Next Article in Journal
Diels–Alder Adducts of Morphinan-6,8-Dienes and Their Transformations
Next Article in Special Issue
A Quinoxaline−Naphthaldehyde Conjugate for Colorimetric Determination of Copper Ion
Previous Article in Journal
Enhanced Oral Bioavailability of β-Caryophyllene in Healthy Subjects Using the VESIsorb® Formulation Technology, a Novel Self-Emulsifying Drug Delivery System (SEDDS)
Previous Article in Special Issue
Ultrafast Excited-State Decay Mechanisms of 6-Thioguanine Followed by Sub-20 fs UV Transient Absorption Spectroscopy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Turn on Fluorescence Sensing of Zn2+ Based on Fused Isoindole-Imidazole Scaffold

1
Department of Chemistry, University of Calcutta, 92, A.P.C. Road, Kolkata 700009, India
2
Department of Chemistry, The Bhawanipur Education Society College, 5, LalaLajpat Rai Sarani, Kolkata 700020, India
3
Departamento de Química, Universidad Autónoma de Madrid, 28049 Madrid, Spain
4
Centro Universitario de la Defensa, Academia General del Aire, Universidad Politécnica de Cartagena, C/Coronel López Peña S/N, Santiago de La Ribera, 30720 Murcia, Spain
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(9), 2859; https://doi.org/10.3390/molecules27092859
Submission received: 22 February 2022 / Revised: 13 April 2022 / Accepted: 25 April 2022 / Published: 30 April 2022
(This article belongs to the Special Issue Photophysics and Photochemistry in Complex Molecular Systems)

Abstract

:
Optical chemosensors caused a revolution in the field of sensing due to their high specificity, sensitivity, and fast detection features. Imidazole derivatives have offered promising features in the literature as they bear suitable donor/acceptor groups for the selective analytes in the skeleton. In this work, an isoindole-imidazole containing a Schiff base chemosensor (1-{3-[(2-Diethylamino-ethylimino)-methyl]-2-hydroxy-5-methyl-phenyl}-2H-imidazo[5,1-a]isoindole-3,5-dione) was designed and synthesized. The complete sensing phenomena have been investigated by means of UV-Vis, fluorescence, lifetime measurement, FT-IR, NMR and ESI-MS spectroscopic techniques. The optical properties of the synthesized ligand were investigated in 3:7 HEPES buffer:DMSO medium and found to be highly selective and sensitive toward Zn2+ ion through a fluorescence turn-on response with detection limit of 0.073 μm. Furthermore, this response is effective in gel form also. The competition studies reveal that the response of the probe for Zn2+ ion is unaffected by other relevant metal ions. The stoichiometric binding study was performed utilizing Job’s method which indicated a 1:1 sensor–Zn2+ ensemble. Computational calculations were performed to pinpoint the mechanism of sensing.

Graphical Abstract

1. Introduction

Metal ions are deeply rooted in both environment and biological systems [1,2,3,4]. Although required in many fundamental processes, their quantitative and qualitative analysis is a critical measure because negative effects are associated if present in excess. In that framework, optical chemosensors are the easiest way to detect analytes as they open the door for cost–effective and fast detection methods [5,6,7]. Among optical chemosensors, fluorescent species are widely used for assessing ions due to their selectivity, high sensitivity, and low detection ability [8,9,10].
The present contribution brings the focus towards the detection of Zn2+. That ion is mostly found in biological systems along with other metal ions including Fe3+, Mg2+, Mn2+ [11,12]. However, Zn2+ specifically participates in a wide range of cellular functions in human body, e.g., cofactor of carbonic anhydrase and zinc-finger proteins (ZNFs), DNA synthesis, RNA transcription, regulation of metalloenzymes, neurophysiology, and apoptosis [13,14,15,16,17]. Zn2+ also plays a core role in brain functionality, a region that accumulates the highest concentration (150 mM–300 nM) as part of the synaptic transmission machinery [18,19,20,21,22,23,24,25]. Other fundamental biological actions of Zn2+ have been detected in immune system and cell growth, cell division, proper development of neutrophils, and as natural killer cells [26,27,28,29,30,31]. As expected, negative effects appear if Zn2+ is out of the natural ratio. Low levels of Zn2+ might lead to serious health concerns such us difficulty in wound healing, alopecia, diarrhea, poor growth, dysfunction of the immune and nervous system with congenital abnormalities and blood disorders, to cite only a few [32]. On the contrary, accumulation of Zn2+ in our bodies might cause anemia, epilepsy, and Alzheimer’s disease [33]. As other ions, Zn2+ also impacts into environment, so that its bioavailability in soil must be monitored. For instance, it has been shown that an excess in Zn2+ concentration induces phytotoxic effect that are concomitant with the degradation of crop quality, which in turn indirectly affects the human life [34,35,36,37]. Unfortunately, due to the filled d orbital of Zn2+, its visual detection is still scarce [38,39]. The available chemosensors for Zn2+ are limited to naphthalene [40], quinoline [41,42], coumarin [43], rhodamine [44], flavonol [45], benzothiazole [46], phenanthroline [47], julolidine [48] building blocks. The determination of Zn2+ remains an essential issue to ensure that its concentration is tuned at the optimal biological and environmental level.
Up until now, the use of imidazole has been less exploited even if those organic molecules containing heterocyclic moiety are excellent as both cation and anion sensor signaling unit [49,50,51]. The acidic NH proton can interact with anion, whereas the nitrogen atom present in the ring can bind the cations [52]. The commercial availability of imidazole derivatives is a plus their use as chemosensors. Indeed, isoindole have been used in the development of drugs [53,54,55,56].
Herein, we report a fused isoindole imidazole fluorescent Schiff base chemosensor (1-{3-[(2-Diethylamino-ethylimino)-methyl]-2-hydroxy-5-methyl-phenyl}-2H-imidazo[5,1-a]isoindole-3,5-dione), which is labelled as IIED. Our hypothesis is based on the exploitation of an imidazole scaffold with Schiff bases, which are widely used in the field of metal ion detection [57,58,59,60,61,62,63,64]. The performed experimental and theoretical work demonstrated that IIED can detect Zn2+ with a limit of 0.073 μM. Aiming at maximizing the use of that new probe, its response in gel form was also assessed.

2. Results

2.1. Synthesis and Characterization

The probe IIED (1-{3-[(2-Diethylamino-ethylimino)-methyl]-2-hydroxy-5-methyl-phenyl}-2H-imidazo[5,1-a]isoindole-3,5-dione) was synthesized by condensing 3-(3,5-Dioxo-2,5-dihydro-3H-imidazo[5,1-a]isoindol-1-yl)-2-hydroxy-5-methyl-benzaldehyde (compound 2) and N,N-Diethylethylenediamine. Compound 1 (1-(2-Hydroxy-5-methyl-phenyl)-2H-imidazo[5,1-a]isoindole-3,5-dione) was prepared from the standard procedure available elsewhere (Figure S1) [65]. Compound 2 was prepared by Duff reaction as previously reported by Chang et al. [66]. In the final step reaction, the Schiff base condensation product IIED was found in a powdery light orange form with a percentage of yield 57.8% (Scheme 1). The preparation of compound 1 and compound 2 were monitored by 1H NMR spectroscopic studies (Figures S1 and S2). The final step compound IIED were characterized by 1H NMR, 13C NMR, FT–IR, ESI–MS spectroscopic study (Figures S3–S6). FT-IR spectrum confirms the peak at 1642 cm−1 for the generation of imine bond (C=N). The 1H NMR peak at δH 8.53 is for aldehydic proton signal, δH 7.81–7.31 peaks are for the aromatic protons and and δH 3.6, 2.6, 2.5, 2.2, 0.95. peaks appeared for aliphatic protons. ESI-MS spectrum shows peak at 419.22 amu for [C24H27N4O3]+species. Further synthetic details can be found in Section 3 and in the Supporting Information.

2.2. Photophysical Studies of IIED towards Zn2+

The entire spectral behaviour was assessed in 3:7 HEPES buffer:DMSO medium (Figure 1). The UV–Visible spectra of the probe IIED (10−5 M) show peaks at 405 nm and 480 nm. The UV–Visible spectra of the probe in presence of various metal ions (Na+, K+, Ca2+, Mg2+, Hg2+, Ni2+, Fe3+, Cu2+, Co2+, Cd2+, Zn2+, Mn2+, Pb2+, Al3+, Cr3+) were recorded. The probe IIED changes its appearance from colorless to very pale yellow in the presence of Cu2+, Zn2+, Hg2+ ions a new peak at 467 nm had appeared for these three cations (Figure S8).
The UV–Visible titration experiment of IIED was carried out for Zn2+ by the gradual addition of Zn2+ up to 2.5 µM. The absorbance peak at 467 nm gradually increased with the presence of isosbestic point at 419 nm which indicates that only one equilibrium is present between the IIED and IIED in presence of Zn2+ (Figure 2).
Benesi–Hildebrand (B–H) relation has been subsequently combined with the absorption titration spectral outputs according to Equation (1) [67]:
1 ( A A 0 ) = 1 { K ( A max A 0 ) [ C ] } + 1 ( A max A 0 )
where, A0 is the absorbance of ligand IIED only, A is the observed absorbance at that distinct wavelength in the presence of a particular concentration of the metal ion [C], Amax is the maximum absorbance value of the complex formed. K is the association constant (M−1) which was calculated from the slope of the linear plot and [C] is the concentration of the Zn2+ ion added during titration studies. The linear fit of the B–H plot of 1/(A − A0) vs. 1/[Zn2+] indicates 1:1 complex formation (Figure S12) between IIED and Zn2+ and association constant was found to be 1.2 × 104 M−1 for absorbance. The stoichiometric ratio using absorbance value was further confirmed by Job’s plot analyses. The molar fraction of the ligand was shown the highest value 0.51, which also indicates 1:1 stoichiometry between Zn2+ and IIED (Figure S17). All these accumulated measures back up that the novel probe accommodates one Zn2+ ion only, an experimental finding that guide our molecular models.
A larger dissimilarity is observed for the fluorescence regime. We first confirmed that the emission spectra of the probe IIED (10−6 M) show no peak intensity (non–fluorescent) upon excitation at 480 nm. The emission spectra of the probe was subsequently measured in presence of all selected metal ions (Na+, K+, Ca2+, Mg2+, Hg2+, Ni2+, Fe3+, Cu2+, Co2+, Cd2+, Zn2+, Mn2+, Pb2+, Al3+, Cr3+). Emission signatures were recorded in 3:7 HEPES buffer solution:DMSO. Under UV light the probe shows dark to green emission upon addition of Zn2+. This optical response can be recognized by naked eye, a result that was not observed with any other ion. Such qualitative conclusion was confirmed by a quantitative analysis of the emission signatures. The probe IIED turns on fluorescence intensity at 558 nm (λex = 480 nm) in the presence of Zn2+ ion only, among other cations mentioned above. It should be underlined that a complete competitive study in emission had been performed with IIED (Figures S9 and S10). The use of our probe with other metal ions demonstrated the selective output for Zn2+.
The fluorescence titration of IIED in presence of an increasing concentration of Zn2+ had been performed by the gradual addition of Zn2+ up to 25 µM in 3:7 HEPES buffer solution:DMSO. As illustrated in Figure 3, the 558 nm characteristic peak of the probe with Zn2+ directly correlates with Zn2+ concentration.
The emission spectral response towards IIED in presence of various counter anions as salts of Zn2+ (Zn(ClO4)2, ZnCl2, ZnBr2, Zn (NO3)2 and Zn(OAc)2) shows peaks at 558 nm, but the intensity varies slightly indicating negligible counter anion effect (Figure S11). This is a remarkable finding: IIED is proved to be versatile enough to reveal the presence of Zn2+ with no restriction on the parent-salt pair.
Consistently with absorption titration, emission titration spectral output was combined with a Benesi–Hildebrand (B–H) equation to find out the association constant (K) upon the complexation of IIED with Zn2+ ion according to Equation (2):
1 F X F 0 = 1 F max F 0 + 1 K [ C ] ( 1 F max F 0 )
where F0 is the emission of ligand IIED only, Fx is the emission of ligand at an intermediate Zn2+ concentration, and Fmax is the maximum emission value of the complex formed. K is the binding constant and [C] is the concentration of Zn2+ ions. According to B–H expression, the measured emission [(Fmax – F0)/(Fx – F0)] at 558 nm varied as a function of 1/[Zn2+] in a linear relationship, also indicated the formation of 1:1 stoichiometry between Zn2+and probe IIED. The association constant is found to be 3 × 104 M−1 for emission (Figure S13), a results that is consistent with the predicted value by using absorbance, e.g., 1.2 × 104 M−1 (see above).
The fluorescence quantum yield (Φ) was determined using quinine sulphate in 0.05 M H2SO4 as reference using the following Equation (3) [68]:
ϕ s = ϕ R A s A R × Abs R Abs S × η S 2 η R 2
where A indicates the integrated area under the fluorescence curve, Abs stand for absorbance, η is the refractive index of the medium and Φ is the fluorescence quantum yield. Subscripts S and R stand for the respective specification for the studied sample and reference, respectively. The probe exhibits low fluorescence emission peak at 558 nm with quantum yield Φ = 0.036 upon excitation at 480 nm but in Zn2+ environment the quantum yield rises to Φ = 0.69.
The emission intensity of probe IIED in presence of Zn2+ was assessed at several pH values in the range of 4–10. In acidic region (pH < 6.0) there is no fluorescence enhancement in presence of Zn2+. On the contrary, in basic region (pH > 9.0) emission intensity gradually deceases. The effectivity of Zn2+ was observed in the range pH 7–8 and highest in pH 8. These findings suggest that the complexation phenomenon is optimal (i.e., the most stable) at the physiological pH region (Figure S14).
The limit of detection has been evaluated from the titration experiments for both absorption and emission study, using the Equation (4):
DL = K × σ S
where a threshold of K = 3 is imposed to ensure an acceptable signal-to-noise rate, and σ is the standard deviation of the blank solution and S is the slope of the calibration curve [69]. The calculated limit of detection for emission study of IIED with Zn2+ was 0.073 μM while for the absorption regime was 0.29 μm (Figures S15 and S16).
The fluorescence study is eventually completed by recording lifetime for IIED parent probe and its IIED + Zn2+ counterpart. For the records, these measure have been carried out in 3:7 HEPES buffer solution:DMSO at 480 nm, 298 K. Average fluorescence lifetimes (τavg) were calculated from the decay times and pre–exponential factors using the following Equation (5):
τ avg = α i τ i 2 α i τ i
where αi is the pre–exponential factor corresponding to the ith decay time constant, τi.
The radiative decay rate constant kr and the total nonradiative decay rate constant knr of IIED and IIED + Zn2+ complex was calculated according to the equations τ−1 = kr + knr and kr = Φf. The fluorescence decay curves for IIED and its Zn2+ complex were fitted by three exponential functions. From the decay curve and fitting data, the average fluorescence lifetime (τ) of IIED and IIED + Zn2+ were estimated 2.74 ns and 3.65 ns. The values of all the fitting data were shown in Figure 4 and Table S1. The values of radiative and nonradiative decay rate constants for IIED indicating the nonradiative decay is the predominant process in the excited states which goes indeed reverse in the presence of Zn2+ where radiative decay is the predominant one, resulting in a strong fluorescence response.

2.3. Theoretical Calculations

As illustrated in Figure 5, two alternative structures become possible upon the Zn2+ coordination to IIED, labeled as L(N) and L(NH) according to their protonation state. Despite all accumulated experimental evidence, the participation of -NH proton in the binding of Zn2+ is not confirmed by the experimental optical signatures.
Density functional theory (DFT) and time dependent-DFT (TD-DFT) calculations were performed to determine the main form responsible of the measured photophysical activity. All structures were first fully optimized without imposing any symmetry restriction. We considered both the neutral (LN-) and the N-deprotonated (LNH) species for the free ligand. The computed pKa was in the range of 7.2–7.4. Such a value is significantly below the expected range for lactams (pKa = 11–18), which can be explained by the stabilization of the N- site by a hydrogen bond with the phenol group. This hypothesis was confirmed by analyzing a conformer where the hydrogen bond formation was impossible, leading to a pKa of ~15. We further checked that the O-deprotonated compound was less stable in DMSO solution than the N-deprotonated one. The spectra of both the neutral and deprotonated ligands were computed using the optimized structures in the gas phase (Figure 6). A phenomenological broadening of 0.25 eV was applied to the 30 computed transitions.
Our simulations indicate that both the neutral and deprotonated forms give rise to the experimental spectrum. Concretely, the deprotonated form leads to a low energy band, ~0.4 eV (~70 nm) below the neutral form, and it is consistent with the shoulder observed in the experimental spectrum around 500 nm. Indeed, a semi-quantitative agreement is achieved by combining both spectra with the deprotonated form scaled by 0.5. This scaling corresponds to a dissociation fraction of 0.333. Considering that the pH is buffered at 7.4, such a ratio would correspond to a pKa = 7.7, which agrees very well with our theoretical estimation.
We next turned to the complexes, considering in all cases that Zn binds to the phenol oxygen, replacing the hydrogen. In contrast, the amide site is considered both protonated (LNH) and deprotonated (LN-). The latter lead to a tetradentate chelate, while LNH binds Zn cation by three sites, as shown in Figure 7.
The spectra for the free and complex compounds with N-protonated sites is shown in Figure 8. The complex shows a large pKa (>15), typical for lactams; the contribution of the N-deprotonated complex is consequently not expected. Although the predicted shift is smaller than the observed signatures, which is the consequence of introducing the metal Zn into the model system, our simulations correctly mimic a red-shift of the complex compared to the free ligand, agreeing with the experimental observation.

2.4. Sensing Mechanism

The absorption profile of Compound 1 was originally characterized by Ray et al., which exhibits two absorption peaks at 320 nm and 395 nm [70]. IIED is a modified version of Compound 1 where we introduced an imine (-C=N) moiety and also increasing the conjugation. As a consequence, these absorption peaks are red shifted at 405 nm and 480 nm in the novel IIED probe. That trend is correctly reproduced by our TD-DFT calculations. As discussed above, the IIED showed no emission band in 3:7 HEPES buffer:DMSO solution upon excitation at 480 nm while fluorescence is switched on after Zn2+ coordination. That dissimilarity in the emission arises from a ESIPT (Excited State Intramolecular Proton Transfer) process. Indeed, earlier works demonstrated that probes with hydroxy-imine moieties are associated to lower fluorescence because of ESIPT equilibria [71]. Such scenario is correctly reproduced by our computational approach, which shows that the spectroscopic transition is characterized by a partial transfer of electron density from the phenolic oxygen to the imine nitrogen, as concluded from the inspection of the molecular orbitals involved (Figure S19). That electronic rearrangement activates the ESIPT and consequently reduces fluorescence. The substitution of the proton by the metal in the IIED + Zn2+ complex cancels the ESIPT and results in the chelation enhanced fluorescence (CHEF) phenomena at 558 nm.

2.5. Reversibility and Application in Gel Phase

Our main goal is to report a probe with application in real scenarios. A critical prerequisite for optimizing the use of IEED, which was performed by using sodium salt of ethylenediaminetetraacetic acid (Na2EDTA) solution in emission spectra. In the presence of EDTA, the 558 nm peak of the IIED + Zn2+ complex had been quenched and reappear upon addition of Zn2+ solution again (Figure S18). The process has been repeated for few cycles, indicating a reversible coordination between IIED and Zn2+.
Finally, we have also checked the stability of our probe in gel medium. The gel has been prepared as the method described in literature using Poloxamer 407 [72,73]. In this gel phase also the probe IIED shows change from dark to green fluorescence under UV light in presence of Zn2+ and the naked eye output is shown in Figure 9.

3. Materials and Methods

3.1. Materials and Instruments

Ninhydrin, p–cresol, HMTA (Hexamethylenetetramine), TFA (Trifluoroacetic acid), N, N-Diethylethylenediamine was purchased from Sigma–Aldrich, Kolkata, India, urea purchased from Merck, Kolkata, India and all were used as received. The salts of the cations were also purchased from Sigma Aldrich, Kolkata, India. Solvents for the syntheses were purchased from commercial sources and used as received. 1H and 13C NMR spectrum were recorded in DMSO-d6 with TMS as internal standard on a Bruker, AV300 Supercon Digital NMR system with dual probe. The FT–IR spectra were recorded from KBr pellets in the range of 400–4000 cm−1 on a Perkin-Elmer Spectrum 100 spectrometer. Elemental analyses for C, H and N were performed on a Perkin–Elmer 2400 II analysers. The ESI-MS experiments were performed on Waters Xevo G2-S QTOF mass spectrometer. The absorption and emission spectral studies were performed on a Hitachi UV–Vis U–3501 spectrophotometer and a Perkin–Elmer LS55 fluorimeter, respectively. Time-resolved fluorescence lifetime measurement were performed in Horiba Jobin Yvon Fluorocube-01-NLtime-correlated single photon counting (TCSPC) set up with picosecond delta diode (DD–375L) operating at λex = 480 nm and a repetition rate of 1 MHz as excitation source.

3.2. Computational Details

Computational methods were implemented to predict the absorption spectra of both the new IIED probe as well as two the possible complexes after Zn2+ complexation, which were labelled as L(N) and L(NH) depending on their protonation state. These calculations were conducted in the framework of the density functional theory (DFT) and the time dependent (TD-DFT) schemes by using the PCM-PBE0/cc-pVTZ level of theory [74,75]. For the records, the addition of a basis with pseudopotentials was also assessed. The use of LANL2TZ+ on the Zn2+ center has a minor effect on the structures, e.g., predicts the same chelation as that not using pseudopotentials. The nature of all located structures was confirmed as real minima (stable) forms by analyzing the associated vibrational modes. The absence of imaginary frequencies confirm that all optimized structures are real minima in the potential energy surface rather than saddle points. Geometries were eventually used to assess the vertical transition energies used for predicting the absorption spectra of these species. The pKa values were evaluated at two levels of theory. Our first choice is based on the protocol described by Rossini et al. [76], computing frequencies at B3LYP/6-31G(d,p) and single point energies with PCM at B3LYP/cc-pVQZ [77]. For the free energy of the proton in gas phase a value of 6.28 kcal/mol was used, while for the solvation free energy of the proton in DMSO, we adopted the theoretical estimation provided by Rossini et al., −266.4 kcal/mol [76]. The empirical pKa prediction module by Schrödinger is also used as a second theoretical framework [78,79,80]. These schemes yield to very similar pKa values, 7.2 and 7.4, respectively. All DFT and TD-DFT calculations were performed with Gaussian 16 [81].

3.3. Synthesis of the Ligand

The IIED synthesized in three simple steps. The 1-(2-Hydroxy-5-methyl-phenyl) -2H-imidazo[5,1-a]isoindole-3,5-dione (Compound 1) was prepared by the procedure found in literature [65]. The 3-(3,5-Dioxo-2,5-dihydro-3H-imidazo[5,1-a]isoindol-1-yl)-2-hydroxy-5-methyl-benzaldehyde (Compound 2) was prepared by Duff reaction mechanism by HMTA and TFA found in literature [66]. In methanolic solution of Compound 2 (160 mg, 0.5 mmol), N,N-Diethylethylenediamine (72 µL, 0.5 mmol) was added dropwise and the color changes from yellow to orange. Then the solution was refluxed for 4 hrs and light orange colored precipitate was filtered through suction filtration and air dried. yield = 57.8% (120 mg, 0.29 mmol). Anal. calc. for C24H26N4O3: C, 68.88; H, 6.26; N, 13.39; Found: C, 68.7; H, 6.24; N, 13.4; 1H NMR (300 MHz, DMSO-d6, 290 K, TMS) δH(ppm): 8.53 (s, 1 H), 7.81−7.59 (m, 2 H), 7.45−7.31 (m, 4 H), 3.6(t, 2 H), 3.51 (s, −NH), 2.6 (t, 2 H), 2.52(q, 4 H), 2.27(s, 3 H), 0.95(m, 6 H); 13C NMR (75 MHz, DMSO–d6) δC: 160.58, 158.79, 156.03, 142.37, 129.38, 128.97, 128.48, 127.47, 126.52, 122.62, 121.67, 120.97, 117.65, 115.63, 113.00, 111.85, 111.82, 49.88, 48.14, 42.66, 26.18, 24.95, 15.58, 7.24; FT–IR (KBr, cm−1):3518, 3237, 2966, 1762, 1682, 1642, 1607, 1441, ESI-MS (m/z), ion Calculated: 418.20 amu; Found: 419.22 amu [IIED + H+].

3.4. Sample Preparation for Spectroscopic Studies

Spectroscopy graded solvents were used for preparing all the stock and working solutions. Spectroscopy graded DMSO were used for the stock solution preparation and mixture of DMSO with HEPES buffer (pH = 7.4, 25 mM) was used in the spectral studies. The absorption and emission experiments of probe IIED was carried out in HEPES buffer: DMSO (3:7, v/v) mixture. All experiments were carried out at room temperature of 298 K. The stock solutions of the cations were prepared by the perchlorate salts. In UV–Visible and fluorometric experiment, a stock solution of 10−3 (M) and 10−4 (M) IIED was filled in a quartz optical cell of 1.0 cm optical path length to achieve a final concentration of the solution of IIED (10−5 (M) and 10−6 (M)) in 2000 µL. After 1 min, each spectral data was recorded with the addition of cation solution by using micropipette.

4. Conclusions

In this paper, we have designed, and synthesized an isoindole-imidazole skeleton based chemosensor that is highly selective for Zn2+ in 3:7 HEPES buffer: DMSO medium. The probe itself shows a weak fluorescence intensity (quantum yield, 0.036), and the addition of zinc ion enhanced the fluorescence intensity 19-fold (quantum yield, 0.69). The competitive study confirms high selectivity for Zn2+ against other common cations. The detection limit for zinc ion was 0.073 μm, which is significantly lower than the WHO guideline (76.5 mM). Addition of EDTA to the probe: Zn2+ ensemble quenched the fluorescence, indicating the reversibility of Zn2+ binding. Zn2+ titrations and Job’s plot analysis indicated the formation of a 1:1 probe: Zn2+ complex with a binding constant of 3 × 104 M−1. The sensing application has been expanded successfully to gel form also. DFT/TD-DFT calculations also pointed towards the sensing mechanism along with the probable binding mode. If low concentrations of Zn2+ are required, our recommendation is to exploit the emission ability of the IIED probe.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27092859/s1, 1H, 13C NMR, ESI-MS, FT-IR spectrum of probe IIED, spectrophotometric study, frontier molecular orbitals.

Author Contributions

Conceptualization, Y.S., S.S. and R.B.; software, J.P.C.-C. and J.C.; validation, J.P.C.-C. and S.G.; formal analysis, S.S. and R.B.; investigation, S.S., R.B. and Y.S.; writing—original draft preparation, S.S.; writing—review and editing, R.B., Y.S., J.C., J.P.C.-C. and S.G.; supervision, S.G. and J.P.C.-C.; funding acquisition, R.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Shailabala Biswas Research Foundation, University of Calcutta, India (ET/T/385-dt-27/02/2018) and University Grants Commission, India (Sanction No. F1-17.1/2017-18/RGNF-2017-18-SC-WES-37660/(SA-III/Website)).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Acknowledgments

S.S. is thankful to Shailabala Biswas Research Foundation, University of Calcutta, Kolkata, India. R.B. is thankful to University Grants Commission, India (Sanction No. F1-17.1/2017-18/RGNF-2017-18-SC-WES-37660/(SA-III/Website)) for the financial support. The CAS–V (UGC) of the Department of Chemistry, University of Calcutta, is gratefully acknowledged. DST-FIST and DST-PURSE are acknowledged for providing Single crystal XRD and ESI-MS facility at Department of Chemistry, University of Calcutta. J.C. acknowledges the MICINN Project PID2019-110091GB-I00 for financial support. This work has used the computational facilities provided by the PlataformaAndaluza de Bioinformática installed at the Universidad of Málaga and by the Centro de Cálculo Científico at Universidad Autónoma de Madrid (CCC-UAM).

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Jung, J.H.; Lee, J.H.; Shinkai, S. Functionalized magnetic nanoparticles as chemosensors and adsorbents for toxic metal ions in environmental and biological fields. Chem. Soc. Rev. 2011, 40, 4464–4474. [Google Scholar] [CrossRef] [PubMed]
  2. Kan, C.; Wu, L.; Shao, X.; Wang, X.; Zhang, Y.; Zhu, J.; Qiu, S. A new reversible fluorescent chemosensor based on rhodamine for rapid detection of Al(III) in natural environmental water samples and living organisms. Tetrahedron Lett. 2020, 61, 152407. [Google Scholar] [CrossRef]
  3. Roy, A.; Dey, S.; Roy, P. A ratiometricchemosensor for Al3+ based on naphthalene–quinoline conjugate with the resultant complex as secondary sensor for F: Interpretation of molecular logic gates. Sens. Actuators B Chem. 2016, 237, 628–642. [Google Scholar] [CrossRef]
  4. Li, C.; Qin, J.; Wang, B.; Bai, X.; Yang, Z. Fluorescence chemosensor properties of two coumarin–based compounds for environmentally and biologically important Al3+ ion. J. Photochem. Photobiol. A 2017, 332, 141–149. [Google Scholar] [CrossRef]
  5. Sharma, H.; Kaur, N.; Singh, A.; Kuwar, A.; Singh, N. Optical chemosensors for water sample analysis. J. Mater. Chem. C 2016, 4, 5154–5194. [Google Scholar] [CrossRef]
  6. Lee, S.A.; Lee, J.J.; Shin, J.W.; Min, K.S.; Kim, C. A colorimetric chemosensor for the sequential detection of copper(II) and cysteine. Dyes Pigm. 2015, 116, 131–138. [Google Scholar] [CrossRef]
  7. Kang, J.H.; Lee, S.Y.; Ahn, H.M.; Kim, C. A novel colorimetric chemosensor for the sequential detection of Ni2+ and CN in aqueous solution. Sens. Actuators B Chem. 2017, 242, 25–34. [Google Scholar] [CrossRef]
  8. Cao, D.; Liu, Z.; Verwilst, P.; Koo, S.; Jangjili, P.; Kim, J.S.; Lin, W. Coumarin-based small-molecule fluorescent chemosensors. Chem. Rev. 2019, 18, 10403–10519. [Google Scholar] [CrossRef]
  9. Roy, P. Recent advances in the development of fluorescent chemosensors for Al3+. Dalton Trans. 2021, 50, 7156–7165. [Google Scholar] [CrossRef]
  10. Roy, P. Fluorescent chemosensors based on 4-methyl-2,6-diformylphenol. Coord. Chem. Rev. 2021, 427, 213562. [Google Scholar] [CrossRef]
  11. Kim, M.J.; Kaur, K.; Singh, N.; Jang, D.O. Benzimidazole-based receptor for Zn2+ recognition in a biological system: A chemosensor operated by retarding the excited state proton transfer. Tetrahedron 2012, 68, 5429–5433. [Google Scholar] [CrossRef]
  12. Jiang, P.; Guo, Z. Fluorescent detection of zinc in biological systems: Recent development on the design of chemosensors and biosensors. Coord. Chem. Rev. 2004, 248, 205–229. [Google Scholar] [CrossRef]
  13. Cassandri, M.; Smirnov, A.; Novelli, F.; Pitolli, C.; Agostini, M.; Malewicz, M.; Melino, G.; Raschellà, G. Zinc-finger proteins in health and disease. Cell Death Discov. 2017, 3, 17071. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Patil, M.; Bothra, S.; Sahoo, S.K.; Rather, H.A.; Vasita, R.; Bendre, R.; Kuwar, A. Highly selective nicotinohydrazide based ‘turn-on’ chemosensor for the detection of bioactive zinc(II): Its biocompitability and bioimaging application in cancer cells. Sens. Actuators B Chem. 2018, 270, 200–206. [Google Scholar] [CrossRef]
  15. Ma, L.; Liu, G.; Pu, S.; Zheng, C.; Fan, C. Solvent-dependent selective fluorescence sensor for Zn2+ and Al3+ based on a new diarylethene with a salicylal schiff base group. Tetrahedron 2017, 73, 1691–1697. [Google Scholar] [CrossRef]
  16. Andreini, C.; Bertini, I. A bioinformatics view of zinc enzymes. J. Inorg. Biochem. 2012, 111, 150–156. [Google Scholar] [CrossRef]
  17. Kimura, E.; Aoki, S.; Kikuta, E.; Koike, T. A macrocyclic zinc(II) fluorophore as a detector of apoptosis. Proc. Natl. Acad. Sci. USA 2003, 100, 3731–3736. [Google Scholar] [CrossRef] [Green Version]
  18. Weiss, J.H.; Sensi, S.L.; Koh, J.Y. Zn2+: A novel ionic mediator of neural injury in brain disease. Trends Pharmacol. Sci. 2000, 21, 395–401. [Google Scholar] [CrossRef]
  19. Assaf, S.Y.; Chung, S.H. Release of endogenous Zn2+ from brain tissue during activity. Nature 1984, 308, 734–736. [Google Scholar] [CrossRef]
  20. Yamasaki, S.; Hasegawa, A.; Hojyo, S.; Ohashi, W.; Fukada, T.; Nishida, K.; Hirano, T. A novel role of the L-Type calcium channel α1D subunit as a gatekeeper for intracellular zinc signaling: Zinc wave. PLoS ONE 2012, 7, e39654. [Google Scholar] [CrossRef]
  21. Li, Y.V. Zinc and insulin in pancreatic beta-cells. Endocrine 2014, 45, 178–189. [Google Scholar] [CrossRef] [PubMed]
  22. Frederickson, C.J.; Koh, J.-Y.; Bush, A.I. The neurobiology of zinc in health and disease. Nat. Rev. Neurosci. 2005, 6, 449–462. [Google Scholar] [CrossRef] [PubMed]
  23. Lee, S.R.; Noh, S.J.; Pronto, J.R.; Jeong, Y.J.; Kim, H.K.; Song, I.S.; Xu, Z.; Kwon, H.Y.; Kang, S.C.; Sohn, E.; et al. The critical roles of zinc: Beyond impact on myocardial signaling. Korean J. Physiol. Pharmacol. 2015, 19, 389–399. [Google Scholar] [CrossRef] [PubMed]
  24. Keen, C.L.; Gershwin, M.E. Zinc deficiency and immune function. Annu. Rev. Nutr. 1990, 10, 415–431. [Google Scholar] [CrossRef] [PubMed]
  25. Choi, J.Y.; Kim, D.; Yoon, J. A highly selective “turn–on” fluorescent chemosensor based on hydroxy pyrene–hydrazone derivative for Zn2+. Dyes Pigm. 2013, 96, 176–179. [Google Scholar] [CrossRef]
  26. Prasad, A.S. Zinc in Human Health: Effect of Zinc on Immune Cells. Mol. Med. 2008, 14, 353–357. [Google Scholar] [CrossRef] [PubMed]
  27. Park, G.J.; Kim, H.; Lee, J.J.; Kim, Y.S.; Lee, S.Y.; Lee, S.; Noh, I.; Kim, C. A highly selective turn–on chemosensor capable of monitoring Zn2+ concentrations in living cells and aqueous solution. Sens. Actuators B Chem. 2015, 215, 568–576. [Google Scholar] [CrossRef]
  28. Frederickson, C.J. Neurobiology of zinc and zinc–containing neurons. Int. Rev. Neurobiol. 1989, 31, 145–238. [Google Scholar]
  29. Auld, D.S. Zinc coordination sphere in biochemical zinc sites. Zinc Biochem. Physiol. Homeost. 2001, 14, 85–127. [Google Scholar]
  30. Cox, E.H.; McLendon, G.L. Zinc–dependent protein folding. Curr. Opin. Chem. Biol. 2000, 4, 162–165. [Google Scholar] [CrossRef]
  31. Xie, X.; Smart, T.G. A physiological role for endogenous zinc in rat hippocampal synaptic neurotransmission. Nature 1991, 349, 521–524. [Google Scholar] [CrossRef] [PubMed]
  32. Mammadova-Bach, E.; Braun, A. Zinc Homeostasis in Platelet–Related Diseases. Int. J. Mol. Sci. 2019, 20, 5258. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Ahmed, N.; Zareen, W.; Zhang, D.; Yang, X.; Ye, Y. A DCM–based NIR sensor for selective and sensitive detection of Zn2+ in living cells. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2020, 243, 118758. [Google Scholar] [CrossRef] [PubMed]
  34. Reichman, S.M. The responses of plants to metal toxicity: A review focusing on copper, manganese and zinc. Aust. Miner. Energy Environ. Found. 2002, 14, 1–54. [Google Scholar]
  35. Baran, A. Assessment of Zea mays sensitivity to toxic content of zinc in soil. Pol. J. Environ. Stud. 2013, 22, 77–83. [Google Scholar]
  36. Baran, A. Assessment of zinc content and mobility in maize. Ecol. Chem. Eng. 2012, 19, 699–706. [Google Scholar]
  37. Baran, A.; Wieczorek, J.; Mazurek, R.; Urbański, K.; Klimkowicz-Pawlas, A. Potential ecological risk assessment and predicting zinc accumulation in soils. Environ. Geochem. Health 2018, 40, 435–450. [Google Scholar] [CrossRef] [Green Version]
  38. Sakamaki, M.; Aikawa, S.; Fukushima, Y. Colorimetric chemosensor for Zn2+ based on pyrogallol red and poly(diallyldimethylammonium chloride) in aqueous solution. Polym. Bull. 2018, 75, 1667–1680. [Google Scholar] [CrossRef]
  39. Xu, Z.; Yoon, J.; Spring, D.R. Fluorescent chemosensors for Zn2+. Chem. Soc. Rev. 2010, 39, 1996–2006. [Google Scholar] [CrossRef] [Green Version]
  40. Liu, Y.; Li, Y.; Feng, Q.; Li, N.; Li, K.; Hou, H.; Zhang, B. ‘Turn–on’ fluorescent chemosensors based on naphthaldehyde–2–pyridinehydrazone compounds for the detection of zinc ion in water at neutral pH. Luminescence 2018, 33, 29–33. [Google Scholar] [CrossRef]
  41. Chae, J.B.; Yun, D.; Kim, S.; Lee, H.; Kim, M.; Lim, M.H.; Kim, K.; Kim, C. Fluorescent determination of zinc by a quinoline–based chemosensor in aqueous media and zebrafish. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2019, 219, 74–82. [Google Scholar] [CrossRef] [PubMed]
  42. Sarkar, D.; Pramanik, A.; Jana, S.; Karmakar, P.; Mondal, T.K. Quinoline based reversible fluorescent ‘turn-on’ chemosensor for the selective detection of Zn2+: Application in living cell imaging and as INHIBIT logic gate. Sens. Actuators B Chem. 2015, 209, 138–146. [Google Scholar] [CrossRef]
  43. Li, H.; Gao, S.; Xi, Z. A colorimetric and “turn–on” fluorescent chemosensor for Zn(II) based on coumarin Schiff–base derivative. Inorg. Chem. Commun. 2009, 12, 300–303. [Google Scholar] [CrossRef]
  44. Wang, K.P.; Jin, Z.H.; Shang, H.S.; Lv, C.D.; Zhang, Q.; Chen, S.; Hu, Z.Q. A highly selective fluorescent chemosensor for Zn2+ based on the rhodamine derivative incorporating coumarin group. J. Fluoresc. 2017, 27, 629–633. [Google Scholar] [CrossRef]
  45. Gupta, V.K.; Mergu, N.; Singh, A.K. Fluorescent chemosensors for Zn2+ ions based on flavonol derivatives. Sens. Actuators B Chem. 2014, 202, 674–682. [Google Scholar] [CrossRef]
  46. Jin, Y.; Wang, S.; Zhang, Y.; Song, B. Highly selective fluorescent chemosensor based on benzothiazole for detection of Zn2+. Sens. Actuators B Chem. 2016, 225, 167–173. [Google Scholar] [CrossRef]
  47. Qian, H.; Tao, T.; Wang, Y.; Yin, G.; Huang, W. A fluorescent chemosensor for Zn2+ based on 3, 8-bis(4-methoxyphenyl)-1,10-phenanthroline. Inorg. Chem. Commun. 2015, 58, 99–102. [Google Scholar] [CrossRef]
  48. Choi, Y.W.; You, G.R.; Lee, J.J.; Kim, C. Turn-on fluorescent chemosensor for selective detection of Zn2+ in an aqueous solution: Experimental and theoretical studies. Inorg. Chem. Commun. 2016, 63, 35–38. [Google Scholar] [CrossRef]
  49. Cheng, D.; Liu, X.; Yang, H.; Zhang, T.; Han, A.; Zang, L. A Cu2+ selective probe based on phenanthro–imidazole derivative. Sensors 2017, 17, 35. [Google Scholar] [CrossRef] [Green Version]
  50. Liu, H.; Zhang, B.; Tan, C.; Liu, F.; Cao, J.; Tan, Y.; Jiang, Y. Simultaneous bioimaging recognition of Al3+ and Cu2+ in living–cell, and further detection of F and S2− by a simple fluorogenicbenzimidazole–based chemosensor. Talanta 2016, 161, 309–319. [Google Scholar] [CrossRef]
  51. Li, D.; Zhong, Z.; Zheng, G.; Tian, Z. A naphthalene benzimidazole–based chemosensor for the colorimetric and on–off fluorescent detection of fluoride ion. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2017, 185, 173–178. [Google Scholar] [CrossRef] [PubMed]
  52. Molina, P.; Tárraga, A.; Otón, F. Imidazole derivatives: A comprehensive survey of their recognition properties. Org. Biomol. Chem. 2012, 10, 1711–1724. [Google Scholar] [CrossRef] [PubMed]
  53. Siwach, A.; Verma, P.K. Synthesis and therapeutic potential of imidazole containing compounds. BMC Chem. 2021, 15, 12. [Google Scholar] [CrossRef] [PubMed]
  54. Baire, B.; Niu, D.W.; Willoughby, P.H.; Woods, B.P.; Hoye, T.R. Synthesis of complex benzenoids via the intermediate generation of o-benzynes through the hexadehydro-Diels-Alder reaction. Nat. Protoc. 2013, 8, 501–509. [Google Scholar] [CrossRef] [Green Version]
  55. Crossley, S.W.M.; Shenvi, R.A. A longitudinal study of alkaloid synthesis reveals functional group interconversions as bad actors. Chem. Rev. 2015, 115, 9465–9531. [Google Scholar] [CrossRef]
  56. Ziegler, S.; Pries, V.; Hedberg, C.; Waldmann, H. Target identification for small bioactive molecules: Finding the needle in the haystack. Angew. Chem. Int. Ed. 2013, 52, 2744–2792. [Google Scholar] [CrossRef]
  57. Sahu, S.; Sikdar, Y.; Bag, R.; Maiti, D.K.; Cerón-Carrasco, J.P.; Goswami, S. Visual detection of fluoride ion based on ICT mechanism. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2019, 213, 354–360. [Google Scholar] [CrossRef]
  58. Sahu, S.; Sikdar, Y.; Bag, R.; Maiti, D.K.; Cerón-Carrasco, J.P.; Goswami, S. A novel quinoxaline–rhodamine conjugate for a simple and efficient detection of hydrogen sulphate ion. Compounds 2021, 1, 29–40. [Google Scholar] [CrossRef]
  59. Bag, R.; Sikdar, Y.; Sahu, S.; Saha, P.; Bag, J.; Pal, K.; Goswami, S. A quinoxaline–diaminomaleonitrile conjugate system for colorimetric detection of Cu2+ in 100% aqueous medium: Observation of aldehyde to acid transformation. Dalton Trans. 2019, 48, 5656–5664. [Google Scholar] [CrossRef]
  60. Bag, R.; Sikdar, Y.; Sahu, S.; Maiti, D.K.; Frontera, A.; Bauzá, A.; Drew, M.G.B.; Goswami, S. A versatile quinoxaline derivative serves as a colorimetric sensor for strongly acidic pH. Dalton Trans. 2018, 47, 17077–17085. [Google Scholar] [CrossRef]
  61. Mondal, S.; Mandal, S.M.; Ojha, D.; Chattopadhyay, D.; Sinha, C. Water soluble sulfaguanidine based Schiff base as a “Turn–on” fluorescent probe for intracellular recognition of Zn2+ in living cells and exploration for biological activities. Polyhedron 2019, 172, 28–38. [Google Scholar] [CrossRef]
  62. Wang, X.; Ding, G.; Duan, Y.; Wang, M.; Zhu, G.; Li, X.; Zhang, Y.; Qin, X. Novel ‘naked–eye’ bis–Schiff base fluorescent chemosensors for sensitive detection of Zn2+ and bio–imaging in living cells. Tetrahedron 2020, 76, 131108. [Google Scholar] [CrossRef]
  63. Li, Z.-Y.; Su, H.-K.; Zhou, K.; Yang, B.-Z.; Xiao, T.; Sun, X.-Q.; Jiang, J.; Wang, L. Oxo–spirocyclic structure bridged ditopic Schiff base: A turn–on fluorescent probe for selective recognition of Zn(II) and its application in biosensing. Dyes Pigm. 2018, 149, 921–926. [Google Scholar] [CrossRef]
  64. Jiao, S.-Y.; Kong, L.-M.; Liu, G.-Q.; Jia, X.; Tian, J.; Liu, Y.-G.; Zhang, L.-X.; Zhang, W.-X.; Li, Y.-H.; Huang, Z. A simple and an easy–to–synthesize turn–on fluorescent probe for rapid detection of Zn2+ and its application in bioimaging. Tetrahedron Lett. 2020, 61, 152507. [Google Scholar] [CrossRef]
  65. Das, S.; Fröhlich, R.; Pramanik, A. Synthesis and fluorescent properties of a new class of heterocycles of isoindole fused imidazoles with phenolic subunits. Org. Lett. 2006, 8, 4263–4266. [Google Scholar] [CrossRef]
  66. Chang, I.J.; Hwang, K.S.; Chang, S.-K. Selective Hg2+ signaling via dithiane to aldehyde conversion of an ESIPT fluorophore. Dyes Pigm. 2017, 137, 69–74. [Google Scholar] [CrossRef]
  67. Wang, B.D.; Hai, J.; Liu, Z.C.; Wang, Q.; Yang, Z.Y.; Sun, S.H. Selective detection of Iron(III) by rhodamine–modified Fe3O4 nanoparticles. Angew. Chem. Int. Ed. 2010, 49, 4576–4579. [Google Scholar] [CrossRef]
  68. Williams, A.T.R.; Winfield, S.A.; Miller, J.N. Relative fluorescence quantum yields using a computer–controlled luminescence spectrometer. Analyst 1983, 78, 1067–1071. [Google Scholar] [CrossRef]
  69. Thomsen, V.; Schatzlein, D.; Mercuro, D. Limits of detection in spectroscopy. Spectroscopy 2003, 18, 112–114. [Google Scholar]
  70. Ray, D.; Pramanik, A.; Guchhait, N. Lactim–lactam tautomerism through four member hydrogen bonded network in isoindole fused imidazole system: A combined spectroscopic and theoretical approach to photophysical properties. J. Photochem. Photobiol. A 2014, 274, 33–42. [Google Scholar] [CrossRef]
  71. Venkatesan, V.; Kumar, R.S.; Ashok Kumar, S.K.; Sahoo, S.K. Highly selective turn-on fluorogenic chemosensor for Zn2+ based on chelation enhanced fluorescence. Inorg. Chem. Comm. 2019, 102, 171–179. [Google Scholar] [CrossRef]
  72. Bhowmik, M.; Kumari, P.; Sarkar, G. Effect of xanthan gum and guar gum on in situ gelling ophthalmic drug delivery system based on poloxamer-407. Int. J. Biol. Macromol. 2013, 62, 117–123. [Google Scholar] [CrossRef] [PubMed]
  73. Pereira, G.G.; Kechinski, C.P. Formulation and characterization of poloxamer 407: Thermoreversible gel containing polymeric microparticles and hyaluronic acid. Quim. Nova 2013, 36, 1121–1125. [Google Scholar] [CrossRef] [Green Version]
  74. Adamo, C. Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys. 1999, 110, 6158–6170. [Google Scholar] [CrossRef]
  75. Zarabadi-Poor, P.; Barroso-Flores, J. Theoretical Assessment of the Selective Fluorescence Quenching of 1-Amino-8-naphthol-3, 6-disulfonic Acid (H-Acid) Complexes with Zn2+, Cd2+, and Hg2+: A DFT and TD-DFT Study. J. Phys. Chem. A 2014, 118, 12178–12183. [Google Scholar] [CrossRef] [PubMed]
  76. Rossini, E.; Bochevarov, A.D.; Knapp, E.W. Empirical conversion of pKa values between different solvents and interpretation of the parameters: Application to water, acetonitrile, dimethyl sulfoxide, and methanol. ACS Omega 2018, 3, 1653–1662. [Google Scholar] [CrossRef] [Green Version]
  77. Becke, A.D. A new mixing of Hartree-Fock and local density-functional theories. J. Chem. Phys. 1993, 98, 1372–1377. [Google Scholar] [CrossRef]
  78. Schrödinger Release 2021-3: Epik; Schrödinger, LLC: New York, NY, USA, 2021.
  79. Greenwood, J.R.; Calkins, D.; Sullivan, A.P.; Shelley, J.C. Towards the comprehensive, rapid, and accurate prediction of the favorable tautomeric states of drug-like molecules in aqueous solution. J. Comput. Aided Mol. Des. 2010, 24, 591–604. [Google Scholar] [CrossRef]
  80. Shelley, J.C.; Cholleti, A.; Frye, L.; Greenwood, J.R.; Timlin, M.R.; Uchimaya, M. Epik: A software program for pKa prediction and protonation state generation for drug-like molecules. J. Comp. Aided Mol. Des. 2007, 21, 681–691. [Google Scholar] [CrossRef]
  81. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.J.; et al. Gaussian 16, Revision, C.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
Scheme 1. Synthetic procedure of IIED.
Scheme 1. Synthetic procedure of IIED.
Molecules 27 02859 sch001
Figure 1. The changes under UV light (above)and day light (below)after addition of various cations to IIED.
Figure 1. The changes under UV light (above)and day light (below)after addition of various cations to IIED.
Molecules 27 02859 g001
Figure 2. Absorbance titration spectra of IIED (10−5 M) in the presence of various concentrations of Zn2+ (up to 2.5 equiv.). Upward arrow indicates gradual increase of absorption at 467 nm with increasing concentration of Zn2+. [Inset: Absorbance of IIED vs. [Zn2+] plot at 467 nm].
Figure 2. Absorbance titration spectra of IIED (10−5 M) in the presence of various concentrations of Zn2+ (up to 2.5 equiv.). Upward arrow indicates gradual increase of absorption at 467 nm with increasing concentration of Zn2+. [Inset: Absorbance of IIED vs. [Zn2+] plot at 467 nm].
Molecules 27 02859 g002
Figure 3. Emission titration spectra of IIED (10−6 M) in presence of various concentrations of Zn2+ (up to 25 equiv.). Upward arrow indicates gradual increase of emission at 558 nm with increasing concentration of Zn2+ ion (λex = 480 nm). [Inset: Emission of IIED vs. [Zn2+] plot at 558 nm].
Figure 3. Emission titration spectra of IIED (10−6 M) in presence of various concentrations of Zn2+ (up to 25 equiv.). Upward arrow indicates gradual increase of emission at 558 nm with increasing concentration of Zn2+ ion (λex = 480 nm). [Inset: Emission of IIED vs. [Zn2+] plot at 558 nm].
Molecules 27 02859 g003
Figure 4. Time-resolved fluorescence decay of IIED in the absence and presence of Zn2+ in 3:7 HEPES buffer:DMSO solvent at 480 nm.
Figure 4. Time-resolved fluorescence decay of IIED in the absence and presence of Zn2+ in 3:7 HEPES buffer:DMSO solvent at 480 nm.
Molecules 27 02859 g004
Figure 5. Left panel: parent IIED probe. Right panel: Complexes with Zn2+ as L(N) and L(NH) forms, displayed on the top and bottom, respectively.
Figure 5. Left panel: parent IIED probe. Right panel: Complexes with Zn2+ as L(N) and L(NH) forms, displayed on the top and bottom, respectively.
Molecules 27 02859 g005
Figure 6. Spectra simulated for the ligand in lactam form, both neutral and N-deprotonated. Stick transitions computed with TD—PBE0 were convoluted with Gaussian functions with HWHM = 0.25 eV. The spectra simulated for the N—deprotonated form was scaled by 0.5 to match the experimental spectra (shown as dashed lines).
Figure 6. Spectra simulated for the ligand in lactam form, both neutral and N-deprotonated. Stick transitions computed with TD—PBE0 were convoluted with Gaussian functions with HWHM = 0.25 eV. The spectra simulated for the N—deprotonated form was scaled by 0.5 to match the experimental spectra (shown as dashed lines).
Molecules 27 02859 g006
Figure 7. Optimized structures of Zn-Ligand. The amide site is considered protonated (LNH, left (a)) or deprotonated (LN-, right (b)).
Figure 7. Optimized structures of Zn-Ligand. The amide site is considered protonated (LNH, left (a)) or deprotonated (LN-, right (b)).
Molecules 27 02859 g007
Figure 8. Experimental and computed absorption spectra of the two possible compounds upon Zn2+ complexation, e.g., L(N) and L (NH), displayed with blue and red lines, respectively.
Figure 8. Experimental and computed absorption spectra of the two possible compounds upon Zn2+ complexation, e.g., L(N) and L (NH), displayed with blue and red lines, respectively.
Molecules 27 02859 g008
Figure 9. Visual change of IIED in the presence of Zn2+ in gel form under UV light.
Figure 9. Visual change of IIED in the presence of Zn2+ in gel form under UV light.
Molecules 27 02859 g009
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sahu, S.; Sikdar, Y.; Bag, R.; Cerezo, J.; Cerón-Carrasco, J.P.; Goswami, S. Turn on Fluorescence Sensing of Zn2+ Based on Fused Isoindole-Imidazole Scaffold. Molecules 2022, 27, 2859. https://doi.org/10.3390/molecules27092859

AMA Style

Sahu S, Sikdar Y, Bag R, Cerezo J, Cerón-Carrasco JP, Goswami S. Turn on Fluorescence Sensing of Zn2+ Based on Fused Isoindole-Imidazole Scaffold. Molecules. 2022; 27(9):2859. https://doi.org/10.3390/molecules27092859

Chicago/Turabian Style

Sahu, Sutapa, Yeasin Sikdar, Riya Bag, Javier Cerezo, José P. Cerón-Carrasco, and Sanchita Goswami. 2022. "Turn on Fluorescence Sensing of Zn2+ Based on Fused Isoindole-Imidazole Scaffold" Molecules 27, no. 9: 2859. https://doi.org/10.3390/molecules27092859

Article Metrics

Back to TopTop