Next Article in Journal
Development and Field Validation of Low-Cost Metal Oxide Nanosensors for Tropospheric Ozone Monitoring in Rural Areas
Next Article in Special Issue
Optimizing the Heavy Metal Ion Sensing Properties of Functionalized Silver Nanoparticles: The Role of Surface Coating Density
Previous Article in Journal
Machine Learning-Based Multi-Level Fusion Framework for a Hybrid Voltammetric and Impedimetric Metal Ions Electronic Tongue
Previous Article in Special Issue
Deferoxamine-Based Materials and Sensors for Fe(III) Detection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Rational Design of Ratiometric Fluorescent Probe for Zn2+ Imaging under Oxidative Stress in Cells

1
State Key Laboratory of Coordination Chemistry Coordination Chemistry, School of Chemistry and Chemical Engineering, Nanjing University, Nanjing 210023, China
2
Chemistry and Biomedicine Innovation Center (ChemBIC), Nanjing University, Nanjing 210023, China
3
Nanchuang (Jiangsu) Institute of Chemistry and Health, Nanjing 210000, China
*
Authors to whom correspondence should be addressed.
Chemosensors 2022, 10(11), 477; https://doi.org/10.3390/chemosensors10110477
Submission received: 20 October 2022 / Revised: 7 November 2022 / Accepted: 8 November 2022 / Published: 13 November 2022
(This article belongs to the Special Issue Chemosensors for Ion Detection)

Abstract

:
Zn2+ is a vital ion for most of the physiological processes in the human body, and it usually has a mutual effect with oxidative stress that often occurs in pathological tissues. Detecting fluctuation of Zn2+ level in cells undergoing oxidative stress could be beneficial to understanding the relationship between them. Herein, a ratiometric fluorescent Zn2+ probe was rationally designed. The wavelength corresponding to the maximum fluorescence intensity bathometrically shifted from 620 nm to 650 nm after coordinating with Zn2+. The intensity ratio of two fluorescence channels changed significantly in cells treated by oxidative stress inducers. It was shown from the results that the labile zinc level was generally elevated under oxidative stress stimulated by various inducers.

1. Introduction

The divalent zinc (Zn2+) is crucial for all creatures. In the cellular level, the spatio-temporal distribution of labile zinc has an effect on the process of signal transduction, gene expression, and cell death by adjusting the extent of zinc–protein and/or zinc–nucleus acid combination [1]. As for human beings, Zn2+ features irreplaceably in numerous biochemical processes, including combining with insulin to guarantee the normal function of pancreas islet [2], inhibiting m-aconitase activity to maintain a high level of citrate in prostate epithelial cells [3], and fluxing during mammalian sperm capacitation to contribute to fertilization [4]. Among them, oxidative stress could be one of the major ways for Zn2+ to affect cell metabolism [5].
Zn2+ acts as an anti-oxidant indirectly at a certain concentration range [6], and it also has pro-oxidant effects at both excessively high and low concentration levels [7]. In turn, however, Zn2+ is released from macromolecules (e.g., metallothionein) during oxidative stress so that the concentration of its labile state elevates [8]. More specifically, if some macromolecules containing Zn2+ are destroyed to lose their chelating ability by some stimulants, Zn2+ will be released, which is common in pathological tissues [9,10,11]. Detecting the Zn2+ released during this process could be helpful for studying the mutual affection between Zn2+ and oxidative stress. Furthermore, the extent of impact of Zn2+-containing macromolecules induced by different chemicals can be learned from Zn2+ imaging.
Fluorescence imaging has the merit of good selectivity, sensitivity, and convenience to present the accurate spatio-temporal distribution of analytes, so that it is frequently used to image the metal ion in cells [12]. While there are lots of brilliant materials that can be used as fluorescent sensors to detect metal ions [13,14,15,16], small molecule probes stand out among the genetically coded probes and nanoparticle probes by its properties of low toxicity, ease of preparation, and being relatively stable [17]. Currently, satisfactory small molecule fluorescent probes for detecting Zn2+ in oxidative context are still rare [18,19,20,21], mainly because it is challenging to simultaneously consider all of the essential factors such as selectivity, stability, and prolonging the wavelength of the spectrum. The laser used for excitation should try to be in a visible or near infrared region to avoid extra damage to cells and enhance the signal-to-background ratio during imaging. Moreover, ratiometric probes are preferred in ion detecting occasions since it has a self-calibration effect to prevent many kinds of interference [22].
Based on the fact described above, we synthesized and screened a fluorescent probe to ratiometrically detect Zn2+ in stressed cells. A BODIPY fluorophore was used since it is known to be much more stable and bright in common cell environments [23]. A mesityl group and two methyl groups were introduced to the meso- and β-position separately to prevent rotating induced relaxation, and a pyrrole ring was directly bound to the α-position of the BODIPY core, so that the wavelengths of both the excitation and emission spectra were prolonged significantly. Among the BODIPY based ratiometric probes, the intramolecular charge transfer (ICT) mechanism was mostly adopted in the designation by locating the recognition group at the 3, 5-position [24], or 8-position [25]. However, this strategy did not always work in the ion-detecting probes because of either the photo-induced electron transfer (PeT) quenching effect or the complexity of the chelating group [26,27]. Therefore, apart from linking a chelating group to the fluorophore, an ester group was added to the ortho-position of the chelating group, which may cause a red shift rather than a blue shift while reacting with a zinc ion [28]. The quantity and position of the ligands were screened to obtain the optimal probe BDP-p-BPEA, while another two probes BDP-p-TMPA and BDP-2BPEA were listed for comparison (Scheme 1).

2. Materials and Methods

2.1. Materials and Instruments

All reagents for probe preparation were of analytic grade and purchased from Energy Chemistry (Shanghai, China) and J&K Chemical (Beijing, China). The 1H NMR and 13C NMR spectra were recorded at 25 °C with a Bruker Avance DRX-400 (German) with TMS as the internal reference. High resolution mass spectrometric data were determined using an Agilent 6540Q-TOF HPLC-MS spectrometer (America). The ultrapure water for spectroscopic and cell culture was obtained from a Millipore system (>18.2 MΩ). Confocal imaging was performed with a Zeiss LSM710 microscope (German).

2.2. Synthesis of Intermediates 14 and Probes BDP-2BPEA, BDP-p-BPEA, and BDP-p-TMPA

2.2.1. Synthesis of 1

To a dichloromethane (16 mL) solution of ethyl 4-methyl-1H-pyrrole-3-carboxylate (3.15 g, 20.5 mmol) were added 33 mL of acetic acid and 0.75 mL of concentrated sulfuric acid under nitrogen atmosphere. After stirring at room temperature for 10 min, 2,4,6-trimethylbenzaldehyde (1.50 g, 10.1 mmol) was added dropwise. The mixture was left to react for 7 h. Upon completion, the solvent was removed by rotary evaporation. Dichloromethane was used to dissolve the residue and the solution was neutralized by a saturated NaHCO3 aqueous solution. The aqueous phase was extracted by the same volume of dichloromethane twice. The combined organic phase was dried with Na2SO4. The solution of crude product was concentrated and purified by silica gel column chromatography using dichloromethane to dichloromethane/methanol = 4/1 as the eluent to give compound 1 as a white solid (3.31 g, 75.1% yield). MS (ESI+) m/z: calculated 436.23621; found 437.24182 [M + H]+, 459.22357 [M + Na]+. 1H NMR (400 MHz, CDCl3) δ 7.83 (s, 2H), 7.23 (d, J = 3.1 Hz, 2H), 6.86 (s, 2H), 5.80 (s, 1H), 4.26 (qd, J = 7.1, 1.4 Hz, 4H), 2.27 (s, 3H), 1.96 (s, 6H), 1.33 (t, J = 7.1 Hz, 6H). 13C NMR (101 MHz, CDCl3) δ 165.64, 136.94, 136.80, 132.34, 130.91, 127.22, 122.06, 117.15, 116.21, 59.52, 36.07, 20.88, 20.76, 14.63, 10.24.

2.2.2. Synthesis of 2

The synthesis of 2 referred to previously reported literature [29]. Compound 1 (2.00 g, 4.58 mmol) was dissolved into dry tetrahydrofuran (THF, 45 mL) under nitrogen atmosphere, and cooled to –78 °C for more than 10 min. Then, N-bromosuccinimide (1.63 g, 9.16 mmol) was added in three portions every 10 min. After stirring at –78 °C for one hour, 2,3-dichloro-5,6-dicyano-p-benzoquinone (1.04 g, 4.58 mmol) was added to the mixture gradually. The mixture reacted for 10 min at −78 °C and was transferred to room temperature and stirred for another 1.5 h. Upon completion, the solvent was evaporated. Forty milliliters of dichloromethane were added, and less soluble components were removed by filtration. The filtrate was concentrated and purified by silica gel column chromatography using petroleum ether/dichloromethane = 2/1 to 0/1 as the eluent to give compound 2 as an orange solid (2.53 g, 93.4% yield). MS (ESI+) m/z: calculated 592.03954; found 593.04224 [M + H]+, 615.02521 [M + Na]+. 1H NMR (400 MHz, CDCl3) δ 14.13 (s, 1H), 6.96 (s, 2H), 4.28 (q, J = 7.1 Hz, 4H), 2.34 (s, 3H), 2.05 (s, 6H), 1.60 (s, 6H), 1.34 (t, J = 7.1 Hz, 6H). 13C NMR (101 MHz, CDCl3) δ 163.47, 146.10, 143.79, 139.47, 137.37, 135.10, 131.79, 130.56, 129.55, 121.95, 60.58, 21.36, 19.65, 14.33, 12.43.

2.2.3. Synthesis of 3

To a solution of compound 2 (2.75 g, 4.64 mmol) in 50 mL of dichloromethane was added diisopropylethylamine (4.9 mL, 27.8 mmol). The mixture was stirred for 10 min, and BF3·OEt2 (4.8 mL, 37.1 mmol) was dripped into it. The mixture was left to react at room temperature for two hours. Fifteen milliliters of water were added to quench the reaction, and was then neutralized by 1 M NaOH. The aqueous phase was extracted by the same volume of dichloromethane twice. The combined organic phase was dried with Na2SO4. The solution of crude product was concentrated and purified by silica gel column chromatography using petroleum ether/dichloromethane = 2/1 to 0/1 as the eluent to give compound 3 as a red solid (1.96 g, 66.0% yield). MS (ESI+) m/z: calculated 640.03782; found 641.04327 [M + H]+, 663.02448 [M + Na]+. 1H NMR (400 MHz, CDCl3) δ 7.02 (s, 2H), 4.32 (q, J = 7.1 Hz, 4H), 2.37 (s, 3H), 2.08 (s, 6H), 1.68 (s, 6H), 1.35 (t, J = 7.1 Hz, 6H). 13C NMR (101 MHz, CDCl3) δ 162.63, 147.60, 147.01, 140.45, 134.44, 134.41, 132.50, 130.02, 129.70, 124.68, 61.15, 21.41, 19.68, 14.32, 12.58.

2.2.4. Synthesis of 4

The synthesis of 4 referred to previously reported literature [30]. Briefly, the mixture of compound 3 (200 mg, 0.31 mmol) and 4 mL of pyrrole was refluxed for 10 h under nitrogen atmosphere. Then, the solvent was removed by rotary evaporation. The crude product was purified by silica gel column chromatography using petroleum ether/dichloromethane = 1/1 to 0/1 as the eluent to give compound 4 as a dark green solid (0.16 g, 82.4% yield). MS (ESI+) m/z: calculated 625.15591; found 626.16199 [M + H]+, 648.14520 [M + Na]+. 1H NMR (400 MHz, CDCl3) δ 10.59 (br, 1H), 7.20 (m, 1H), 6.99 (s, 2H), 6.86 (m, 1H), 6.36 (dt, J = 4.2, 2.3 Hz, 1H), 4.32 (p, J = 7.1 Hz, 3H), 2.36 (s, 3H), 2.11 (s, 6H), 1.66 (s, 3H), 1.47 (s, 3H), 1.35 (t, J = 7.1 Hz, 3H), 1.30 (t, J = 7.1 Hz, 3H). 13C NMR (101 MHz, CDCl3) δ 166.02, 163.47, 148.29, 145.48, 141.42, 140.70, 139.82, 135.29, 133.55, 130.58, 130.41, 129.61, 128.13, 126.41, 126.33, 122.35, 121.40, 119.68, 111.64, 61.88, 60.70, 21.39, 19.79, 14.38, 14.15, 12.54, 11.86.

2.2.5. Synthesis of BDP-2BPEA

A mixture of compound 3 (100 mg, 0.16 mmol), N,N-bis(pyridin-2-ylmethyl)ethane-1,2-diamine (151 mg, 0.62 mmol) and triethylamine (86 μL, 0.62 mmol) in 8 mL of acetonitrile was stirred at 60 °C overnight under nitrogen atmosphere. Then, the solvent was removed by rotary evaporation. The crude product was purified by neutral aluminium oxide column chromatography using dichloromethane/ethanol = 100/1 to 20/1 as the eluent to give BDP-2BPEA as a blue foam (44 mg, 28.6% yield). MS (ESI+) m/z: calculated 962.49384; found 963.50043 [M + H]+, 985.48047 [M + Na]+. 1H NMR (400 MHz, CDCl3) δ 8.53 (br, 2H), 8.48 (ddd, J = 5.0, 1.8, 0.9 Hz, 4H), 7.83 (dt, J = 7.8, 1.1 Hz, 4H), 7.60 (td, J = 7.6, 1.8 Hz, 4H), 7.11 (ddd, J = 7.5, 4.9, 1.2 Hz, 4H), 6.91 (s, 2H), 4.29 (q, J = 7.1 Hz, 4H), 3.96 (t, J = 5.9 Hz, 4H), 3.89 (s, 8H), 2.85 (t, J = 5.8 Hz, 4H), 2.33 (s, 3H), 2.06 (s, 6H), 1.59 (s, 6H), 1.32 (t, J = 7.1 Hz, 6H). 13C NMR (101 MHz, CDCl3) δ 166.54, 159.37, 156.13, 148.87, 142.43, 138.57, 136.67, 136.63, 133.38, 132.67, 128.96, 125.84, 123.52, 122.14, 107.57, 60.18, 60.10, 53.41, 41.62, 21.36, 19.68, 14.53, 12.82.

2.2.6. Synthesis of BDP-p-BPEA

A mixture of compound 4 (160 mg, 0.26 mmol), N,N-bis(pyridin-2-ylmethyl)ethane-1,2-diamine (93 mg, 0.38 mmol) and triethylamine (53 μL, 0.38 mmol) in 10 mL of acetonitrile was stirred at 60 °C for 2 h under nitrogen atmosphere. Then, the solvent was removed by rotary evaporation. The crude product was purified by silica gel column chromatography using dichloromethane/methanol = 1/0 to 10/1 as the eluent to give BDP-p-BPEA as a brown solid (130 mg, 63.4% yield). MS (ESI+) m/z: calculated 787.38289; found 788.38849 [M + H]+, 810.36523 [M + Na]+. 1H NMR (400 MHz, CDCl3) δ 9.98 (br, 1H), 9.06 (br, 1H), 8.52 (d, J = 4.9 Hz, 2H), 7.77 (d, J = 7.9 Hz, 2H), 7.63 (td, J = 7.7, 1.8 Hz, 2H), 7.16 (ddd, J = 7.5, 4.8, 1.2 Hz, 2H), 6.95 (s, 3H), 6.62–6.55 (m, 1H), 6.26 (dt, J = 3.7, 2.6 Hz, 1H), 4.35 (q, J = 7.1 Hz, 2H), 4.20 (q, J = 7.1 Hz, 2H), 4.01 (q, J = 5.6 Hz, 2H), 3.94 (s, 4H), 2.93 (t, J = 5.9 Hz, 2H), 2.34 (s, 3H), 2.09 (s, 6H), 1.71 (s, 3H), 1.44 (s, 3H), 1.35 (t, J = 7.1 Hz, 3H), 1.20 (t, J = 7.1 Hz, 3H). 13C NMR (101 MHz, CDCl3) δ 167.09, 165.87, 158.70, 158.25, 150.21, 148.76, 138.97, 138.75, 136.94, 136.22, 134.83, 133.92, 131.79, 129.75, 129.17, 128.44, 123.72, 122.65, 122.42, 122.37, 119.92, 113.36, 112.53, 109.23, 60.99, 60.77, 59.90, 52.83, 41.69, 21.36, 19.80, 14.40, 14.16, 13.52, 11.26.

2.2.7. Synthesis of BDP-p-TMPA

A mixture of compound 4 (140 mg, 0.22 mmol) and 1-(6-(aminomethyl)pyridin-2-yl)-N,N-bis(pyridin-2-ylmethyl)methanamine (103 mg, 0.32 mmol) in 10 mL of acetonitrile was stirred at 60 °C for 1 h under nitrogen atmosphere. Then, the solvent was removed by rotary evaporation. The crude product was purified by silica gel column chromatography using dichloromethane as the eluent to give BDP-p-TMPA as a purple foam (164 mg, 86.3% yield). MS (ESI+) m/z: calculated 864.40944; found 865.41260 [M + H]+, 887.39386 [M + Na]+. 1H NMR (400 MHz, CDCl3) δ 10.05 (br, 1H), 10.02 (br, 1H), 8.51 (dt, J = 4.9, 1.4 Hz, 2H), 7.67 (t, J = 7.6 Hz, 1H), 7.65–7.61 (m, 2H), 7.63–7.56 (m, 2H), 7.50 (d, J = 7.6 Hz, 1H), 7.20 (d, J = 7.6 Hz, 1H), 7.13 (ddd, J = 6.8, 4.9, 1.6 Hz, 2H), 6.98–6.93 (m, 3H), 6.65–6.58 (m, 1H), 6.28 (dt, J = 3.7, 2.5 Hz, 1H), 5.20 (d, J = 4.5 Hz, 2H), 4.24 (q, J = 7.1 Hz, 2H), 4.19 (q, J = 7.1 Hz, 2H), 3.99 (s, 2H), 3.94 (s, 4H), 2.35 (s, 3H), 2.11 (s, 6H), 1.73 (s, 3H), 1.46 (s, 3H), 1.25 (t, J = 7.3 Hz, 3H), 1.20 (t, J = 7.1 Hz, 3H). 13C NMR (101 MHz, CDCl3) δ 167.11, 165.41, 159.56, 158.65, 158.16, 154.68, 150.67, 149.07, 138.92, 138.45, 137.32, 136.60, 136.26, 134.62, 133.48, 131.83, 130.11, 129.14, 128.36, 123.16, 122.77, 122.27, 122.13, 121.84, 120.36, 119.71, 113.17, 113.09, 109.18, 60.93, 60.70, 60.35, 60.12, 49.03, 42.08, 27.12, 25.10, 21.34, 19.82, 14.32, 14.15, 13.53, 11.22.

2.3. Optical Experimental Method

Stock solutions of metal ions and probes were prepared by dissolving an appropriate amount in ultrapure water and DMSO, respectively. Solvents for spectroscopic study were of spectral grade from Tedia Company Inc. Absorption spectra were recorded using a PerkinElmer E35 spectrophotometer. Fluorescence spectra were determined using a Horiba FluoroMax-4 spectrofluorometer with the 2 nm and 1 nm slit for excitation and emission, respectively. Excitation wavelength was 550 nm.
The detection limit of three probes were determined according to previous literature [31]. Taking BDP-p-BPEA as an example, the fluorescence intensity at 670 nm and 595 nm of the probe was determined 20 times in the identical condition with what was used in the titration experiment. The standard deviation of the twenty F670/F595 value was calculated to be the background noise σ. Then, the limit of detection (LOD) was determined using the following equation:
LOD = 3 σ k ,   k   was   the   slope   of   the   linear   fitting   curve .

2.4. Cell Culture

For oxidative stress inducing experiment, Hela cells were washed by phosphate-buffered saline (PBS) and incubated with 5 µM of probe (in Dulbecco’s modified eagle medium, DMEM) for 4 h at 37 °C, followed by washing another 3 times with PBS. Then, 100 µM of 4,4′-dithiodipyridine (DTDP), 50 µM of N-ethylmaleimide (NEM), 50 µM of TPEN, 200 µM of N’,N’-diethyl-N-hydroxynitrous hydrazide diethyl ammonium salt (DEA-NONOate) and 20 µM of cisplatin (in DMEM) were added separately to different dishes of cells. Then, the culture dish was put back to incubate for 30 min (for DTDP, NEM and TPEN) or 2 h (for DEA-NONOate and cisplatin), and the cells were washed 3 times with PBS before acquiring the fluorescence image.
The cytotoxicity experiment of BDP-p-BPEA in Hela cells was measured using an MTT assay. Cells were seeded into a 96-well cell culture plate at a rate of 4300/well in 5% CO2 at 37 °C for 24 h, followed by adding probe BDP-p-BPEA (0–8 µM), and then allowed to incubate for 24 h, respectively. Subsequently, each well was injected with 40 µL MTT (2.5 mg/mL) and cultured for another 4 h under the same conditions. Then, the culture medium was removed, and 150 µL of DMSO were added to each well to dissolve the formazan. Absorption of 490 nm of each well was obtained on a microplate reader to indirectly calculate the amount of live cells.

3. Results

3.1. Design and Synthesis of the Three Probes

The BODIPY core was constructed in a conventional method, that is, aldehyde and two equivalents of pyrrole were condensed and then functionalized by halogenation and nucleophilic substitution. In this way, we successfully obtained three fluorescent probes that were shown in Scheme 1. All of the newly synthesized compounds were characterized by 1H NMR, 13C NMR, and ESI-MS (Figures S1–S21).

3.2. Spectral Response Towards Zn2+ and Screening of the Probes

With three probes in hand, we obtained their absorption spectra in the range of 400–750 nm in acetonitrile/aqueous (v/v = 1:1) solution (Figure 1a–c). It can be seen that the absorption bands of BDP-p-BPEA and BDP-p-TMPA were broad (full width at half maximum were about 100 nm), and the wavelength of maximum absorption changed for 5–10 nm after chelating with Zn2+. Thus, their ratiometric absorption turned out to be negligible. As for fluorescence spectra, all of them exhibited ratiometric response to Zn2+. BDP-2BPEA could chelate with two equivalents of Zn2+, and the inflection point at the one equivalent position in the fluorescence titration graph inferred the different electron density distribution on the fluorophore between reacting with one and two equivalents of Zn2+ (Figure 1d). The fluorescence spectrum of this compound hypochromatically shifted after chelating, regardless of how much Zn2+ was added. It was supposed that this probe adopted a different binding mode from that of the other two counterparts. In addition, the fluorescence peak of BDP-p-BPEA presented the bathochromic shift from 620 nm to 650 nm when reacting with a sufficient amount of Zn2+ (Figure 1e). When the chelator of the probe was changed to a hexadentate, the bathochromic shift slightly shortened as the electron density change of the ligands that conjugated to fluorophore in average attenuated (Figure 1f). Thus, BDP-p-BPEA had a lower detection limit (29 nM) than BDP-p-TMPA (41 nM) and BDP-2BPEA (50 nM).
Based on the screening results displayed above, the probe BDP-p-BPEA was selected to study the optical properties in detail. The probe reached coordination equilibrium with Zn2+ within 30 s (Figure 2a), and the reversibility, crucial to displaying the dynamic change of Zn2+ level, was also checked by adding ZnCl2 and Zn2+ scavenger N,N,N’,N’-tetrakis(2-pyridinylmethyl)-1,2-ethanediamine (TPEN) solution alternatively to the solution of BDP-p-BPEA for five cycles (Figure 2b). The dissociation constant for BDP-p-BPEA + Zn2+ was calculated to be 3.2 nM (Figure S22), providing sufficient affinity for BDP-p-BPEA to capture the labile Zn2+. Then, the selectivity was tested with common metal ions in the environment (Figure 2c). H2O2 and its mixture with Fe2+ were used to mimic oxidative environment in cells. Except for Cu2+, all of the tested species negligibly affected the fluorescence response of the probe to Zn2+. Fortunately, the Cu2+ level in cells was too low to compete with the coordination site of the probe with Zn2+ [32].

3.3. Exploring the Sensing Mechanism for BDP-p-BPEA

The Zn2+ binding stoichiometry of BDP-p-BPEA was studied by Job’s plot and ESI-MS. The fluorescence ratio reached maximum when c(BDP-p-BPEA)/([Zn2+] + c(BDP-p-BPEA)) equaled to 0.5 (Figure 2d); meanwhile, a [M − H + Zn]+ peak (m/z = 850.30214) and a [M + Cl + Zn]+ peak (m/z = 886.27891) were presented in the mass spectrum of BDP-p-BPEA-Zn2+ complex (Figure S23), proving that the probe chelated with Zn2+ in the manner of 1:1. After that, the 1H NMR spectrum of BDP-p-BPEA in DMSO-d6 before and after being added with ZnCl2 were compared (Scheme 2a). We found that three of the H signals of the pyridine rings moved to the lower field, while the H signal of the 3-position of pyridine upshifted. All of the H on the pyrrole ring downshifted because the chelation lowered the electron density of the whole conjugation system. Chemical shift of methylene hydrogen in the chelating group and ester group also changed in different degrees, indicating that the ester group involved in chelating without hydrolysis. In a word, BDP-p-BPEA chelated with Zn2+ in the way showed in Scheme 2b.

3.4. Imaging of Zn2+ Levels in Cellular Oxidative Stress

As BDP-p-BPEA showed great fluorescent response to Zn2+ in vitro, we next explored its potential to imaging Zn2+ in cells. The intake of the probe was firstly investigated. The fluorescence intensity in cells gradually enhanced and stopped at 4 h (Figure S24), that is, it took about 4 h for BDP-p-BPEA to reach the balance of intake and outlet. The MTT assay was carried out for three times to make sure the cell viability would not obviously decline while imaging. As a result, the survival rate of Hela cells kept more than 80% after being treated by 0–8 μM of BDP-p-BPEA for 24 h (Figure S25), so the cytotoxicity was evaluated to be low enough to apply the probe to cell imaging. Interested about the cell distribution of the probe, we co-stained BDP-p-BPEA with mitochondrial, lysosome, endoplasmic reticulum (ER), Golgi apparatus, and lipid droplet targeting commercial dyes separately. As shown in Figure 3, BDP-p-BPEA mainly distributed in lipid droplet and ER, and other organelles were also stained in varied degrees. This property enabled us to observe the overall fluctuation of Zn2+ level in cells.
Cells fully stained with BDP-p-BPEA exhibited a bright fluorescence signal, so we tried to determine the variation of Zn2+ levels in stress induced by different mechanisms using BDP-p-BPEA. Cisplatin and scavenger of biothiols—DTDP and NEM were used to treat Hela cells. All the reagents listed above were reported to cause oxidative stress at certain concentrations [33,34,35]. Nitric oxide, provided by DEA-NONOate, reacted with metalloproteins, which not only released Zn2+ from storage, but also had a close relationship with the formation of oxidative stress [36,37]. Therefore, it was also adopted to treat Hela cells. We used a ratio graph to show the relative intensity change of the two channels more clearly. The results in Figure 4 showed that NO, DTDP and cisplatin significantly raised ratio value, while NEM induced a relatively low increase. The results indicated that oxidative stress easily converted Zn2+ from a tightly-binding state to labile state. TPEN made the ratio value slightly lower than that of the control group, and it means the original endogenous labile zinc can be detected by BDP-p-BPEA. As a result, it was disclosed with BDP-p-BPEA that the labile Zn2+ level was elevated during oxidative stress.

3.5. Comparison with the Previous Methods

Many fluorescent Zn2+ probes have been designed up to now, and some representative ones were listed below (Table 1). Though cyanine and hemicyanine based probes had excellent hydrophilicity and near infrared absorption, they seldom performed ratiometric responses to Zn2+ [21,38]. Meanwhile, the cyanine structure was fragile to destruction from reactive oxygen species. The combination of coumarin and other fluorophores could be made as a ratiometric probe based on a fluorescence resonance energy transfer (FRET) mechanism [39,40], but they had higher LOD value and had to be excited by a near ultraviolet laser, which may cause extra damage to cells while imaging. Currently, most of the BODIPY based Zn2+ probes blueshifted after their coordination with Zn2+ [41], while BDP-p-BPEA presented a redshift response. It was a pity that a large part of the probes could not simultaneously solve the problems of sensitivity [42], solubility in water [43], and short wavelength [44]. The probe proposed in this work showed great sensitivity, selectivity, and ratiometric response to Zn2+, so it has better imaging performance than other Zn2+ probes to some extent.

4. Conclusions

To sum up, we synthesized and screened a ratiometric fluorescent Zn2+ probe with a stable structure, and it was capable of imaging in the oxidative environment. Its maximum fluorescence signal moved from 620 nm to 650 nm after chelating. Brilliant optical properties were obtained to image the fluctuation of labile zinc in Hela cells undergoing oxidative stress. In addition, cisplatin induced elevation of Zn2+ level in cells was also observed. The probe showed its potential to be used in super-resolution imaging in the future by virtue of its spectra of relatively long wavelength and stability, we will consider modifying the structure of BDP-p-BPEA to further prolong its wavelength so that it can be applied to in vivo Zn2+ imaging of inflammatory tissues, where oxidative stress frequently occurs.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/chemosensors10110477/s1, Figure S1: ESI-MS spectra of 1; Figure S2: 1H NMR spectra of 1; Figure S3: 13C NMR spectra of 1; Figure S4: ESI-MS spectra of 2; Figure S5: 1H NMR spectra of 2; Figure S6: 13C NMR spectra of 2; Figure S7: ESI-MS spectra of 3; Figure S8: 1H NMR spectra of 3; Figure S9: 13C NMR spectra of 3; Figure S10: ESI-MS spectra of 4; Figure S11: 1H NMR spectra of 4; Figure S12: 13C NMR spectra of 4; Figure S13: ESI-MS spectra of BDP-2BPEA; Figure S14: 1H NMR spectra of BDP-2BPEA; Figure S15: 13C NMR spectra of BDP-2BPEA; Figure S16: ESI-MS spectra of BDP-p-BPEA; Figure S17: 1H NMR spectra of BDP-p-BPEA; Figure S18: 13C NMR spectra of BDP-p-BPEA; Figure S19: ESI-MS spectra of BDP-p-TMPA; Figure S20: 1H NMR spectra of BDP-p-TMPA; Figure S21: 13C NMR spectra of BDP-p-TMPA; Figure S22: The Hill plot of BDP-p-BPEA complexation with Zn2+; Figure S23: ESI-MS spectra of BDP-p-BPEA + Zn2+; Figure S24: (a) confocal imaging of Hela cells treated by BDP-p-BPEA for 1–4 h; (b) average fluorescence intensity of cells in green channel of each graph. Green channel = 570–615 nm, red channel = 650–720 nm, scale bar: 10 µm; Figure S25: Viability of Hela cells versus the concentration of BDP-p-BPEA.

Author Contributions

Conceptualization, investigation, data curation and writing—original draft preparation, Y.L.; validation and formal analysis, S.Y.; writing—review and editing, H.F.; supervision and project administration, Y.C.; funding acquisition and resources, W.H., Y.C., and Z.G. All authors have read and agreed to the published version of the manuscript.

Funding

Z.G. was funded by the National Natural Science Foundation of China (21731004, 92153303, 91953201), the Natural Science Foundation of Jiangsu Province (BK20202004), and the Excellent Research Program of Nanjing University (ZYJH004). Y.C. was funded by the National Natural Science Foundation of China (22122701, 21907050) and the Natural Science Foundation of Jiangsu Province (BK20190282). W.H. was funded by the National Natural Science Foundation of China (21977044). H.F. was funded by the National Postdoctoral Program for Innovative Talents (BX2021123), the China Postdoctoral Science Foundation (2021M691505), and the Jiangsu Postdoctoral Research Funding Program (2021K125B).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Valko, M.; Jomova, K.; Rhodes, C.J.; Kuča, K.; Musílek, K. Redox- and non-redox-metal-induced formation of free radicals and their role in human disease. Arch. Toxicol. 2016, 90, 1–37. [Google Scholar] [CrossRef] [PubMed]
  2. Li, Y.V. Zinc and insulin in pancreatic beta-cells. Endocrine 2014, 45, 178–189. [Google Scholar] [CrossRef] [PubMed]
  3. Costello, L.C.; Franklin, R.B. A comprehensive review of the role of zinc in normal prostate function and metabolism; and its implications in prostate cancer. Arch. Biochem. Biophys. 2016, 611, 100–112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Kerns, K.; Zigo, M.; Drobnis, E.Z.; Sutovsky, M.; Sutovsky, P. Zinc ion flux during mammalian sperm capacitation. Nat. Commun. 2018, 9, 2061. [Google Scholar] [CrossRef] [Green Version]
  5. Paithankar, J.G.; Saini, S.; Dwivedi, S.; Sharma, A.; Chowdhuri, D.K. Heavy metal associated health hazards: An interplay of oxidative stress and signal transduction. Chemosphere 2021, 262, 128350. [Google Scholar] [CrossRef]
  6. Yuan, Y.; Niu, F.; Liu, Y.; Lu, N. Zinc and its effects on oxidative stress in Alzheimer’s disease. Neurol. Sci. 2014, 35, 923–928. [Google Scholar] [CrossRef]
  7. Maret, W. The redox biology of redox-inert zinc ions. Free Radical Bio. Med. 2019, 134, 311–326. [Google Scholar] [CrossRef] [Green Version]
  8. Marreiro, D.d.N.; Climaco Cruz, K.J.; Silva Morais, J.B.; Beserra, J.B.; Severo, J.S.; Soares de Oliveira, A.R. Zinc and oxidative stress: Current mechanisms. Antioxidants 2017, 6, 24. [Google Scholar] [CrossRef]
  9. Aizenman, E.; Stout, A.K.; Hartnett, K.A.; Dineley, K.E.; McLaughlin, B.; Reynolds, I.J. Induction of neuronal apoptosis by thiol oxidation: Putative role of intracellular zinc release. J. Neurochem. 2000, 75, 1878–1888. [Google Scholar] [CrossRef] [Green Version]
  10. Hao, Q.; Maret, W. Aldehydes release zinc from proteins. A pathway from oxidative stress/lipid peroxidation to cellular functions of zinc. FEBS J. 2006, 273, 4300–4310. [Google Scholar] [CrossRef]
  11. McCord, M.C.; Aizenman, E. The role of intracellular zinc release in aging, oxidative stress, and Alzheimer’s disease. Front. Aging Neurosci. 2014, 6, 77. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Wang, F.; Wang, K.; Kong, Q.; Wang, J.; Xi, D.; Gu, B.; Lu, S.; Wei, T.; Chen, X. Recent studies focusing on the development of fluorescence probes for zinc ion. Coord. Chem. Rev. 2021, 429, 213636. [Google Scholar] [CrossRef]
  13. Lin, Y.; Yang, Z.; Lake, R.J.; Zheng, C.; Lu, Y. Enzyme-mediated endogenous and bioorthogonal control of a DNAzyme fluorescent sensor for imaging metal ions in living cells. Angew. Chem. Int. Ed. 2019, 58, 17061–17067. [Google Scholar] [CrossRef] [PubMed]
  14. Zhang, Y.; Xu, S.; Li, X.; Zhang, J.; Sun, J.; Tong, L.; Zhong, H.; Xia, H.; Hua, R.; Chen, B. Improved LRET-based detection characters of Cu2+ using sandwich structured NaYF4@NaYF4:Er3+/Yb3+@NaYF4 nanoparticles as energy donor. Sens. Actuators B Chem. 2018, 257, 829–838. [Google Scholar] [CrossRef]
  15. Dong, H.; Zhao, L.; Chen, Y.; Li, M.; Chen, W.; Wang, Y.; Wei, X.; Zhang, Y.; Zhou, Y.; Xu, M. Dual-ligand near-infrared luminescent lanthanide-based metal–organic framework coupled with in vivo microdialysis for highly sensitive ratiometric detection of Zn2+ in a mouse model of Alzheimer’s disease. Anal. Chem. 2022, 94, 11940–11948. [Google Scholar] [CrossRef]
  16. Liu, R.; Kowada, T.; Du, Y.; Amagai, Y.; Matsui, T.; Inaba, K.; Mizukami, S. Organelle-level labile Zn(2+) mapping based on targetable fluorescent sensors. ACS Sens. 2022, 7, 748–757. [Google Scholar] [CrossRef]
  17. Jiang, X.; Wang, L.; Carroll, S.L.; Chen, J.; Wang, M.C.; Wang, J. Challenges and opportunities for small-molecule fluorescent probes in redox biology applications. Antioxid. Redox. Signal. 2018, 29, 518–540. [Google Scholar] [CrossRef]
  18. 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 Zn(2+) and its application to living cells. Molecules 2022, 27, 1287. [Google Scholar] [CrossRef]
  19. Hu, Z.; Yang, G.; Hu, J.; Wang, H.; Eriksson, P.; Zhang, R.; Zhang, Z.; Uvdal, K. Real-time visualizing the regulation of reactive oxygen species on Zn2+ release in cellular lysosome by a specific fluorescent probe. Sensor. Actuat. B Chem. 2018, 264, 419–425. [Google Scholar] [CrossRef]
  20. Zhu, H.; Fan, J.; Zhang, S.; Cao, J.; Song, K.; Ge, D.; Dong, H.; Wang, J.; Peng, X. Ratiometric fluorescence imaging of lysosomal Zn(2+) release under oxidative stress in neural stem cells. Biomater. Sci. 2014, 2, 89–97. [Google Scholar] [CrossRef]
  21. Xiao, H.; Li, P.; Zhang, S.; Zhang, W.; Zhang, W.; Tang, B. Simultaneous fluorescence visualization of mitochondrial hydrogen peroxide and zinc ions in live cells and in vivo. Chem. Commun. 2016, 52, 12741–12744. [Google Scholar] [CrossRef] [PubMed]
  22. Trusso Sfrazzetto, G.; Satriano, C.; Tomaselli, G.A.; Rizzarelli, E. Synthetic fluorescent probes to map metallostasis and intracellular fate of zinc and copper. Coord. Chem. Rev. 2016, 311, 125–167. [Google Scholar] [CrossRef]
  23. Liu, M.; Ma, S.; She, M.; Chen, J.; Wang, Z.; Liu, P.; Zhang, S.; Li, J. Structural modification of BODIPY: Improve its applicability. Chin. Chem. Lett. 2019, 30, 1815–1824. [Google Scholar] [CrossRef]
  24. Deniz, E.; Isbasar, G.C.; Bozdemir, Ö.A.; Yildirim, L.T.; Siemiarczuk, A.; Akkaya, E.U. Bidirectional switching of near IR emitting boradiazaindacene fluorophores. Org. Lett. 2008, 10, 3401–3403. [Google Scholar] [CrossRef]
  25. Hiruta, Y.; Koiso, H.; Ozawa, H.; Sato, H.; Hamada, K.; Yabushita, S.; Citterio, D.; Suzuki, K. Near IR Emitting red-shifting ratiometric fluorophores based on borondipyrromethene. Org. Lett. 2015, 17, 3022–3025. [Google Scholar] [CrossRef]
  26. Bozdemir, O.A.; Guliyev, R.; Buyukcakir, O.; Selcuk, S.; Kolemen, S.; Gulseren, G.; Nalbantoglu, T.; Boyaci, H.; Akkaya, E.U. Selective manipulation of ICT and PET processes in styryl-bodipy derivatives: Applications in molecular logic and fluorescence sensing of metal ions. J. Am. Chem. Soc. 2010, 132, 8029–8036. [Google Scholar] [CrossRef]
  27. Guliyev, R.; Ozturk, S.; Kostereli, Z.; Akkaya, E.U. From virtual to physical: Integration of chemical logic gates. Angew. Chem. Int. Ed. 2011, 50, 9826–9831. [Google Scholar] [CrossRef] [Green Version]
  28. Xia, S.; Shen, J.; Wang, J.; Wang, H.; Fang, M.; Zhou, H.; Tanasova, M. Ratiometric fluorescent and colorimetric BODIPY-based sensor for zinc ions in solution and living cells. Sensor. Actuat. B Chem. 2018, 258, 1279–1286. [Google Scholar] [CrossRef]
  29. Kawashima, H.; Ukai, S.; Nozawa, R.; Fukui, N.; Fitzsimmons, G.; Kowalczyk, T.; Fliegl, H.; Shinokubo, H. Determinant factors of three-dimensional aromaticity in antiaromatic cyclophanes. J. Am. Chem. Soc. 2021, 143, 10676–10685. [Google Scholar] [CrossRef]
  30. Dai, E.; Pang, W.; Zhang, X.F.; Yang, X.; Jiang, T.; Zhang, P.; Yu, C.; Hao, E.; Wei, Y.; Mu, X.; et al. Synthesis and photophysics of BF2 -rigidified partially closed chain bromotetrapyrroles: Near infrared emitters and photosensitizers. Chem. Asian. J. 2015, 10, 1327–1334. [Google Scholar] [CrossRef]
  31. Zhang, S.; Adhikari, R.; Fang, M.; Dorh, N.; Li, C.; Jaishi, M.; Zhang, J.; Tiwari, A.; Pati, R.; Luo, F.T.; et al. Near-infrared fluorescent probes with large Stokes shifts for sensing Zn(II) ions in living cells. ACS Sens. 2016, 1, 1408–1415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Rae, T.D.; Schmidt, P.J.; Pufahl, R.A.; Culotta, V.C.; O’Halloran, T.V. Undetectable intracellular free copper: The requirement of a copper chaperone for superoxide dismutase. Science 1999, 284, 805–808. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Malaiyandi, L.M.; Dineley, K.E.; Reynolds, I.J. Divergent consequences arise from metallothionein overexpression in astrocytes: Zinc buffering and oxidant-induced zinc release. Glia 2004, 45, 346–353. [Google Scholar] [CrossRef] [PubMed]
  34. Kannan, N.; Nguyen, L.V.; Makarem, M.; Dong, Y.; Shih, K.; Eirew, P.; Raouf, A.; Emerman, J.T.; Eaves, C.J. Glutathione-dependent and -independent oxidative stress-control mechanisms distinguish normal human mammary epithelial cell subsets. Proc. Natl. Acad. Sci. USA 2014, 111, 7789–7794. [Google Scholar] [CrossRef] [Green Version]
  35. Yu, W.; Chen, Y.; Dubrulle, J.; Stossi, F.; Putluri, V.; Sreekumar, A.; Putluri, N.; Baluya, D.; Lai, S.Y.; Sandulache, V.C. Cisplatin generates oxidative stress which is accompanied by rapid shifts in central carbon metabolism. Sci. Rep. 2018, 8, 4306. [Google Scholar] [CrossRef] [Green Version]
  36. Chang, C.J.; Jaworski, J.; Nolan, E.M.; Sheng, M.; Lippard, S.J. A tautomeric zinc sensor for ratiometric fluorescence imaging Application to nitric oxide-induced release of intracellular zinc. Proc. Natl. Acad. Sci. USA 2003, 101, 1129–1134. [Google Scholar] [CrossRef] [Green Version]
  37. Wei, T.; Chen, C.; Hou, J.; Xin, W.; Mori, A. Nitric oxide induces oxidative stress and apoptosis in neuronal cells. Biochim. Biophys. Acta 2000, 1498, 72–79. [Google Scholar] [CrossRef] [Green Version]
  38. Fang, M.; Xia, S.; Bi, J.; Wigstrom, T.P.; Valenzano, L.; Wang, J.; Tanasova, M.; Luck, R.L.; Liu, H. Detecting Zn(II) ions in live cells with near-infrared fluorescent probes. Molecules 2019, 24, 1592. [Google Scholar] [CrossRef] [Green Version]
  39. Xu, H.; Zhu, C.; Chen, Y.; Bai, Y.; Han, Z.; Yao, S.; Jiao, Y.; Yuan, H.; He, W.; Guo, Z. A FRET-based fluorescent Zn(2+) sensor: 3D ratiometric imaging, flow cytometric tracking and cisplatin-induced Zn(2+) fluctuation monitoring. Chem. Sci. 2020, 11, 11037–11041. [Google Scholar] [CrossRef]
  40. Wang, X.; Bai, X.; Su, D.; Zhang, Y.; Li, P.; Lu, S.; Gong, Y.; Zhang, W.; Tang, B. Simultaneous fluorescence imaging reveals N-methyl-D-aspartic acid receptor dependent Zn(2+)/H(+) flux in the brains of mice with depression. Anal. Chem. 2020, 92, 4101–4107. [Google Scholar] [CrossRef]
  41. Fang, H.; Li, Y.; Yao, S.; Geng, S.; Chen, Y.; Guo, Z.; He, W. An endoplasmic reticulum-targeted ratiometric fluorescent molecule reveals Zn(2+) micro-dynamics during drug-induced organelle ionic disorder. Front. Pharmacol. 2022, 13, 927609. [Google Scholar] [CrossRef] [PubMed]
  42. Feng, J.; Li, J.Z.; Mao, X.M.; Wang, Q.; Li, S.P.; Wang, C.Y. Real-time detection and imaging of exogenous and endogenous Zn(2+) in the PC12 cell model of depression with a NIR fluorescent probe. Analyst 2021, 146, 3971–3976. [Google Scholar] [CrossRef] [PubMed]
  43. Miao, W.; Dai, E.; Sheng, W.; Yu, C.; Hao, E.; Liu, W.; Wei, Y.; Jiao, L. Direct synthesis of dipyrrolyldipyrrins from SNAr reaction on 1,9-dihalodipyrrins with pyrroles and their NIR fluorescence “turn-on” response to Zn2+. Org. Lett. 2017, 19, 6244–6247. [Google Scholar] [CrossRef] [PubMed]
  44. Fang, L.; Trigiante, G.; Crespo-Otero, R.; Hawes, C.S.; Philpott, M.P.; Jones, C.R.; Watkinson, M. Endoplasmic reticulum targeting fluorescent probes to image mobile Zn2+. Chem. Sci. 2019, 10, 10881–10887. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Synthesis route of the three probes.
Scheme 1. Synthesis route of the three probes.
Chemosensors 10 00477 sch001
Figure 1. (ac) Absorption spectra of three probes (5 µM) before (black line) and after (red line) being supplemented with ZnCl2 solution, one equivalent of ZnCl2 for BDP-p-BPEA and BDP-p-TMPA, and two equivalents for BDP-2BPEA. (a) BDP-2BPEA; (b) BDP-p-BPEA; (c) BDP-p-TMPA; (df) fluorescence spectra of three probes (5 µM) titrated with ZnCl2 solution, λex = 550 nm; (d) BDP-2BPEA; (e) BDP-p-BPEA; (f) BDP-p-TMPA. All the spectra of the probe were tested in acetonitrile/aqueous media (v/v = 1:1, pH 7.4, 0.05 M HEPES with 0.10 M KNO3); Inset: ratio value of fluorescence intensity against the concentration of Zn2+, n = 3.
Figure 1. (ac) Absorption spectra of three probes (5 µM) before (black line) and after (red line) being supplemented with ZnCl2 solution, one equivalent of ZnCl2 for BDP-p-BPEA and BDP-p-TMPA, and two equivalents for BDP-2BPEA. (a) BDP-2BPEA; (b) BDP-p-BPEA; (c) BDP-p-TMPA; (df) fluorescence spectra of three probes (5 µM) titrated with ZnCl2 solution, λex = 550 nm; (d) BDP-2BPEA; (e) BDP-p-BPEA; (f) BDP-p-TMPA. All the spectra of the probe were tested in acetonitrile/aqueous media (v/v = 1:1, pH 7.4, 0.05 M HEPES with 0.10 M KNO3); Inset: ratio value of fluorescence intensity against the concentration of Zn2+, n = 3.
Chemosensors 10 00477 g001
Figure 2. Equilibration time, reversibility, and selectivity of BDP-p-BPEA. (a) ratio of fluorescence intensity F670/F605 of 5 µM BDP-p-BPEA against the time after the addition of 5 µM ZnCl2; (b) ratio of fluorescence intensity F670/F605 of 5 µM BDP-p-BPEA with alternative addition of 5 µM ZnCl2 and TPEN; (c) normalized ratio of fluorescence intensity (F670/F605)/(F670 probe/F605 probe) of 5 µM BDP-p-BPEA in the present of 1. blank 2. Cu2+ 3. Fe2+ 4. Fe2++ H2O2 5. H2O2 6. Fe3+ 7. Hg2+ 8. Pb2+ 9. Cd2+ 10. Mn2+ 11. Ni2+ 12. 100 eq Na+ +100 eq K+ +10 eq Ca2+ +10 eq Mg2+ 13. Co2+ 14. Al3+. All the species were presented as 5 µM (1 eq) if not mentioned, n = 3; (d) job’s plot of BDP-p-BPEA, Χprobe = c(BDP-p-BPEA)/([Zn2+] + c(BDP-p-BPEA)), and the total concentration was kept as 10 µM.
Figure 2. Equilibration time, reversibility, and selectivity of BDP-p-BPEA. (a) ratio of fluorescence intensity F670/F605 of 5 µM BDP-p-BPEA against the time after the addition of 5 µM ZnCl2; (b) ratio of fluorescence intensity F670/F605 of 5 µM BDP-p-BPEA with alternative addition of 5 µM ZnCl2 and TPEN; (c) normalized ratio of fluorescence intensity (F670/F605)/(F670 probe/F605 probe) of 5 µM BDP-p-BPEA in the present of 1. blank 2. Cu2+ 3. Fe2+ 4. Fe2++ H2O2 5. H2O2 6. Fe3+ 7. Hg2+ 8. Pb2+ 9. Cd2+ 10. Mn2+ 11. Ni2+ 12. 100 eq Na+ +100 eq K+ +10 eq Ca2+ +10 eq Mg2+ 13. Co2+ 14. Al3+. All the species were presented as 5 µM (1 eq) if not mentioned, n = 3; (d) job’s plot of BDP-p-BPEA, Χprobe = c(BDP-p-BPEA)/([Zn2+] + c(BDP-p-BPEA)), and the total concentration was kept as 10 µM.
Chemosensors 10 00477 g002
Scheme 2. (a) 1H NMR spectrum (400 MHz) of BDP-p-BPEA in DMSO-d6 before (up) and after (down) the addition of ZnCl2; (b) the coordination mode of BDP-p-BPEA-Zn2+.
Scheme 2. (a) 1H NMR spectrum (400 MHz) of BDP-p-BPEA in DMSO-d6 before (up) and after (down) the addition of ZnCl2; (b) the coordination mode of BDP-p-BPEA-Zn2+.
Chemosensors 10 00477 sch002
Figure 3. The colocalization experiment of BDP-p-BPEA with Mito-Tracker Green, Lyso-Tracker Green, ER-Tracker Green, Golgi-Tracker Green and lipid droplet targeting commercial dye BODIPY 493/503, respectively. For BDP-p-BPEA, λex = 543 nm, λem = 570–615 nm, for commercial dyes, λex = 488 nm, λem = 500–550 nm, scale bar: 10 µm.
Figure 3. The colocalization experiment of BDP-p-BPEA with Mito-Tracker Green, Lyso-Tracker Green, ER-Tracker Green, Golgi-Tracker Green and lipid droplet targeting commercial dye BODIPY 493/503, respectively. For BDP-p-BPEA, λex = 543 nm, λem = 570–615 nm, for commercial dyes, λex = 488 nm, λem = 500–550 nm, scale bar: 10 µm.
Chemosensors 10 00477 g003
Figure 4. (a) Cells incubated with 5 μM BDP-p-BPEA for four hours and supplemented with DTDP, NEM, TPEN (for 0.5 h), DEA-NONOate, and cisplatin (for 2 h) separately before acquiring fluorescence imaging. Green Channel = 570–615 nm, red channel = 650–720 nm, scale bar: 10 µm; (b) average fluorescence ratio value of cells in each graph, ratio = Fred/Fgreen, ** p < 0.01 and *** p < 0.001.
Figure 4. (a) Cells incubated with 5 μM BDP-p-BPEA for four hours and supplemented with DTDP, NEM, TPEN (for 0.5 h), DEA-NONOate, and cisplatin (for 2 h) separately before acquiring fluorescence imaging. Green Channel = 570–615 nm, red channel = 650–720 nm, scale bar: 10 µm; (b) average fluorescence ratio value of cells in each graph, ratio = Fred/Fgreen, ** p < 0.01 and *** p < 0.001.
Chemosensors 10 00477 g004
Table 1. Performance comparison between BDP-p-BPEA and other Zn2+ probes.
Table 1. Performance comparison between BDP-p-BPEA and other Zn2+ probes.
Detection Modeλex (nm)/λem (nm)FluorophoreLinear Range (µM)LOD (nM)Testing MediaKd(nM)Ref.
Enhanced750/779cyanine0.1–5.01125 mM HEPES1.42 × 103[21]
Enhanced687/703hemicyanine0–1.40.45HEPES buffer solution (10 mM, pH 7.0)1.05 × 103[38]
Ratiometric415/560 to 480coumarin & benzo[c][1,2,5]oxadiazole0–582DMSO: aqueous solution (50 mM HEPES 100 mM KNO3, pH = 7.40) v:v = 1:990.014[39]
Enhanced390/460coumarin & rhodamine0–103.2×1020.1 M Tris, pH 7.43.94 × 103[40]
Ratiometric610 to 575/700 to 660BODIPY0–1031.8DMSO: aqueous solution (50 mM HEPES 100 mM KNO3, pH = 7.2) v:v = 3:23.65[41]
Ratiometric430 to 475/660isophorone0–8106DMSO: aqueous solution (10 mM HEPES, pH = 7.46) v:v = 1:17.7×103[42]
Enhanced544 to 607/634dipyrrolyldipyrrin(No data)210DMF13[43]
Enhanced346/414naphthalimide0–0.0030.047DMSO: aqueous solution (10 mM HEPES, pH 7.4) v:v = 1:94.70[44]
Ratiometric550/620 to 650BODIPY0–329MeCN: aqueous solution (50 mM HEPES 100 mM KNO3, pH 7.4) v:v = 1:13.2This work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, Y.; Yao, S.; Fang, H.; He, W.; Chen, Y.; Guo, Z. Rational Design of Ratiometric Fluorescent Probe for Zn2+ Imaging under Oxidative Stress in Cells. Chemosensors 2022, 10, 477. https://doi.org/10.3390/chemosensors10110477

AMA Style

Li Y, Yao S, Fang H, He W, Chen Y, Guo Z. Rational Design of Ratiometric Fluorescent Probe for Zn2+ Imaging under Oxidative Stress in Cells. Chemosensors. 2022; 10(11):477. https://doi.org/10.3390/chemosensors10110477

Chicago/Turabian Style

Li, Yaheng, Shankun Yao, Hongbao Fang, Weijiang He, Yuncong Chen, and Zijian Guo. 2022. "Rational Design of Ratiometric Fluorescent Probe for Zn2+ Imaging under Oxidative Stress in Cells" Chemosensors 10, no. 11: 477. https://doi.org/10.3390/chemosensors10110477

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop