V.; Walia, A.; Kumar, S. N,N-dimethylaminoethylaminoanthrone—A chromofluorogenic chemosensor for estimation of Cu2+ in aqueous medium and HeLa cells imaging. Sens. Actuators B 2013, 177, 904–912. [Google Scholar] [CrossRef]
  20. Arena, G.; Mendola, D.L.; Pappalardo, G.; Sóvágó, I.; Rizzarelli, E. Interactions of Cu2+ with prion family peptide fragments: Considerations on affinity, speciation and coordination. Coord. Chem. Rev. 2012, 256, 2202–2218. [Google Scholar] [CrossRef]
  21. Kardos, J.; Héja, L.; Simon, Á.; Jablonkai, I.; Kovács, R.; Jemnitz, K. Copper signalling: Causes and consequences. Cell Commun. Signal. 2018, 16, 1–22. [Google Scholar] [CrossRef]
  22. Zheng, W.; Monnot, A.D. Regulation of brain iron and copper homeostasis by brain barrier systems: Implication in neurodegenerative diseases. Pharmacol. Ther. 2012, 133, 177–188. [Google Scholar] [CrossRef]
  23. Squitti, R. Copper dysfunction in Alzheimer’s disease: From meta-analysis of biochemical studies to new insight into genetics. J. Trace Elem. Med. Biol. 2012, 26, 93–96. [Google Scholar] [CrossRef]
  24. Madsen, E.; Gitlin, J.D. Copper and Iron Disorders of the Brain. Annu. Rev. Neurosci. 2007, 30, 317. [Google Scholar] [CrossRef]
  25. 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. [Google Scholar] [CrossRef]
  26. Hordyjewska, A.; Popiołek, Ł.; Kocot, J. The many ”faces” of copper in medicine and treatment. Biometals 2014, 27, 611–621. [Google Scholar] [CrossRef] [PubMed]
  27. Lamichhane, J.R.; Osdaghi, E.; Behlau, F.; Jones, J.K.J.B.; Aubertot, J.-N. Thirteen decades of antimicrobial copper compounds applied in agriculture. A review. Agron. Sustain. Dev. 2018, 38, 1–18. [Google Scholar] [CrossRef]
  28. World Health Organization. Guidelines for Drinking Water Quality, 3rd ed.; World Health Organization: Geneva, Switzerland, 2004; p. 188. [Google Scholar]
  29. Prodi, L.; Bolletta, F.; Montalti, M.; Zaccheroni, N. Luminescent chemosensors for transition metal ions. Coord. Chem. Rev. 2000, 205, 59. [Google Scholar] [CrossRef]
  30. Kumar, R.; Bawa, R.; Gahlyan, P.; Dalela, M.; Jindal, K.; Jha, P.K.; Tomar, M.; Gupta, V. Pyrene appended bis-triazolylated 1,4-dihydropyridine as a selective fluorogenic sensor for Cu2+. Dye. Pigment. 2019, 161, 162–171. [Google Scholar] [CrossRef]
  31. Feng, Y.; Yang, Y.; Wang, Y.; Qiu, F.; Song, X.; Tang, X.; Zhang, G.; Liu, W. Dual-functional colorimetric fluorescent probe for sequential Cu2+ and S2− detection in bio-imaging. Sens. Actuators B Chem. 2019, 288, 27–37. [Google Scholar] [CrossRef]
  32. Kang, D.E.; Lim, C.S.; Kim, J.Y.; Kim, E.S.; Chun, H.J.; Cho, B.R. Two-Photon Probe for Cu2+ with an Internal Reference: Quantitative Estimation of Cu2+ in Human Tissues by Two-Photon Microscopy. Anal. Chem. 2014, 86, 5353–5359. [Google Scholar] [CrossRef]
  33. Zeng, S.; Liu, X.; Kafuti, Y.S.; Kim, H.; Wang, J.; Peng, X.; Li, H.; Yoon, J. Fluorescent dyes based on rhodamine derivatives for bioimaging and therapeutics: Recent progress, challenges, and prospects. Chem. Soc. Rev. 2023, 52, 5607–5651. [Google Scholar] [CrossRef]
  34. Huang, Z.; Yang, L.; Kong, L.; Yang, J.-X. Two-photon fluorescent detection of Cu2+ in live cells through ZnS-microhybrid constructed from interfacial coordination bridge of thiocyanate. Dye. Pigment. 2020, 172, 107831. [Google Scholar] [CrossRef]
  35. Kumar, A.; Vanita, V.; Walia, A.; Chae, P.S.; Kumar, S. Pyridoanthrone-based chromo-fluorogenic amphiphiles for selective CN detection and their bioimaging application. Sens. Actuators B. Chem. 2020, 304, 127396. [Google Scholar] [CrossRef]
  36. Anbu, S.; Paul, A.; Surendranath, K.; Solaiman, N.S.; Pombeiro, A.J.L. A benzimidazole-based new fluorogenic differential/sequential chemosensor for Cu2+, Zn2+, CN, P2O74−, DNA, its live-cell imaging and pyrosequencing applications. Sens. Actuators B 2021, 337, 129785. [Google Scholar] [CrossRef]
  37. Li, D.; Liu, A.; Xing, Y.; Li, Z.; Luo, Y.; Zhao, S.; Dong, L.; Xie, T.; Guo, K.; Li, J. A smart chemosensor with different response mechanisms to multi-analytes: Chromogenic and fluorogenic recognition of Cu2+, Fe3+, and Zn2+. Dye. Pigment. 2023, 213, 111180. [Google Scholar] [CrossRef]
  38. Kucukbasmaci, H.; Aydin, D. Optical and quantitative sensing capability of phenolphthalein derived Schiff base chromo-fluorogenic sensor for Cu2+. J. Photochem. Photobiol. A 2023, 437, 114460. [Google Scholar] [CrossRef]
  39. Kumar, A.; Chae, P.S. A bis(fluorenyl-triazole)-conjugated naphthoquinoline-dione probe for a cascade detection of Cu2+ and F− and its logic circuit with a memory unit. J. Photochem. Photobiol. A 2022, 431, 114048. [Google Scholar] [CrossRef]
  40. Kumar, A.; Kumar, S.; Chae, P.S. A novel anthrapyridone diamine-based probe for selective and distinctive Cu2+ and Hg2+ sensing in aqueous solution; utility as molecular logic gates. Dye. Pigment. 2020, 181, 108522. [Google Scholar] [CrossRef]
  41. Drisya, V.; Shurooque, K.S.; Das, S.; Chakkumkumarath, L. Fmoc-phenylalanine-based fluorimetric and colorimetric dual-channel probes for the detection of Fe3+, Cu2+ and F. J. Photochem. Photobiol. A 2023, 442, 114796. [Google Scholar] [CrossRef]
  42. Sivaraman, G.; Iniya, M.; Anand, T.; Kotla, N.G.; Sunnapu, O.; Singaravadivel, S.; Gulyani, A.; Chellappa, D. Chemically diverse small molecule fluorescent chemosensors for copper ion. Coord. Chem. Rev. 2018, 357, 50–104. [Google Scholar] [CrossRef]
  43. Singh, G.; Mohit; Pawan; Sharma, S.; Devi, A.; Devi, S.; Yadav, R.; Sehgal, R.; Mohan, B. Click derived organosilane assembled with nano platform for the detection of Cu2+ ions: Biological evaluation and molecular docking approach. Spectrochim. Acta Part A 2023, 295, 122618. [Google Scholar] [CrossRef]
  44. Kumar, A.; Hur, W.; Seong, G.H.; Kumar, S.; Chae, P.S. Chromofluorogenic naphthoquinolinedione-based probes for sensitive detection and removal of Hg2+ in aqueous solutions. Dye. Pigment. 2022, 198, 110025. [Google Scholar] [CrossRef]
  45. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  46. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B Condens. Matter 1988, 37, 785–789. [Google Scholar] [CrossRef]
  47. Schmidt, M.W.; Balbridge, K.K.; Boatz, J.A.; Elbert, S.T.; Gordon, M.S.; Jensen, J.H.; Koseki, S.; Matsunaga, N.; Nguyen, K.A.; Su, S.; et al. General atomic and molecular electronic structure system. J. Comput. Chem. 1993, 14, 1347–1363. [Google Scholar] [CrossRef]
  48. Huber, R.G.; Margreiter, M.A.; Fuchs, J.E.; Grafenstein, S.V.; Tautermann, C.S.; Liedl, K.R. Heteroaromatic π-Stacking Energy Landscapes. J. Chem. Inf. Model 2014, 54, 1371–1379. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Synthetic scheme of probe 1: (a) K2CO3, DMSO, BrCH2COOtBu, TBAHSO4, RT, 2 days, and (b) BrCH2COOtBu, K2CO3, THF, TBAHSO4, RT, 3 days.
Scheme 1. Synthetic scheme of probe 1: (a) K2CO3, DMSO, BrCH2COOtBu, TBAHSO4, RT, 2 days, and (b) BrCH2COOtBu, K2CO3, THF, TBAHSO4, RT, 3 days.
Molecules 29 00808 sch001
Figure 1. (A) Partial 1H NMR spectra of probe 1 and compounds 2 and 3 with assignments of the main peaks. CDCl3 was used as the NMR solvent for probe 1 and compound 2, while DMSO-d6 was used for probe 3. (B,C) The partial 1H-1H NOESY spectrum of probe 1 dissolved in CDCl3 and (D,E) the 1H-1H COSY spectrum of probe 1 dissolved in DMSO-d6. The letters a–g in (C,E) correspond to the aromatic protons of probe 1, as indicated in the chemical structure in (B). Strong correlations between the two neighboring protons are indicated by the colored square boxes.
Figure 1. (A) Partial 1H NMR spectra of probe 1 and compounds 2 and 3 with assignments of the main peaks. CDCl3 was used as the NMR solvent for probe 1 and compound 2, while DMSO-d6 was used for probe 3. (B,C) The partial 1H-1H NOESY spectrum of probe 1 dissolved in CDCl3 and (D,E) the 1H-1H COSY spectrum of probe 1 dissolved in DMSO-d6. The letters a–g in (C,E) correspond to the aromatic protons of probe 1, as indicated in the chemical structure in (B). Strong correlations between the two neighboring protons are indicated by the colored square boxes.
Molecules 29 00808 g001
Figure 2. (A) Bar diagram showing the changes in fluorescence intensity of probe 1 (5 µM, ACN) upon the addition of different metal ions. The fluorescence intensity changes are represented as fluorescence intensity ratios (Io/I), where I and Io are the fluorescence intensities of the probe at 528 nm in the presence and absence of individual metal ions, respectively. The inset shows luminescence quenching of the probe 1 solution upon the addition of Cu2+. A Cu2+ stock solution was prepared by dissolving Cu(ClO4)2·6H2O in ACN. (B) Fluorescence spectral changes during the titration of probe 1 with Cu2+. The inset shows a linear correlation between fluorescence intensity at 528 nm (I528) and [Cu2+] in a low [Cu2+] range (0–10 µM). The points and line represent the experimental values and curved fit, respectively, where λex = 470 nm. (C) UV–Vis absorption spectral changes of probe 1 (50 µM, ACN) upon the addition of various metal ions individually. (D) Image of the probe solution showing a color change from yellow-green to pale green upon the addition of Cu2+ in daylight, as detected by the naked eye. (E) UV–Vis absorption spectral changes during the titration of probe 1 (50 µM, ACN) with Cu2+ and (F) changes in absorption values at both 476 and 413 nm (A476 and A413, respectively) with increasing [Cu2+]. The points and line represent experimental values and non-linear curve fitting, respectively.
Figure 2. (A) Bar diagram showing the changes in fluorescence intensity of probe 1 (5 µM, ACN) upon the addition of different metal ions. The fluorescence intensity changes are represented as fluorescence intensity ratios (Io/I), where I and Io are the fluorescence intensities of the probe at 528 nm in the presence and absence of individual metal ions, respectively. The inset shows luminescence quenching of the probe 1 solution upon the addition of Cu2+. A Cu2+ stock solution was prepared by dissolving Cu(ClO4)2·6H2O in ACN. (B) Fluorescence spectral changes during the titration of probe 1 with Cu2+. The inset shows a linear correlation between fluorescence intensity at 528 nm (I528) and [Cu2+] in a low [Cu2+] range (0–10 µM). The points and line represent the experimental values and curved fit, respectively, where λex = 470 nm. (C) UV–Vis absorption spectral changes of probe 1 (50 µM, ACN) upon the addition of various metal ions individually. (D) Image of the probe solution showing a color change from yellow-green to pale green upon the addition of Cu2+ in daylight, as detected by the naked eye. (E) UV–Vis absorption spectral changes during the titration of probe 1 (50 µM, ACN) with Cu2+ and (F) changes in absorption values at both 476 and 413 nm (A476 and A413, respectively) with increasing [Cu2+]. The points and line represent experimental values and non-linear curve fitting, respectively.
Molecules 29 00808 g002
Figure 3. (A,B) Luminescence image of probe 1 solutions (5 µM) prepared using the various solvents under 365 nm irradiation before and after the addition of Cu2+ (20 equiv.) and (C) changes in fluorescence emission intensity (IoI) of probe 1 (5 µM) dissolved in various solvents upon the addition of Cu2+ (20 equiv.), where λex = 470 nm. Io and I are the fluorescence intensities of probe 1 at the maximum emission wavelength before and after the addition of Cu2+, respectively. A Cu2+ stock solution was prepared by dissolving Cu(ClO4)2·6H2O in ACN.
Figure 3. (A,B) Luminescence image of probe 1 solutions (5 µM) prepared using the various solvents under 365 nm irradiation before and after the addition of Cu2+ (20 equiv.) and (C) changes in fluorescence emission intensity (IoI) of probe 1 (5 µM) dissolved in various solvents upon the addition of Cu2+ (20 equiv.), where λex = 470 nm. Io and I are the fluorescence intensities of probe 1 at the maximum emission wavelength before and after the addition of Cu2+, respectively. A Cu2+ stock solution was prepared by dissolving Cu(ClO4)2·6H2O in ACN.
Molecules 29 00808 g003
Figure 4. (A,B) Luminescence images obtained under illumination at 365 nm and of the probe 1 solution (5 µM) in a binary mixture of ACN and water. Luminescence images of the probe 1 solution in the absence (A) and presence of Cu2+ (20 equiv.) A Cu2+ stock solution was prepared by dissolving Cu(ClO4)2·6H2O in water. A, W, and the numbers in the vial designations indicate ACN, water, and water percentages in the solutions, respectively, where λex = 470 nm. (C) Changes in fluorescence emission intensity (IoI) of the probe 1 solution (5 µM) at the maximum emission wavelength (λem) upon the addition of 20 equiv. of Cu2+. Io and I represent the fluorescence intensities of the probe at λem before and after the addition of Cu2+ ions, respectively.
Figure 4. (A,B) Luminescence images obtained under illumination at 365 nm and of the probe 1 solution (5 µM) in a binary mixture of ACN and water. Luminescence images of the probe 1 solution in the absence (A) and presence of Cu2+ (20 equiv.) A Cu2+ stock solution was prepared by dissolving Cu(ClO4)2·6H2O in water. A, W, and the numbers in the vial designations indicate ACN, water, and water percentages in the solutions, respectively, where λex = 470 nm. (C) Changes in fluorescence emission intensity (IoI) of the probe 1 solution (5 µM) at the maximum emission wavelength (λem) upon the addition of 20 equiv. of Cu2+. Io and I represent the fluorescence intensities of the probe at λem before and after the addition of Cu2+ ions, respectively.
Molecules 29 00808 g004
Figure 5. (A) Bar diagram showing changes in fluorescence intensity of probe 1 (5 µM, HEPES (pH 7.4): ACN = 1:4)) upon individual additions of various metal ions. The inset shows the solution color in the absence and presence of Cu2+. A Cu2+ stock solution was prepared by dissolving Cu(ClO4)2·6H2O in water. (B) Fluorescence titration of probe 1 with Cu2+ when excited at 470 nm. The inset shows a linear correlation between I550 and [Cu2+] at a low metal ion concentration range (0–10 µM).
Figure 5. (A) Bar diagram showing changes in fluorescence intensity of probe 1 (5 µM, HEPES (pH 7.4): ACN = 1:4)) upon individual additions of various metal ions. The inset shows the solution color in the absence and presence of Cu2+. A Cu2+ stock solution was prepared by dissolving Cu(ClO4)2·6H2O in water. (B) Fluorescence titration of probe 1 with Cu2+ when excited at 470 nm. The inset shows a linear correlation between I550 and [Cu2+] at a low metal ion concentration range (0–10 µM).
Molecules 29 00808 g005
Figure 6. Changes in 1H NMR spectra (partial) of probe 1 during titration with Cu2+(ClO4)2·6H2O in DMSO-d6. The letters (a–g) represent the aromatic protons of probe 1, as indicated in the chemical structure).
Figure 6. Changes in 1H NMR spectra (partial) of probe 1 during titration with Cu2+(ClO4)2·6H2O in DMSO-d6. The letters (a–g) represent the aromatic protons of probe 1, as indicated in the chemical structure).
Molecules 29 00808 g006
Scheme 2. Schematic representation of a plausible mechanism responsible for the fluorescence quenching of probe 1 upon interaction with Cu2+. The Cu2+ binding to the probe allows electron transfer from the naphthoquinoline-dione ring to the metal binding site, leading to fluorescence quenching.
Scheme 2. Schematic representation of a plausible mechanism responsible for the fluorescence quenching of probe 1 upon interaction with Cu2+. The Cu2+ binding to the probe allows electron transfer from the naphthoquinoline-dione ring to the metal binding site, leading to fluorescence quenching.
Molecules 29 00808 sch002
Figure 7. (A) Energy-minimized structures of probe 1 and the [1-Cu2+] complex in a ball and spoke model and (B) electronic distributions and energies of their frontier molecular orbitals (FMOs; HOMOs and LUMOs) obtained from DFT calculations at the B3LYP/6-31G* level. Individual atoms are color coded for a clear view: N = sky blue (cyan), O = red, C = yellow, and H = light purple. The letters (a-d) represent individual bond lengths between the metal ion and the chelating atoms.
Figure 7. (A) Energy-minimized structures of probe 1 and the [1-Cu2+] complex in a ball and spoke model and (B) electronic distributions and energies of their frontier molecular orbitals (FMOs; HOMOs and LUMOs) obtained from DFT calculations at the B3LYP/6-31G* level. Individual atoms are color coded for a clear view: N = sky blue (cyan), O = red, C = yellow, and H = light purple. The letters (a-d) represent individual bond lengths between the metal ion and the chelating atoms.
Molecules 29 00808 g007
Figure 8. (A) Quenching effect of Cu2+ on the luminescence of ‘Cu2+’ written on the TLC plate using probe 1 (5 µM, ACN) solution. (B,C) Quenching effect of various metal ions on the luminescence of probe 1-coated cotton buds (B) and silica gels placed in the wells of ELISA plates (C). The probe-coated cotton buds and silica gels were prepared by applying the probe 1 solution (5 μM) onto the cotton buds and silica gels. (D,E) Quenching effect of Cu2+ on the luminescent spot of probe 1 (5 μM) on the TLC plate (D) and Whatman filter paper (E). Different amounts of metal ions were applied to the spots, and the images were obtained under 365 nm illumination.
Figure 8. (A) Quenching effect of Cu2+ on the luminescence of ‘Cu2+’ written on the TLC plate using probe 1 (5 µM, ACN) solution. (B,C) Quenching effect of various metal ions on the luminescence of probe 1-coated cotton buds (B) and silica gels placed in the wells of ELISA plates (C). The probe-coated cotton buds and silica gels were prepared by applying the probe 1 solution (5 μM) onto the cotton buds and silica gels. (D,E) Quenching effect of Cu2+ on the luminescent spot of probe 1 (5 μM) on the TLC plate (D) and Whatman filter paper (E). Different amounts of metal ions were applied to the spots, and the images were obtained under 365 nm illumination.
Molecules 29 00808 g008
Figure 9. Confocal microscopic images of HeLa cells loaded with probe 1 in the absence (top) and presence of Cu2+ (20 µM, bottom). The HeLa cells were preloaded with probe 1 (5 µM) for 30 min and then treated with Cu2+ (20 µM), followed by the acquisition of fluorescence images under excitation at 488 nm. The probe stock solution is prepared using DMSO, and the final DMSO concentration was 0.5% in the sample medium.
Figure 9. Confocal microscopic images of HeLa cells loaded with probe 1 in the absence (top) and presence of Cu2+ (20 µM, bottom). The HeLa cells were preloaded with probe 1 (5 µM) for 30 min and then treated with Cu2+ (20 µM), followed by the acquisition of fluorescence images under excitation at 488 nm. The probe stock solution is prepared using DMSO, and the final DMSO concentration was 0.5% in the sample medium.
Molecules 29 00808 g009
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

Kumar, A.; Chae, P.S. A Naphthoquinoline-Dione-Based Cu2+ Sensing Probe with Visible Color Change and Fluorescence Quenching in an Aqueous Organic Solution. Molecules 2024, 29, 808. https://doi.org/10.3390/molecules29040808

AMA Style

Kumar A, Chae PS. A Naphthoquinoline-Dione-Based Cu2+ Sensing Probe with Visible Color Change and Fluorescence Quenching in an Aqueous Organic Solution. Molecules. 2024; 29(4):808. https://doi.org/10.3390/molecules29040808

Chicago/Turabian Style

Kumar, Ashwani, and Pil Seok Chae. 2024. "A Naphthoquinoline-Dione-Based Cu2+ Sensing Probe with Visible Color Change and Fluorescence Quenching in an Aqueous Organic Solution" Molecules 29, no. 4: 808. https://doi.org/10.3390/molecules29040808

Article Metrics

Back to TopTop