Next Article in Journal
Identification of New Regulators of Pancreatic Cancer Cell Sensitivity to Oxaliplatin and Cisplatin
Next Article in Special Issue
Sensitive Assay for the Lactonase Activity of Serum Paraoxonase 1 (PON1) by Harnessing the Fluorescence Turn-On Characteristics of Bioorthogonally Synthesized and Geometrically Controlled Chemical Probes
Previous Article in Journal
The Difference of Volatile Compounds in Female and Male Buds of Herpetospermum pedunculosum Based on HS-SPME-GC-MS and Multivariate Statistical Analysis
Previous Article in Special Issue
Novel Boronate Probe Based on 3-Benzothiazol-2-yl-7-hydroxy-chromen-2-one for the Detection of Peroxynitrite and Hypochlorite
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

High-Affinity Ratiometric Fluorescence Probe Based on 6-Amino-2,2′-Bipyridine Scaffold for Endogenous Zn2+ and Its Application to Living Cells

1
Laboratory of Analytical Chemistry, Faculty of Pharmaceutical Sciences, Mukogawa Women’s University, 11-68 Koshien Kyubancho, Nishinomiya 663-8179, Japan
2
Division of Medical Innovation, Translational Research Center for Medical Innovation, 1-5-4 Minatojima-minamimachi, Chuo-ku, Kobe 650-0047, Japan
3
Graduate School of Science and Engineering, Saitama University, 255 Shimo-Okubo, Sakura-ku, Saitama 338-8570, Japan
4
Graduate School of Pharmaceutical Sciences, Kyoto University, 46-29 Yoshida-Shimoadachi-cho, Sakyo-ku, Kyoto 606-8501, Japan
5
Laboratory of Biophysical Chemistry, Kobe Pharmaceutical University, 4-19-1 Motoyamakita Machi, Higashinada-ku, Kobe 658-8558, Japan
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(4), 1287; https://doi.org/10.3390/molecules27041287
Submission received: 12 January 2022 / Revised: 1 February 2022 / Accepted: 12 February 2022 / Published: 14 February 2022
(This article belongs to the Special Issue Fluorescent Probes as Powerful Tools in Medicinal Chemistry)

Abstract

:
Zinc is an essential trace element involved in many biological activities; however, its functions are not fully understood. To elucidate the role of endogenous labile Zn2+, we developed a novel ratiometric fluorescence probe, 5-(4-methoxyphenyl)-4-(methylsulfanyl)-[2,2′-bipyridin]-6-amine (6 (rBpyZ)) based on the 6-amino-2,2′-bipyridine scaffold, which acts as both the chelating agent for Zn2+ and the fluorescent moiety. The methoxy group acted as an electron donor, enabling the intramolecular charge transfer state of 6 (rBpyZ), and a ratiometric fluorescence response consisting of a decrease at the emission wavelength of 438 nm and a corresponding increase at the emission wavelength of 465 nm was observed. The ratiometric probe 6 (rBpyZ) exhibited a nanomolar-level dissociation constant (Kd = 0.77 nM), a large Stokes shift (139 nm), and an excellent detection limit (0.10 nM) under physiological conditions. Moreover, fluorescence imaging using A549 human lung adenocarcinoma cells revealed that 6 (rBpyZ) had good cell membrane permeability and could clearly visualize endogenous labile Zn2+. These results suggest that the ratiometric fluorescence probe 6 (rBpyZ) has considerable potential as a valuable tool for understanding the role of Zn2+ in living systems.

1. Introduction

Zinc is an essential trace element for living organisms and plays pivotal roles in many biological processes, such as gene expression, apoptosis, enzyme regulation, and neurotransmission [1,2,3,4]. Most zinc is tightly bound to proteins and is involved in structural and catalytic functions [2,3]. In contrast, labile (free) zinc exits at low concentrations (picomolar to nanomolar) and is reported to play important roles in signaling, cell proliferation, and differentiation [4,5,6]. The amount of labile zinc present is regulated by zinc homeostasis; however, it has been suggested that zinc deficiency induced by insufficient dietary intake, poor absorption, and heredity disorders is related to hypogeusia, growth retardation, and carcinogenesis [7,8,9]. Therefore, to elucidate the details of the function and behavior of zinc in cells and tissues, a reliable method is urgently required.
Biological imaging using fluorescence probes is a robust method that enables the sensitive, real-time, and noninvasive detection of target molecules. Organic molecule-based fluorescence probes are gaining importance as simple cost-effective probes with good biocompatibility [10,11]. Their properties, including target binding ability and fluorescence response, are relatively easy to optimize by modifying their molecular structure and substituents [10,11]. As zinc has a closed shell structure (3d104s0) and cannot be directly analyzed using conventional methods, such as ultraviolet/visible (UV/vis) and nuclear magnetic resonance (NMR) spectroscopies, the development of organic molecule-based fluorescence probes for zinc has attracted attention [12,13]. A variety of fluorescence Zn2+ probes based on fluorophore platforms, such as quinoline, fluorescein, cyanine, and biheteroaryl have been reported, some of which have contributed to the understanding of the biological function and behavior of zinc [12,13,14,15,16]. However, even with these probes, the detailed biological functions of zinc are yet to be fully understood. Therefore, the development of novel fluorescence Zn2+ probes that demonstrate higher affinity toward Zn2+ as well as increased solubility and cell membrane permeability than the currently available probes is essential. Previously, we reported a 6-amino-2,2′-bipyridine-based fluorescence Zn2+ probe 3, in which the core structure acted as both the chelating agent for Zn2+ and the fluorescent moiety. The resulting probe had a small molecular weight, making it highly soluble and cell membrane-permeant [17]. This probe also achieved a nanomolar-range dissociation constant toward Zn2+ by controlling intramolecular amino-imino tautomerization; however, the high background fluorescence of the probe was a significant drawback for detecting endogenous Zn2+ [17]. The use of ratiometric probes enables highly sensitive measurements through self-calibration at different emission wavelengths in the presence and absence of analyte [18,19]. Generally, the ratiometric property is based on forming an intramolecular charge transfer (ICT) state. This is generally achieved by complex electron interaction between an electron donor and an electron acceptor within molecules [18]. In this manner, the ratiometric probe can minimize the effect of its background fluorescence, which allows for the accurate detection of analytes, including Zn2+.
We report a novel ratiometric Zn2+ probe based on the 6-amino-2,2′-bipyridine scaffold. The introduction of an electron-donating or electron-withdrawing group significantly affected the fluorescence property of the 6-amino-2,2′-bipyridine-based fluorescence probe, and 5-(4-methoxyphenyl)-4-(methylsulfanyl)-[2,2′-bipyridin]-6-amine, 6 (rBpyZ), exhibited a favorable ratiometric property and high binding affinity toward Zn2+. In this study, we demonstrate the design, synthesis, and fluorescence properties of 6 (rBpyZ) and evaluate its usefulness in detecting endogenous Zn2+ through fluorescence imaging using A549 human lung adenocarcinoma cells.

2. Results and Discussion

2.1. Synthesis of 6-Amino-2,2′-Bipyridine-Based Fluorescence Compounds

The 6-Amino-2,2′-bipyridine-based compounds 4–6 were synthesized using a one-pot synthesis method we developed for 3 [17], involving the reaction of 3,3-bis(methylsulfanyl)-1-(pyridin-2-yl)prop-2-en-1-one (1) with phenyl acetonitriles (2b–d) in the presence of sodium hydroxide, followed by treatment with an ammonia solution (Scheme 1). Compared to the yield of the unsubstituted compound 3 (28%), those of 4 and 5, which bear electron-withdrawing groups (Cl and CF3) on the allyl group at position 5 of the pyridine ring, decreased to 15% and 5%, respectively. In contrast, the yield of 6 (rBpyZ), which bears an electron-donating group (OMe), increased to 44%. As these reactions are initiated by the nucleophilic attack of phenyl acetonitriles on 1, it is likely that the electron density of the substituent in phenyl acetonitriles has a significant effect on its nucleophilicity, resulting in the difference in yields.

2.2. Spectroscopic Studies

We analyzed the UV/Vis absorption and fluorescence properties of 46 (rBpyZ) in HEPES buffer (100 mM, 50% EtOH, pH = 7.4) and compared them with those of compound 3. As shown in Figure 1, the absorption peaks at wavelengths 260 and 345 nm of free 3 gradually increased with the addition of Zn2+, accompanied by a small bathochromic shift of 5–7 nm (Figure 1a). Compounds 4 and 5 also exhibited absorption maxima at 260 and 345 nm, and similar changes to 3 were observed after adding Zn2+ (Figure 1b,c). In compound 6 (rBpyZ), the absorption peak at the longer wavelength exhibited a hypsochromic shift of 20 nm compared to that of 3, and a new peak appeared at 349 nm with the addition of Zn2+ (Figure 1d).
Figure 2 shows the fluorescence titration spectra with Zn2+. Similar to the spectrum of 3, the fluorescence intensities of 4 and 5 gradually increased in a Zn2+ concentration-dependent manner, and the maximum emission wavelength of 4 and 5 exhibited small bathochromic shifts of 14 and 5 nm, respectively. The fluorescence intensities of 4 and 5 were stronger than that of 3, being enhanced 10- and 18-fold, respectively.
In contrast, a significant ratiometric response to Zn2+ was observed in compound 6 (rBpyZ) (Figure 2d). In the absence of Zn2+, 6 (rBpyZ) exhibited an emission peak at 438 nm, which was blue-shifted by 30 nm compared to those of compounds 4 and 5. Upon adding Zn2+, the emission peak at 438 nm gradually decreased, and a new emission peak at 465 nm concomitantly appeared (Figure 3). In addition, 6 (rBpyZ) showed a significantly larger Stokes shift (139 nm), which enabled sensitive detection without interference from the excitation light wavelength.
These results indicated that a chelation-enhanced fluorescence (CHEF) effect occurred with the addition of Zn2+, and the simple replacement of substituent influenced the fluorescence emission profile. The analysis of all compounds revealed that the complex formed between the compound and Zn2+ possessed a 1:1 stoichiometry (Figure S1, Supplementary Materials). From the fluorescence titration data, the dissociation constants (Kd) of 46 (rBpyZ) were calculated to be at a nanomolar level, comparable to that of compound 3 (2.2 nM) (Table 1; Figure S2).
We previously reported on the properties of the Zn2+-coordinated bipyridine moiety of 3 and its structural analogs using 1H-NMR [17]. As 6 (rBpyZ) exhibited a fluorescence response distinct from 3–5, we investigated the Zn2+ coordination site of 6 (rBpyZ) using 1H-NMR spectroscopy. As shown in Figure 4, the protons of the 3- and 3′-positions of 6 (rBpyZ) shifted 0.04 or 0.9 ppm downfield when 1.0 equiv. of Zn2+ was added. The NH2 proton peak at 5.26 ppm moved to 6.29 ppm without any change in integral values. These results were similar to the NMR results of 3, indicating that the Zn2+ coordinated with the 2,2′-bipyridine moiety of 6 (rBpyZ) and that amino-imino tautomerism did not occur. These results also suggested that the ratiometric property of 6 (rBpyZ) was not due to a change of the Zn2+ coordination site but was rather caused by the stabilization of the highest occupied molecular orbital (HOMO) energy via the introduction of an electron-donating group, similar to that observed in other ratiometric probes [20]. Ratiometric probes that emit at two distinct wavelengths are beneficial for the sensitive and selective detection of targets. Subsequently, evaluations including a metal selectivity experiment, a test of the influence of pH on fluorescence, and the analysis of the fluorescence imaging of cells, were performed on the ratiometric compound 6 (rBpyZ).
To investigate the selectivity of 6 (rBpyZ) toward cations, the fluorescence spectra of 6 (rBpyZ) in the presence of various cations (Al3+, Ca2+, Cd2+, Co2+, Cu2+, Fe2+, Fe3+, K+, Mg2+, Mn2+, Na+, Ni2+, and Zn2+) were recorded, and are displayed in Figure 5. We note that the addition of Cd2+ to 6 yields a similarly enhanced fluorescence spectrum to that produced with the addition of Zn2+. It has been reported that various Zn2+ fluorescence probes also detect Cd2+ owing to both ions being in the same column of the periodic table [12]. However, the concentration of free Cd2+ is very low in living systems and can therefore be assumed to have a negligible influence on cellular Zn2+ imaging. The addition of alkali ions (Na+; K+) and alkaline-earth ions (Ca2+; Mg2+), which exist ubiquitously at millimolar concentrations in living systems, cause no change to the spectrum of 6 (rBpyZ). In the transition metal cations (Co2+, Cu2+, Fe2+, Fe3+, Mn2+, Ni2+, and Al3+), fluorescence quenching and small bathochromic shifts in the emission wavelength were observed, indicating that these cations also form a complex with 6 (rBpyZ). However, binary competition experiments of 6 (rBpyZ) between Zn2+ and other cations, using the ratio of fluorescence intensity at 465 nm to that at 438 nm (F465/F438) (Figure 6), showed that in the presence of Co2+, Fe2+, Fe3+, Mn2+, and Al3+, 6 (rBpyZ) selectively detects Zn2+. 6 (rBpyZ) is not selective for Zn2+ when in the presence of either Cu2+ or Ni2+. Still, this lack of preference may have little impact on visualizing cellular Zn2+ as the cellular concentrations of free Cu2+ and Ni2+ are also very low.
Figure 7 shows the effects of pH on the Zn2+ detection ability of 6 (rBpyZ). F465/F438 of 6 (rBpyZ) was measured in buffer solutions at different pH values in the absence and presence of Zn2+. Under both conditions, F465/F438 was approximately constant between pH 5.0–8.0. In addition, a limit of detection (LOD) of 0.10 nM at pH 7.0 buffer solution was calculated using the equation LOD = 3σ/slope. These results indicated that 6 (rBpyZ) could function as a sensitive Zn2+ probe under physiological conditions. We subsequently explored its usefulness in cellular applications, which we report in the next section.

2.3. Zn2+ Cellular Imaging

Figure 8 shows fluorescence microscopy images of human lung adenocarcinoma cells (A549). The cells were exposed to 100 µM 6 (rBpyZ) for 30 min at 37 °C, which produced very weak fluorescence (Figure 8a). In contrast, the cells incubated with 100 µM 6 (rBpyZ) for 30 min and 100 µM Zn2+ for another 30 min exhibited a bright fluorescence signal in the cells (Figure 8b). The fluorescence signal was reduced by treatment with a cell membrane permeable Zn2+ chelator, N,N,N’,N’-tetrakis(2-pyridylmethyl)ethylenediamine (TPEN) [21], indicating that 6 (rBpyZ) exhibited fluorescence in response to intracellular Zn2+ (Figure 8c). In addition, no significant cytotoxicity was observed during this study.
Therefore, we further investigated the possibility of detecting endogenous Zn2+ in the apoptotic cells. Zn2+ is known as a regulator of apoptosis and is released from intracellular stores during apoptosis [22]. As shown in Figure 9, in the H2O2-induced apoptotic cells, a bright fluorescence signal was observed after incubation with 100 µM 6 (rBpyZ). It was confirmed that this observed fluorescence signal does not originate from the direct effect of H2O2 on 6 (rBpyZ) (Figure S3). These data indicated that ratiometric probe 6 (rBpyZ) could detect endogenous Zn2+.

3. Materials and Methods

3.1. Materials and Instruments

All reagents and chemicals were of the highest grade available. 1H and 13C-NMR spectra were recorded on JEOL ECP-400 NMR and Varian Gemini 300 systems at room temperature. The chemical shifts are reported in parts per million (ppm) relative to tetramethylsilane (TMS) as the reference. Mass spectra (MS) and high-resolution mass spectrometry (HRMS) were performed using a JMS-700 spectrometer (JEOL, Tokyo, Japan). UV/Vis absorption spectra were acquired with a UV/Vis spectrophotometer (UV-2450, Shimadzu, Kyoto, Japan). Fluorescence spectra were acquired using a fluorescence spectrophotometer (FP-8300ST, Jasco, Tokyo, Japan).

3.2. Synthesis of 4-(Methylsulfanyl)-5-Phenyl-[2,2′-Bipyridin]-6-Amine (3)

To a solution of 3,3-bis(methylsulfanyl)-2-(phenylsulfonyl)acrylonitrile (1) (560 mg, 2.5 mmol) and 2-phenylacetonitrile (2a) (439 mg, 3.8 mmol) in DMSO (20 mL), powdered sodium hydroxide (200 mg, 5.0 mmol) was added, and the mixture was stirred for 2 h at room temperature. The reaction solution was poured into 200 mL of ice water and neutralized with a 10% hydrochloric acid solution. The mixture was extracted with chloroform (3 × 50 mL), washed with brine (50 mL), and dried over anhydrous sodium sulfate. After concentration under reduced pressure conditions, a mixture of the residue and 150 mL of 28% aqueous ammonia was refluxed for 1 h. After the evaporation of the ammonia and water, the residue was chromatographed on a silica gel column using toluene as an eluent to obtain 205 mg (0.7 mmol, 28%) of 3 as colorless needles. Mp: 119–120 °C. 1H-NMR (DMSO-d6, 400 MHz) δ 2.50 (s, 3H), 5.26 (brs, 2H), 7.27 (d, J = 7.3 Hz, 2H), 7.41 (m, 2H), 7.50 (dd, J = 7.3, 7.8 Hz, 2H), 7.70 (s, 1H), 7.91 (dd, J = 7.8, 7.8 Hz, 1H), 8.30 (d, J = 7.8 Hz, 1H), 8.67 (d, J = 4.4 Hz, 1H). 13C-NMR (DMSO-d6, 100 MHz) δ 13.9, 104.8, 117.5, 120.4, 123.9, 128.4, 129.3, 129.9, 129.9, 135.3, 137.0, 149.0, 149.7, 152.4, 155.4, 155.8. MS m/z: 294 [M + H+]. HRMS Calcd for C17H16N3S [M + H+]: 294.1065. Found: 294.1040.

3.3. Synthesis of 5-(4-Chlorophenyl)-4-(Methylsulfanyl)-[2,2′-Bipyridin]-6-Amine (4)

Compound 4 was prepared from 1 (565 mg, 2.5 mmol) and 2-(4-chlorophenyl)acetonitrile (2b) (574 mg, 3.8 mmol) in a manner similar to that described for the synthesis of 3. The residue was chromatographed on a silica gel column using toluene as an eluent to obtain 121 mg (0.4 mmol, 15%) of 4 as colorless needles. Mp: 243–244 °C. 1H-NMR (DMSO-d6, 400 MHz) δ 2.50 (s, 3H), 5.41 (brs, 2H), 7.27 (d, J = 8.4 Hz, 2H), 7.47–7.49 (m, 2H), 7.49 (dd, J = 8.4 Hz, 2H), 7.65 (s, 1H), 7.99 (dd, J = 6.8, 8.0 Hz, 1H), 8.26 (d, J = 8.4 Hz, 1H), 8.74 (d, J = 4.4 Hz, 1H). 13C-NMR (DMSO-d6, 100 MHz) δ 14.8, 120.4, 121.1, 124.9, 128.1, 129.3, 132.0, 132.1, 132.2, 133.9, 137.0, 137.7, 149.5, 151.9, 152.7, 159.4. MS m/z: 328 [M + H+]. HRMS Calcd for C17H15ClN3S [M + H+]: 328.0675. Found: 328.0610.

3.4. Synthesis of 4-(Methylsulfanyl)-5-(4-(Trifluoromethyl)Phenyl)-[2,2′-Bipyridin]-6-Amine (5)

Compound 5 was prepared from 1 (565 mg, 2.5 mmol) and 2-(4-(trifluoromethyl)phenyl)acetonitrile (2c) (693 mg, 3.8 mmol) in a manner similar to that described for the synthesis of 3. The residue was chromatographed on a silica gel column using toluene as an eluent to obtain 42 mg (0.1 mmol, 5%) of 5 as colorless needles. Mp: 228–229 °C. 1H-NMR (CDCl3, 300 MHz) δ 2.49 (s, 3H), 6.89 (brs, 2H), 7.37–7.42 (m, 2H), 7.53 (d, J = 7.84 Hz, 2H), 7.69 (d, J = 7.8 Hz, 2H), 7.79 (m, 3H), 7.91 (d, J = 8.4 Hz, 1H), 8.66 (d, J = 5.4 Hz, 1H). 13C-NMR (DMSO-d6, 75 MHz) δ 15.5, 100.2, 119.9, 122.6, 125.1, 125.5, 125.6, 126.2, 130.0, 130.4, 130.9, 137.6, 138.6, 140.7, 147.8, 149.7, 152.8, 160.1. MS m/z: 362 [M + H+]. HRMS Calcd for C18H15F3N3S [M + H+]: 362.0938. Found: 362.0902.

3.5. Synthesis of 5-(4-Methoxyphenyl)-4-(Methylsulfanyl)-[2,2′-Bipyridin]-6-Amine (6, rBpyZ)

Compound 6 (rBpyZ) was prepared from 1 (565 mg, 2.5 mmol) and 2-(4-2-(4-methoxyphenyl)acetonitrile (2d) (551 mg, 3.8 mmol) in a manner similar to that described for the synthesis of 3. The residue was chromatographed on a silica gel column using toluene as an eluent to obtain 364 mg (1.1 mmol, 44%) of 6 (rBpyZ) as colorless needles. Mp 233–234 °C. 1H-NMR (CDCl3, 300 MHz) δ 2.45 (s, 3H), 3.84 (s, 3H), 4.35 (brs, 2H), 7.02 (d, J = 8.7 Hz, 2H), 7.25–7.30 (m, 1H), 7.34 (d, J = 8.7 Hz, 2H), 7.72 (s, 1H), 7.76–7.85 (m, 1H), 8.30 (d, J = 8.1 Hz, 1H), 8.65 (m, 1H). 13C-NMR (CDCl3, 75 MHz) δ 14.9, 55.5, 100.0, 107.0, 114.1, 115.1, 121.3, 123.6, 127.5, 131.4, 131.5, 149.2, 151.4, 153.4, 155.7, 156.5, 159.8. MS m/z: 324 [M + H+]. HRMS calcd for C18H17N3OS [M + H+]: 324.1170. Found: 324.1195.

3.6. Spectroscopic Studies

The solutions of perchlorate salts of metal ions (Na+, K+, Mg2+, Ca2+, Fe2+, Fe3+, Ni2+, Zn2+, Cd2+, Co2+, Cu2+, Mn2+, and Al3+) were prepared by dissolving them in deionized water. To prepare compound stock solutions (1 × 10−2 M), each compound was dissolved in DMSO. The UV/Vis and fluorescence analysis of each compound was performed in the absence and presence of metal ions in EtOH/H2O (1:1, v/v). Job plot experiments were performed to determine the complex stoichiometry of 4–6 (rBpyZ) to Zn2+. The dissociation constants (Kd) were determined using a 0–9 mM ZnSO4/10 mM nitrilotriacetic acid (NTA) system and Benesi–Hildebrand plots [23,24,25]. The metal ion selectivity of 6 (rBpyZ) was investigated using 10−2 M metal cation solutions. The effects of pH change on the fluorescence of 6 (rBpyZ) were evaluated using various buffers: citrate buffer (pH 3.0–6.0), HEPES buffer (pH 7.0–8.0), and tris-hydrochloric acid buffer (pH 9.0).

3.7. Microscope Imaging of Zinc Fluorescence in Cells

The cell line A549 human lung adenocarcinoma from RIKEN BRC (Ibaraki, Japan) was cultured in Dulbecco’s modified Eagle’s medium (DMEM) supplemented with 10% fetal bovine serum and 1% penicillin. Cells were incubated at 37 °C in a humidified atmosphere containing 5% CO2. To evaluate the cell membrane permeability of 6 (rBpyZ), the cells were incubated with 6 (rBpyZ) (100 μM) for 30 min at 37 °C, washed with phosphate-buffered saline (PBS), and incubated with Zn2+/pyrithione (100 μM) for 30 min at 37 °C. For TPEN studies, the cells were incubated with TPEN (100 μM) for an additional 30 min at 37 °C after incubation with Zn2+/pyrithione (100 μM). To detect endogenous Zn2+ in A549 cells, the cells were incubated with hydrogen peroxide (100 µM) in DMEM for 24 h at 37 °C. After washing with PBS, the treated cells were incubated with 6 (rBpyZ) (100 μM) for 30 min at 37 °C. The fluorescence images of the incubated cells were acquired using a fluorescence microscope BZ-X710 (Keyence, Osaka, Japan).

4. Conclusions

To reduce the influence of background fluorescence and achieve the sensitive detection of endogenous Zn2+ in cells, we designed and synthesized 6-amino-2,2′-bipyridine-based small molecular weight Zn2+ fluorescence compounds with electron-withdrawing or electron-donating groups on the allyl group at position 5 of the central pyridine ring. Among the synthesized compounds, compound 6 (rBpyZ) bearing a methoxy group exhibited ratiometric fluorescence profiles at 438 nm and 465 nm in the absence or presence of Zn2+. In addition, 6 (rBpyZ) showed a high Zn2+ binding affinity (Kd = 0.77 nM), a large Stokes shift (over 139 nm), high Zn2+ selectivity, and high stability under physiological pH conditions, making it useful as a fluorescence Zn2+ probe. Fluorescence microscopy studies indicated that 6 (rBpyZ) possessed cell membrane permeability in living human lung adenocarcinoma cells and enabled the visualization of endogenous labile Zn2+ in the cells during apoptosis. We expect that 6 (rBpyZ) will be a valuable and efficient probe for elucidating the biological functions of Zn2+.

Supplementary Materials

The following are available online: Figure S1: Job’ plot analysis of 4, 5, and 6 (rBpyZ); Figure S2: Benesi–Hildebrand analysis of 4, 5, and 6 (rBpyZ); Figure S3: Fluorescence spectra of 6 (rBpyZ) upon the addition of H2O2. 1H- and 13C-NMR spectra are available online.

Author Contributions

Conceptualization, M.H. and T.M.; data curation, M.H., N.M. and F.H.; formal analysis; M.H., N.M. and F.H.; investigation, M.H., N.M., F.H. and T.F.; writing—original draft preparation, M.H.; writing, reviewing, and editing, M.H., N.M., T.F., H.S. and T.M.; supervision, H.S. and T.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was partly funded by the Konica Minolta Imaging Science Encouragement Award of Konica Minolta Science and Technology Foundation and JSPS KAKENHI grant number 20K08139.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the present article and Supplementary Files.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Yang, T.C.; Wu, P.C.; Chung, I.F.; Jiang, J.H.; Fann, M.J.; Kao, L.S. Cell death caused by the synergistic effects of zinc and dopamine is mediated by a stress sensor gene Gadd45b—Implication in the pathogenesis of Parkinson’s disease. J. Neurochem. 2016, 139, 120–133. [Google Scholar] [CrossRef] [PubMed]
  2. Zhu, B.; Wang, J.; Zhou, F.; Liu, Y.; Lai, Y.; Wang, J.; Chen, X.; Chen, D.; Luo, L.; Hua, Z.-C. Zinc depletion by TPEN induces apoptosis in human acute promyelocytic NB4 cells. Cell Physiol. Biochem. 2017, 42, 1822–1836. [Google Scholar] [CrossRef] [PubMed]
  3. Bafaro, E.; Liu, Y.; Xu, Y.; Dempski, R.E. The emerging role of zinc transporters in cellular homeostasis and cancer. Signal Transduct. Target. Ther. 2017, 2, e17029. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Tóth, K. Zinc in neurotransmission. Annu. Rev. Nutr. 2011, 31, 139–153. [Google Scholar] [CrossRef] [PubMed]
  5. Turan, B.; Tuncay, E. Impact of labile zinc on heart function: From physiology to pathophysiology. Int. J. Mol. Sci. 2017, 18, 2395. [Google Scholar] [CrossRef] [Green Version]
  6. Kim, B.; Lee, W.W. Regulatory role of zinc in immune cell signaling. Mol. Cells 2021, 44, 335–341. [Google Scholar] [CrossRef]
  7. Yanagisawa, H. Zinc deficiency and clinical practice—Validity of zinc preparations. Yakugaku Zasshi 2008, 128, 333–339. [Google Scholar] [CrossRef] [Green Version]
  8. Kolenko, V.; Teper, E.; Kutikov, A.; Uzzo, R. Zinc and zinc transporters in prostate carcinogenesis. Nat. Rev. Urol. 2013, 10, 219–226. [Google Scholar] [CrossRef] [Green Version]
  9. Abdollahi, M.; Ajami, M.; Abdollahi, Z.; Kalantari, N.; Houshiarrad, A.; Fozouni, F.; Fallahrokni, A.; Mazandaranid, F.S. Zinc supplementation is an effective and feasible strategy to prevent growth retardation in 6 to 24 month children: A pragmatic double blind, randomized trial. Heliyon 2019, 5, e02581. [Google Scholar] [CrossRef] [Green Version]
  10. 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]
  11. Rao, J.; Dragulescu-Andrasi, A.; Yao, H. Fluorescence imaging in vivo: Recentadvances. Curr. Opin. Biotechnol. 2007, 18, 17–25. [Google Scholar] [CrossRef] [PubMed]
  12. Nolan, E.M.; Lippard, S.J. Small-molecule fluorescent sensors for investigating zinc metalloneurochemistry. Acc. Chem. Res. 2010, 42, 193–203. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Li, J.; Yin, C.; Huo, F. Development of fluorescent zinc chemosensors based on various fluorophores and their applications in zinc recognition. Dyes Pigm. 2016, 131, 100–133. [Google Scholar] [CrossRef]
  14. Yin, J.; Kwon, Y.; Kim, D.; Lee, D.; Kim, G.; Hu, Y.; Ryu, J.H.; Yoon, J. Cyanine-based fluorescent probe for highly selective detection of glutathione in cell cultures and live mouse tissues. J. Am. Chem. Soc. 2014, 136, 5351–5358. [Google Scholar] [CrossRef] [PubMed]
  15. Hagimori, M.; Mizuyama, N.; Tominaga, Y.; Mukai, T.; Saji, H. A low-molecular-weight fluorescent sensor with Zn2+ dependent bathochromic shift of emission wavelength and its imaging in living cells. Dyes Pigm. 2015, 113, 205–209. [Google Scholar] [CrossRef]
  16. Hagimori, M.; Taniura, M.; Mizuyama, N.; Karimine, Y.; Kawakami, S.; Saji, H.; Mukai, T. Synthesis of a novel pyrazine–pyridone biheteroaryl-based fluorescence sensor and detection of endogenous labile zinc Ions in lung cancer cells. Sensors 2019, 19, 2049. [Google Scholar] [CrossRef]
  17. Hagimori, M.; Temma, T.; Mizuyama, N.; Uto, T.; Yamaguchi, Y.; Tominaga, Y.; Mukai, T.; Saji, H. A high-affinity fluorescent Zn2+ sensor improved by the suppression of pyridine-pyridone tautomerism and its application in living cells. Sens. Act. B Chem. 2015, 213, 45–52. [Google Scholar] [CrossRef]
  18. 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]
  19. Andina, D.; Leroux, J.C.; Luciani, P. Ratiometric fluorescent probes for the detection of reactive oxygen species. Chem. Eur. J. 2017, 23, 13549–13573. [Google Scholar] [CrossRef]
  20. Li, X.; Baryshnikov, G.; Deng, C.; Bao, X.; Wu, B.; Zhou, Y.; Ågren, H.; Zhu, L. A three-dimensional ratiometric sensing strategy on unimolecular fluorescence-thermally activated delayed fluorescence dual emission. Nat. Commun. 2019, 10, 731. [Google Scholar] [CrossRef]
  21. Golovine, K.; Uzzo, R.G.; Makhov, P.; Crispen, P.L.; Kunkle, D.; Kolenko, V.M. Depletion of intracellular zinc increases expression of tumorigenic cytokines VEGF, IL-6 and IL-8 in prostate cancer cells via NF-κB dependent pathway. Prostate 2008, 68, 1443–1449. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Zalewski, P.D.; Forbes, I.J.; Seamark, R.F.; Borlinghaus, R.; Betts, W.H.; Lincoln, S.F.; Ward, A.D. Flux of intracellular labile zinc during apoptosis (gene-directed cell death) revealed by a specific chemical probe. Chem. Biol. 1994, 1, 53–61. [Google Scholar] [CrossRef]
  23. Hanaoka, K.; Kikuchi, K.; Kojima, H.; Urano, Y.; Nagano, T. Selective detection of zinc ions with novel luminescent lanthanide probes. Angew. Chem. Int. Ed. 2003, 42, 2996–2999. [Google Scholar] [CrossRef] [PubMed]
  24. Benesi, H.A.; Hildebrand, J.H. A spectrophotometric investigation of the interaction of iodine with aromatic hydrocarbons. J. Am. Chem. Soc. 1949, 71, 2703–2707. [Google Scholar] [CrossRef]
  25. Kumawat, L.M.; Asif, M.; Gupta, V.K. Dual ion selective fluorescence sensor with potential applications in sample monitoring and membrane sensing. Sens. Act. B Chem. 2017, 241, 1090–1098. [Google Scholar] [CrossRef] [Green Version]
Scheme 1. Preparation of 6-amino-2,2′-bipyridine-based compounds 36 (rBpyZ).
Scheme 1. Preparation of 6-amino-2,2′-bipyridine-based compounds 36 (rBpyZ).
Molecules 27 01287 sch001
Figure 1. UV/Vis absorption spectra of 10 µM of (a) 3, (b) 4, (c) 5, and (d) 6 (rBpyZ), in HEPES buffer (100 mM, 50% EtOH, pH = 7.4), The arrows indicate the direction in which the spectra shift when in the presence of Zn2+ as its perchlorate salt (0–1000 µM).
Figure 1. UV/Vis absorption spectra of 10 µM of (a) 3, (b) 4, (c) 5, and (d) 6 (rBpyZ), in HEPES buffer (100 mM, 50% EtOH, pH = 7.4), The arrows indicate the direction in which the spectra shift when in the presence of Zn2+ as its perchlorate salt (0–1000 µM).
Molecules 27 01287 g001
Figure 2. Fluorescence spectra of (a) 3 (1 μM, Exmax = 359 nm); (b) 4 (1 μM, Exmax = 360 nm); (c) 5 (1 μM, Exmax = 359 nm); (d) 6 (rBpyZ) (1 μM, Exmax = 326 nm) upon the addition of Zn2+ as the perchlorate salt (0–100 μM) in HEPES buffer (100 mM, 50% EtOH, pH = 7.4). Exmax = excitation wavelength.
Figure 2. Fluorescence spectra of (a) 3 (1 μM, Exmax = 359 nm); (b) 4 (1 μM, Exmax = 360 nm); (c) 5 (1 μM, Exmax = 359 nm); (d) 6 (rBpyZ) (1 μM, Exmax = 326 nm) upon the addition of Zn2+ as the perchlorate salt (0–100 μM) in HEPES buffer (100 mM, 50% EtOH, pH = 7.4). Exmax = excitation wavelength.
Molecules 27 01287 g002
Figure 3. Normalized fluorescence intensity of 6 (rBpyZ) at 438 nm (▲) and 465 nm (●).
Figure 3. Normalized fluorescence intensity of 6 (rBpyZ) at 438 nm (▲) and 465 nm (●).
Molecules 27 01287 g003
Figure 4. 1H-NMR spectra of (a) 6 (rBpyZ), (b) 6 (rBpyZ) + 0.5 equiv. Zn2+, and (c) 6 (rBpyZ) + 1.0 equiv. Zn2+ in DMSO-d6.
Figure 4. 1H-NMR spectra of (a) 6 (rBpyZ), (b) 6 (rBpyZ) + 0.5 equiv. Zn2+, and (c) 6 (rBpyZ) + 1.0 equiv. Zn2+ in DMSO-d6.
Molecules 27 01287 g004
Figure 5. Fluorescence spectra of 6 (rBpyZ) (1 µM) in HEPES buffer (100 mM, 50% EtOH, pH = 7.4, excitation wavelength: 326 nm) upon the addition of different metal cations (100 µM).
Figure 5. Fluorescence spectra of 6 (rBpyZ) (1 µM) in HEPES buffer (100 mM, 50% EtOH, pH = 7.4, excitation wavelength: 326 nm) upon the addition of different metal cations (100 µM).
Molecules 27 01287 g005
Figure 6. Competitive binding experiments of 6 (rBpyZ), in which competing metal ions (10−3 M) and Zn2+ (10−3 M) were added to 6 (rBpyZ) (10−6 M) in HEPES buffer (100 mM, 50% EtOH, pH = 7.4).
Figure 6. Competitive binding experiments of 6 (rBpyZ), in which competing metal ions (10−3 M) and Zn2+ (10−3 M) were added to 6 (rBpyZ) (10−6 M) in HEPES buffer (100 mM, 50% EtOH, pH = 7.4).
Molecules 27 01287 g006
Figure 7. Effects of pH on the fluorescence properties of 6 (rBpyZ) in the absence (□) and presence (■) of Zn2+.
Figure 7. Effects of pH on the fluorescence properties of 6 (rBpyZ) in the absence (□) and presence (■) of Zn2+.
Molecules 27 01287 g007
Figure 8. Fluorescence images of exogenously added Zn2+ in A549 cells: (a) fluorescence image of cells with 6 (rBpyZ) (100 μM); (b) fluorescence image of cells incubated with 6 (rBpyZ) (100 μM) in the presence of Zn2+/pyrithione (zinc ionophore; 100 μM); (c) fluorescence image of cells in (b) after treatment with TPEN (100 μM). Scale bars = 300 μm.
Figure 8. Fluorescence images of exogenously added Zn2+ in A549 cells: (a) fluorescence image of cells with 6 (rBpyZ) (100 μM); (b) fluorescence image of cells incubated with 6 (rBpyZ) (100 μM) in the presence of Zn2+/pyrithione (zinc ionophore; 100 μM); (c) fluorescence image of cells in (b) after treatment with TPEN (100 μM). Scale bars = 300 μm.
Molecules 27 01287 g008
Figure 9. Fluorescence images of endogenous Zn2+ in A549 cells: (a) bright-field image of cells with 6 (rBpyZ) (100 μM); (b) fluorescence image of cells with 6 (rBpyZ) (100 μM); (c) bright field image incubated with 6 (rBpyZ) (100 μM) after incubation with H2O2 (100 μM) for 24 h; (d) fluorescence image incubated with 6 (rBpyZ) (100 μM) after incubation with H2O2 (100 μM) for 24 h. Scale bars = 300 μm.
Figure 9. Fluorescence images of endogenous Zn2+ in A549 cells: (a) bright-field image of cells with 6 (rBpyZ) (100 μM); (b) fluorescence image of cells with 6 (rBpyZ) (100 μM); (c) bright field image incubated with 6 (rBpyZ) (100 μM) after incubation with H2O2 (100 μM) for 24 h; (d) fluorescence image incubated with 6 (rBpyZ) (100 μM) after incubation with H2O2 (100 μM) for 24 h. Scale bars = 300 μm.
Molecules 27 01287 g009
Table 1. Fluorescence data for compounds 36 (rBpyZ).
Table 1. Fluorescence data for compounds 36 (rBpyZ).
Compoundsλ maxlog єEx maxEm maxSS aΦ b Kd (nM)
(nm)(nm)(nm)(nm)
33524.153594661070.07 (−Zn2+)2.2
0.37 (+Zn2+)
43503.983604671070.05 (−Zn2+)1.2
0.35 (+Zn2+)
53503.983594651060.05 (−Zn2+)3
0.52 (+Zn2+)
6 (rBpyZ)3194.19326437 (−Zn2+)111 (−Zn2+)0.45 (−Zn2+)0.77
465 (+Zn2+)139 (+Zn2+)0.63 (+Zn2+)
a Stokes shift. b The fluorescence quantum yield values were measured with respect to quinine sulfated solution as standard.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hagimori, M.; Hara, F.; Mizuyama, N.; Fujino, T.; Saji, H.; Mukai, T. High-Affinity Ratiometric Fluorescence Probe Based on 6-Amino-2,2′-Bipyridine Scaffold for Endogenous Zn2+ and Its Application to Living Cells. Molecules 2022, 27, 1287. https://doi.org/10.3390/molecules27041287

AMA Style

Hagimori M, Hara F, Mizuyama N, Fujino T, Saji H, Mukai T. High-Affinity Ratiometric Fluorescence Probe Based on 6-Amino-2,2′-Bipyridine Scaffold for Endogenous Zn2+ and Its Application to Living Cells. Molecules. 2022; 27(4):1287. https://doi.org/10.3390/molecules27041287

Chicago/Turabian Style

Hagimori, Masayori, Fumiko Hara, Naoko Mizuyama, Takeshi Fujino, Hideo Saji, and Takahiro Mukai. 2022. "High-Affinity Ratiometric Fluorescence Probe Based on 6-Amino-2,2′-Bipyridine Scaffold for Endogenous Zn2+ and Its Application to Living Cells" Molecules 27, no. 4: 1287. https://doi.org/10.3390/molecules27041287

Article Metrics

Back to TopTop