DOI:
10.1039/C5RA15353K
(Paper)
RSC Adv., 2015,
5, 72659-72669
A new pyridoxal based fluorescence chemo-sensor for detection of Zn(II) and its application in bio imaging†
Received
1st August 2015
, Accepted 18th August 2015
First published on 18th August 2015
Abstract
This paper describes the activity of a Schiff base ligand, derived from pyridoxal, as a promising fluorescence probe for biologically important Zn(II) ion sensing. A physiologically compatible pyridoxal based chemosensor PydDmen was synthesized and evaluated for its fluorescent response towards metal ions. Chemosensor PydDmen exhibits a selective turn-on type response in the presence of Zn2+ in ethanol–water mixture. The addition of EDTA quenches the fluorescence of receptor PydDmen-Zn2+, making the chemosensor PydDmen reversible. The response is specific for Zn(II) ions, and remains almost unaffected by the presence of alkali and alkaline earth metals but is suppressed to varying degrees by transition metal ions. The selectivity mechanism of PydDmen for Zn2+ is the combined effects of proton transfer between the prevailing tautomeric forms, C
N isomerization and CHEF. The DFT optimized structure of the complex is compatible with elemental analysis, mass spectrometry, FT-IR, electronic and NMR spectra. The experimental and theoretical support in terms of NMR spectroscopy and DFT are provided to establish the existence of Zn2+ induced transformation of PydDmen to a 3-pyridone tautomeric form.
Introduction
The zinc ion (Zn2+) is the second most abundant heavy metal ion in the human body and the cellular biochemistry of Zn2+ is diverse and far ranging.1 On the other hand, misregulation of Zn2+ is also implicated in human health disorders. It is believed that a lack of zinc ions can result in an increased risk of several diseases such as those effecting stature, mental retardation and digestive dysfunction because the majority of biological zinc ions are tightly sequestered by proteins.2 Additionally, the presence of excess “free zinc” in certain cells may be related to severe neurological disorders such as Alzheimer's and Parkinson's diseases.3a,b Therefore, it is necessary to get an insight into the vital roles of Zn2+ in biological processes, resulting in the great demand regarding the design and development of efficient systems that can selectively and sensitively detect Zn2+ in living systems. Several analytical methods have played a vital role in the detection of Zn2+ including UV-Vis-spectroscopy,3c potentiometry,3d flame atomic absorption spectrometry.3e However, its investigation can be facilitated by the use of fluorescent probes as fluorescence detection, in particular, is considered to be the most effective tool for sensing applications owing to the high sensitivity, easy visualization, short response time for detection and most importantly can be used for real time bio-imaging.4 Unfortunately, zinc ions are not intrinsically fluorescent, making direct quantitative detection a difficult task.5,6 A variety of fluorescent sensors for Zn2+ have been reported based on various fluorophores.7–9 Continuous effort has been dedicated improving the effectiveness of Zn2+ sensors.
We are mainly concerned with Schiff bases as probes because Schiff bases are suitable ligands for metal ions. Schiff bases are inherently non-/poorly fluorescent due to conventional modes of non-radiative decay pathways such as, isomerisation of C
N bond in the excited state and ESIPT involving phenolic proton.10,11 Therefore, it is reasonable to expect that if we limit to simple Schiff bases as prospective probes for metal cations, the probable signaling pathways involve restriction of C
N isomerisation, ESIPT and CHEF (Scheme 1). But, Schiff base related probes can also coordinate strongly other physiologically available metal ions. Therefore, new strategies should be developed to improve Zn2+ selectivity of the probe. In order to make analyte binding more specific and favourable, it would be desirable if the binding pattern of the probe with the analyte is unique, i.e. the probe is transformed after binding to the analyte of choice. Such type of phenomena enhances the selectivity manifolds. Tautomerization is an efficient transformable factor in such cases.12 Spring and co-workers have beautifully and efficiently exploited the imidic acid and amide tautomeric forms for selective binding with metal ions.12 Till date, reports on tautomerism during analyte binding is relatively scarce in the literature (Chart 1†).13
 |
| Scheme 1 Probable signaling pathways of probe PydDmen. | |
In order to enhance the bio-compatibility in bioimaging, we have already started working with pyridoxal14 containing Schiff bases as chemosensors and reported a chemosensor that selectively detects Cu2+ ion.14,15 The pyridoxal and its derivatives exhibit a wide range biological properties and have been used as substrates in many biological transformations.16–21 In order to explore further this less ventured path, we herein constructed a novel turn-on sensor for Zn2+, which can exhibit emission at longer wavelength (483 nm). The 3-hydroxy pyridine moiety present in the probe permits 3-hydroxy pyridine–3-pyridone tautomeric equilibrium to exist.22 The spectroscopic behaviour of Schiff base and PydDmen [((2-(dimethylamino)ethylimino)methyl)-5-(hydroxymethyl)-2-methylpyridin-3-ol] shows that it is an excellent chemosensor for Zn2+ in ethanol–water and can be used for Zn2+ monitoring in living cells. The NMR spectroscopy and DFT study are utilized to establish the Zn2+ induced transformation of PydDmen to 3-pyridone tautomeric form in solution. To the best of our knowledge, this is the first report that establishes such type of tautomeric transformation as mechanistic rationale by means of 1H NMR, 13C NMR and 13C DEPT methods.
Results and discussion
Synthesis and FTIR spectral characterization of PydDmen
The chemosensor PydDmen has been synthesized by condensing pyridoxal hydrochloride with N,N-dimethylethylenediamine under refluxing conditions in ethanol medium (Scheme 2) and the structure of PydDmen was well characterized by 1H NMR and FTIR spectroscopy (Fig. S1†). The PydDmen-Zn2+ complex was characterized by recording FTIR and ESI-MS spectra (Fig. S2†).
 |
| Scheme 2 Preparation of chemosensor PydDmen. | |
FTIR spectra of PydDmen showed the characteristic band due to ν(C
N) at 1650 cm−1. In the Zn2+ complex, the ν(C
N) absorption appears at lower energy (ca. 1631 cm−1) indicating possible coordination to the metal center. The complexes also display broad band of medium intensity around 3400 cm−1 attributable to the –OH stretching vibration of the –CH2OH group of the pyridoxal part of the chemosensor PydDmen.23
UV-Vis spectroscopic investigation of PydDmen
In order to ascertain the complexation of Zn2+ by PydDmen, absorption titrations were carried out by adding varied concentrations of Zn(NO3)2·6H2O to a fixed concentration of PydDmen. Fig. 1 depicts spectrophotometric changes upon titrating a fixed concentration of PydDmen (5 × 10−5 M) with incremental additions of Zn(NO3)2·6H2O (55 × 10−5 M) in EtOH/H2O (4
:
1, v/v, 25 mM Tris buffer, pH 7.4). The absorption spectrum of PydDmen in same solvent displayed sharp absorption bands centered at 253 nm and 335 nm, which are assigned to the π–π* transitions. The absorption bands at 415 nm is attributed to the n–π* transitions of azomethine group.
 |
| Fig. 1 UV-Vis spectral changes of sensor PydDmen (c = 5 × 10−5 M) in EtOH/H2O (4 : 1, v/v, 25 mM Tris buffer, pH 7.4) solutions upon addition of Zn2+ ions (0–55 equivalent) (c = 0–55 × 10−5 M) in EtOH/H2O (4 : 1, v/v) at pH 7.4. | |
However, addition of Zn2+ induced dramatic modification both in the maxima and shape of the said bands of PydDmen. The band at 253 nm is decreased slightly, accompanied by a blue shift to 246 nm. The band maxima at 335 nm is gradually decreased and a new band appeared at 295 nm. Another broad and moderately intense band around 390 nm could be assigned to O− (phenolate) ↔ Zn2+ (LMCT or MLCT). These spectral changes, spanning the 240–390 cm−1 region, are indicative of Zn2+ coordination induced perturbation of the –(OH)C
C–C
N– portion of the ligand during the course of the reaction.24 Therefore, these absorption peaks were expected to correspond to coordination of PydDmen with Zn2+ generating the PydDmen-Zn2+ coordinated species. The accompanying isosbestic points at 268, 302, 356 and 435 nm clearly indicate that the transition between the free and the complexed species occurs and a stable complex resulted at a certain composition. The stoichiometry of the complex formed between PydDmen and Zn2+ is 1
:
1 based on Job's plot (Fig. S3†).
Binding behaviour analysed by fluorescence spectroscopy
The chemosensor PydDmen is poorly fluorescent in nature when excited at 411 nm in EtOH/H2O (4
:
1, v/v, 25 mM Tris buffer, pH 7.4), which may be attributed to the combined effect of C
N isomerisation and ESIPT as commonly encountered in Schiff bases.10 The fluorescence sensing behaviour of PydDmen towards Zn2+ was investigated in buffer solution at physiological pH in EtOH/H2O (4
:
1, v/v, 25 mM Tris buffer, pH 7.4) using a 5 × 10−6 M solution (Fig. 2). Zn(NO3)2·6H2O was chosen as the representative Zn2+ species in the following experiment. Upon addition of 14 equivalents of Zn(NO3)2·6H2O an enhancement in fluorescence spectrum was observed at 483 nm and maximum emissive wavelength shifts from 477 to 483 nm.
 |
| Fig. 2 Fluorescence emission changes of PydDmen (c = 5 × 10−6 M) upon addition of Zn2+ ions (c = 0–70 × 10−6 M) in EtOH/H2O (4 : 1, v/v) in Tris buffer at pH 7.4 (λex = 411 nm). | |
Under the same conditions, the selective sensing behaviour of PydDmen was validated using a variety of other metal ions in place of Zn2+, viz., Li+, Na+, K+, Sr2+, Ba2+, Ca2+, Cr3+, Mn2+, Fe2+, Co2+, Ni2+, Cu2+, Hg2+, Pb2+ and Al3+. But they do not show any significant change of fluorescence intensity of PydDmen, whereas Mg2+ and Cd2+ produce moderate enhancements (Fig. 3). From Fig. 3 it is clear that the Zn2+ ion gives rise to the largest fluorescence enhancement among these metal cations.
 |
| Fig. 3 Emission spectra of PydDmen (c = 5 × 10−6 M) in presence of Zn2+, Li+, Na+, K+, Ca2+, Sr2+, Al3+, Pb2+, Cr3+, Mn2+, Fe2+,Fe3+, Co2+, Ni2+, Cu2+, Cd2+, Hg2+ and Mg2+(c = 70 × 10−6 M each metal ion) in EtOH/H2O (4 : 1, v/v, 25 mM Tris buffer, pH 7.4) (λex = 411 nm). | |
The competitive studies of PydDmen towards Zn2+ over other metal ions are carried out by adding 14 equivalents of Zn2+ to the solution of PydDmen (5 × 10−6 M) in the presence of 14 equivalents of other metal ions, viz., Li+, Na+, K+, Sr2+, Ba2+, Ca2+, Cr3+, Mn2+, Fe2+, Fe3+, Co2+, Ni2+, Cu2+, Hg2+, Pb2+, Al3+, Mg2+ and Cd2+. The competition experiments showed that the emission profile of the PydDmen-Zn2+ complex is more or less unaffected in the presence of other cations (Fig. S4†).
Titration of PydDmen with Zn2+ (concentration increased from 0–70 μM) revealed enhancement (up to 52 fold enhancement) of fluorescence intensity at 483 nm as a function of the added Zn2+ concentration (Fig. 2), suggesting a sensitive and selective recognition of Zn2+ by PydDmen. The chelation of Zn2+ by PydDmen occurs via Namine/imine and Oalkoxo giving rigidity to the binding core. Therefore, C
N isomerisation as well as ESIPT is inhibited, reducing the non-radiative decay processes and increasing the possibility of fluorescence emission. This phenomenon results in chelation enhanced fluorescence (CHEF).
The linear relationship of the fluorescence titration showed that PydDmen responded to Zn2+ in 1
:
1 stoichiometry as evident from the Job's plot from absorption studies (Fig. S3†). The association constant for Zn2+ was estimated to be 1.18 × 104 M−1 by the linear Benesi–Hildebrand equation F0/(F − F0) = F0/[ PydDmen] + F0/[PydDmen] × Ka × [Zn2+]. F is the change in the fluorescence intensity at 483 nm, Ka is the association constant, and [PydDmen] and [Zn2+] are the concentration of PydDmen and Zn2+ respectively. By plotting F0/(F − F0) against the reciprocal of the concentration of Zn2+, the association constant value Ka is obtained from the ratio intercept/slope with a good linear correlation coefficient (R2 = 0.998) (Fig. 4).
 |
| Fig. 4 Benesi–Hildebrand expression fitting of fluorescence titration curve of PydDmen (c = 5 × 10−6 M) upon addition of Zn2+ ions (c = 6, 10, 15, 20, 25, 30 and 35 × 10−6 M) in EtOH/H2O (4 : 1, v/v) in Tris buffer at pH 7.4 (λex = 411 nm). | |
Since many fluorescent probes are sensitive to pH, it is necessary to investigate the pH effect to find the optimal condition when the fluorescence measurements are to be carried out. From the pH dependence of fluorescence study (Fig. S5†), it was found that the fluorescence intensity of PydDmen at 483 nm in EtOH/H2O remains unaffected at pH 7.4 which makes it suitable for application under physiological conditions. These results indicate that PydDmen can be used as a selective fluorescent probe to recognize and distinguish Zn2+ in the presence of various metal ions at pH 7.4.
We have also performed a reversibility experiment (Fig. S6†) which proved that the binding of Zn2+ to PydDmen is reversible which is the key requirement of an ideal biologically relevant chemosensor because binding of guest molecule must occur reversibly. In the presence of Na2EDTA, a strong chelating ligand, due to its strong affinity towards Zn2+, decomposition of the PydDmen-Zn2+ entity takes place thereby reproducing non fluorescent PydDmen. As shown in Fig. S6† after the addition of Na2EDTA, the emission intensity of the original ligand was gradually lost. This phenomena certainly gives a tacit support towards the reversible binding of PydDmen with Zn2+.
To determine the detection limit, following equation was used. DL = K × Sb/S where K = 2 or 3 (we take 3 in this case), Sb is the standard deviation of the blank solution and S is the slope of the calibration curve.25 Here, the detection limit of PydDmen (Fig. S7†) as a chemosensor for Zn2+ was found to be 40.78 × 10−7 M which is sufficiently low for the detection of submillimolar concentrations of Zn2+ ions found in many chemical systems.26
The fluorescence quantum yields (Φf)27 of PydDmen and PydDmen-Zn2+ states were found to be 0.094 and 0.408 respectively (Table S1†). This substantial increase in the quantum yield of PydDmen in the presence of Zn2+ advocates its credibility as an efficient Zn2+ sensor.
We have also examined the anion independency of PydDmen by using Cl−, Br−, I−, NO3−, CH3COO−, ClO4− salts and the spectral output is represented in Fig. S8.†
Time resolved measurement
A picosecond time-resolved fluorescence technique has been used to examine the decay process of free sensor PydDmen and PydDmen-Zn2+ in EtOH/H2O (4
:
1, v/v, 25 mM Tris buffer, pH 7.4, 298 K). According to the equations τ−1 = kr + knr and kr = Φf/τ, the radiative decay rate constant kr and the total nonradiative decay rate constant knr of PydDmen and Zn2+-bound species were calculated. The decay curve of the fluorescence intensity of PydDmen and fitting data were shown in Fig. 5 and Table S1.†
 |
| Fig. 5 Time resolved fluorescence decay of sensor PydDmen (red) and PydDmen-Zn2+ (green) and prompt (black) (λex = 375 nm). | |
The decay curve and fitting data of PydDmen suggested that there were three main isomeric components of PydDmen to absorb light and emit fluorescence photons of lifetime at 0.821 ns, 3.375 ns and 9.334 ns. The average fluorescence lifetime (τ) of PydDmen was estimated as 3.79 ns. The radiative and nonradiative decay rate constants are calculated to be 2.48 × 107 and 2.39 × 108 s−1 respectively indicating that the nonradiative decay is the predominant process in the excited states.28 In the presence of 70 × 10−6 M Zn2+, the time-resolved fluorescence decay showed significant change, which indicated two components corresponding to PydDmen-Zn2+ at 9.078 ns and a small number of isomeric components of PydDmen at 1.479 ns. The lifetime is increased to 8.23 ns, which is longer than that of the free-PydDmen. The radiative and nonradiative decay rate constants changed to 4.96 × 107 and 7.19 × 107 s−1 respectively. This result suggested that both the radiative and nonradiative decay processes became comparative resulting in a strong fluorescence.
1H and 13C NMR titrations and mode of binding present in the 3-pyridone tautomeric form
In order to evaluate the binding mode of PydDmen with Zn2+, 1H and 13C NMR titrations, and 13C-DEPT NMR experiment were performed by gradual addition of Zn(CH3COO)2·2H2O to the DMSO-d6 solution of PydDmen. 1H NMR spectroscopy revealed significant differences as shown in Fig. 6 in the chemical shifts of PydDmen and PydDmen-Zn2+, which could be applied for establishing the structure of PydDmen-Zn2+. As presumed from the UV-Vis study (Scheme 3), the –(OH)C
C–C
N– region of PydDmen is severely perturbed after addition of Zn2+.
 |
| Fig. 6 1H NMR titration experiment of PydDmen with Zn2+. | |
 |
| Scheme 3 Zn2+ induced tautomerism in PydDmen (red color indicate the change in molecular fragment due to Zn2+ binding). | |
Therefore, it may be speculated that addition of Zn2+ induces a transformation to PydDmen. Again, it is well documented that 3-hydroxy pyridine-bearing moieties could undergo tautomerism producing the 3-pyridones.15 If this speculation is true, (Scheme 4), then, it is expected that Ha, Hb, Hd and He protons would be upfield shifted because of through bond propagation effect due to tautomerism. Given the greater distance from the perturbed zone, only small differences in the chemical shift values for He and Hb are expected. Due to conversion of Nimine to Namine, Hg would undergo moderate upfield shift. For the Hh protons, downfield shift is expected as the NMe2 lone pair is used up in the complexation process.
 |
| Scheme 4 Pathway of tautomerism in PydDmen. | |
Analysis of the 1H NMR spectra after titration revealed that the speculated output were in good agreement with the experiment (Fig. 6). Upon addition of 0.5 equivalent of Zn2+ to a solution of PydDmen, the signals for Ha, Hd, Hb, Hg, Hh and He significantly broadened indicating incomplete complexation between PydDmen and Zn2+. After the addition of one equivalent of Zn2+, the resonances of Ha, Hb, Hd, He, Hg protons were found to be shielded relative to resonances for PydDmen indicating enhancement of electron density in the associated regions (Table S2†). Deprotonation of –CH2OH group due to coordination to Zn2+, would accumulate negative charge on the oxygen atom resulting in an upfield shift for Hb. But, subsequent complexation process decreases electron density on Hb. Therefore, the extent of upfield shift is less for Hb. The same logic holds true for Hg protons also. The resonances of methylenic protons Hh, α to –NMe2 fragment, were deshielded and show downfield shift from 2.47 ppm to 2.53 ppm. To understand the fate of the C3–O1Hf, a 13C NMR experiment was executed (Fig. 7). A new peak appeared at δ 174.0, which was disappeared in the 13C-DEPT experiment (Fig. 4). It confirms that C3–O1Hf was transformed to corresponding ketone. The ligand after complexation with zinc transformed to corresponding 3-pyridone–Zn2+ complex i.e. PydDmen-Zn2+ (Scheme 3). The imine bond was subsequently transformed to an amino exocyclic double bond (C5), which appeared at δ 166.1. There were no further changes when more than one equivalent of Zn2+ were added, which was indicative of the 1
:
1 binding ratio between the sensor and Zn2+. These features are in accordance with the hypothesis that Zn2+-induced tautomerisation of the PydDmen occurs and coordinating environment of Zn2+ is composed of two Namine and one Oalkoxo of the PydDmen. Moreover, the reaction mechanism was also confirmed by mass spectral analysis, and a peak at m/z 386.08 is assignable to the mass of [PydDmen − Me + Zn2+ + H2O + CH3COO−].
 |
| Fig. 7 (a) 13C NMR of PydDmen and PydDmen-Zn2+ in DMSO-d6. (b) 13C DEPT spectral output after adding 1 equivalent of Zn2+ in PydDmen in DMSO-d6. | |
To further reinforce the Schiff base transformation phenomena and mode of complexation between PydDmen and Zn2+, DFT calculations were carried out. Since attempts to isolate single crystals of PydDmen-Zn2+ suitable for X-ray diffraction analysis were unsuccessful, the optimized structure of the complex was computed by theoretical methods. The ground state structures of PydDmen and PydDmen-Zn2+ were optimized by density functional theory (DFT) as implemented in the Gaussian 09 (B3LYP/6-311G(d,p)) software package.29 Full geometry optimizations were carried out at the UB3LYP level for PydDmen and PydDmen-Zn2+, which are shown in Fig. 8.
 |
| Fig. 8 B3LYP optimized structure of (a) PydDmen and (b) PydDmen-Zn2+. | |
In the optimized structure of PydDmen-Zn2+, it is clear that PydDmen is present in its pyridone tautomeric form and Zn2+ is coordinated by means of two Namine atoms and one Oalkoxo atom. One nitrate anion and two water molecules complete the six coordination geometry around Zn2+. The optimized bond parameters are given in Table S3.† Energy of different molecular orbitals of PydDmen-Zn2+ were calculated and given in Table S4.† Contour plots of some selected molecular orbitals of PydDmen-Zn2+ are given in Fig. S9.† The spatial distributions of Highest Occupied Molecular Orbital (HOMO) and the Lowest Unoccupied Molecular Orbital (LUMO) of PydDmen and PydDmen-Zn2+ are presented in Fig. 9. The energy gaps between HOMO and LUMO in PydDmen and PydDmen-Zn2+ were 3.80 eV and 3.28 eV respectively. The UV-Vis absorption spectra of PydDmen-Zn2+ were calculated with electronic ground and excited states through time dependent density functional theory calculations (TDDFT) using conductor-like polarizable continuum model (CPCM) in ethanol. The calculated singlet–singlet vertical electronic transitions are summarized in Table S5.† The calculated electronic transitions are very close to the experimental electronic bands. All the transitions in PydDmen-Zn2+ have intra-ligand charge transfer (ILCT) origin.
 |
| Fig. 9 HOMO and LUMO distribution of PydDmen and PydDmen-Zn2+. | |
Biological implications of PydDmen
To demonstrate the potential application of PydDmen, the intracellular Zn2+ imaging behaviour of PydDmen was studied on A549, human lung cancer cell lines by fluorescence microscopy. After incubation with PydDmen (10 μM) at 37 °C for 10 min, the cells displayed no intracellular fluorescence (Fig. 10(B)). However, cells displayed light fluorescence with the addition of low concentration of zinc ions (1 μM) (Fig. 10(C)) and exhibited gradually intensive fluorescence when exogenous Zn2+ was introduced into the cell via incubation with a zinc nitrate salt solution (Fig. 10(D) and (F)). The intensive fluorescence behaviour was, however, strongly suppressed when TPEN (100 μM) was added to the medium. Since TPEN confers having a strong scavenging action on Zn2+ ions, the sensors were competitively inhibited to bind with Zn2+ ions, as a result, the intensive fluorescence disappeared (Fig. 10(H)). This presents the confirmatory evidence of the sensor having the selectivity to sense Zn2+ ions. The fluorescence responses of PydDmen with various concentrations of added Zn2+ proves that such fluorescence intensity can be used as indelible signature of selective sensor response clearly evident from the cellular imaging. Hence, these results indicate that PydDmen is an efficient candidate for monitoring changes in the intracellular Zn2+ concentration under biological conditions.
 |
| Fig. 10 (A) Phase contrast, (B) fluorescence image of A549 cells incubated with 10 μM PydDmen for 10 min at 37 °C. PydDmen (10 μM) incubated cells were washed with PBS and were exposed to the presence of sequentially increased concentrations of added extracellular Zn2+ ion as (C) 1 μM, (D) 10 μM, (E) 20 μM and (F) 50 μM. (G) Represent the merge image of phase contrast and fluorescence image. (H) Represent disappearance of fluorescence intensity in the A549 cells treated with the PydDmen and Zn2+ ion after further addition of 100 μM TPEN. For all imaging, the samples were excited at 410 nm. | |
The cytotoxicity study (MTT assay) in human lung cancer cells treated with various concentrations of PydDmen for up to 12 h as shown in Fig. 11 showed that PydDmen concentrations up to 10 μM did not show significant cytotoxic effects on human lung cancer cells for at least up to 12 h of its treatment. The study suggests that PydDmen can be readily used as an efficient, selective and sensitive tool for bioimaging at the indicated doses and incubation time without cytotoxic effects. Thus the intensity based Zn-sensors can offer promising potential to probe physiological and biochemical consequences of metal dynamics with wide metabolic spectrum in cellular environment with appreciable fidelity.
 |
| Fig. 11 % Cell viability of A549 cells treated with different concentrations (1–100 μM) of PydDmen for 12 h determined by MTT assay. Results are expressed as mean ± S.D of three independent experiments. | |
Conclusion
In conclusion, we have synthesized and characterized a new pyridoxal containing Schiff base turn-on Zn2+ probe, PydDmen. The poorly fluorescent probe responds giving a strong selective, fast and specific fluorescence signal in presence of Zn2+ ions, i.e. there is a zinc triggered fluorescence switching. The C
N isomerisation and ESIPT are inhibited upon binding with Zn2+ ions, which causes CHEF effect, inducing an enhancement in the fluorescence intensity of the chemosensor. The complex formation, stoichiometry, and binding mode have been thoroughly examined by UV-Vis, ESI-MS, and NMR studies, which show formation of an 1
:
1 PydDmen-Zn2+ complex. The chelating agent EDTA can switch off the fluorescence signal by coordinating with the zinc ion, releasing the chemosensor to the solution. 1H, 13C NMR and DEPT analysis indicates Zn2+ induced transformation of the chemosensor to 3-pyridone tautomeric form. The DFT/TDDFT calculation was carried out to demonstrate the electronic properties of the chemosensor and PydDmen-Zn2+ and it supports the prevailing 3-pyridone tautomeric form. Furthermore, we have demonstrated that the probe is applicable for Zn2+ imaging in the living cells. We believe that the development of a new turn-on Zn2+ probe, its outstanding fluorescence enhancement, spectroscopic and DFT studies, and cell imaging will find considerable application in the chemical science and its allied branches.
Experimental section
General information and materials
All reagents were purchased from Sigma-Aldrich and used as received. Solvents were spectroscopic grade and used without purification. Elemental analyses (carbon, hydrogen and nitrogen) were carried out with a Perkin-Elmer CHN analyzer 2400. The 1H and 13C NMR spectra were measured on Bruker-300 MHz spectrometer. IR spectra were recorded in the region 400–4000 cm−1 on a Bruker-Optics Alpha-T spectrophotometer with samples as KBr disks. Electronic spectra were obtained by using a Hitachi U-3501 spectrophotometer. Luminescence property was measured using LS-55 PerkinElmer fluorescence spectrophotometer at room temperature (298 K) by 1 cm path length quartz cell. Fluorescence lifetimes were obtained by the method of Time Correlated Single-Photon Counting (TCSPC) on FluoroCube-01-NL spectrometer (Horiba Jobin Yvon) using a nanoLED as light source (340 nm) and the signals were collected at the magic angle of 54.7° to eliminate any considerable contribution from fluorescence anisotropy decay. The typical time resolution of our experimental set up is 800 ps. The decays were deconvoluted using DAS-6 decay analysis software. The acceptability of the fits was judged by χ2 criteria (fitting analysis having χ2 beyond the range 1.20 < χ2 < 1.00 has been neglected) and visual inspection of the residuals of the fitted function to the data. Mean (average) fluorescence lifetimes were calculated using the following equation:
in which ai is the pre-exponential factor corresponding to the ith decay time constant, τi.
Reagents for cell study
A549, human lung cancer cell lines were collected from National Center for Cell Science, Pune, India, and used throughout the experiments. Cells were grown in DMEM (Himedia) supplemented with 10% FBS (Himedia), and an antibiotic mixture (1%) containing PSN (Himedia) at 37 °C in a humidified incubator with 5% CO2 and cells were grown to 80–90% confluence, harvested with 0.025% trypsin (Himedia) and in phosphate-buffered saline (PBS), plated at the desired cell concentration and allowed to grow overnight before any treatment.
Imaging system
Fluorescence images of A549 cells were taken by a fluorescence microscope (Model: LEICA DM4000B, Germany) with an objective lens of 20× magnification.
Cell culture
Cells were rinsed with PBS and incubated with DMEM containing PydDmen making the final concentration up to 10 μM in DMEM [the stock solution (3 mmol) was prepared by dissolving PydDmen into ethanol] for 10 min at 37 °C. After incubation, bright field and fluorescence images of A549 cells were taken by a fluorescence microscope (Model: LEICA DM4000B, Germany) with an objective lens of 20× magnification. Similarly, fluorescence images of A549 cells (pre-incubated with 10 μM PydDmen) were taken with addition of different concentrations (1–50 μM) of zinc nitrate salt at 10 minutes interval. A merged image between phase contrast and fluorescent images at 50 μM salt concentration were taken and consequently fluorescence images were taken after further addition of TPEN (100 μM).
Cell cytotoxicity assay
In order to test the cytotoxicity of PydDmen, 3-(4, 5-dimethylthiazol-2-yl)-2,S-diphenyltetrazolium bromide (MTT) assay was performed in A549 cells according to standard procedure.30 Briefly, after treatment of overnight culture of A549 cells (103 cells in each well of 96-well plate) with PydDmen (1, 10, 20, 50 and 100 μM) for 12 h, 10 μL of a MTT solution (1 mg mL−1 in PBS) was added in each well and incubated at 37 °C continuously for 3 h. All media were removed from wells and 100 μL of acidic isopropyl alcohol was added into each well. The intracellular formazan crystals (blue-violet) formed were solubilized with 0.04 N acidic isopropyl alcohol and absorbance of the solution was measured at 595 nm wavelength with a microplate reader (Model: THERMO MULTI SCAN EX). The cell viability was expressed as the optical density ratio of the treatment to control. Values are mean ± standard deviation of three independent experiments. The cell cytotoxicity was calculated as % cell cytotoxicity = 100% − % cell viability.
Computational studies
All geometries for PydDmen and PydDmen-Zn2+ were optimized by density functional theory (DFT) calculations using Gaussian 09 (B3LYP/6-311G(d,p)) software package.31 The vibrational frequency calculations were performed to ensure that the optimized geometries represent the local minima and that there is only positive Eigen values. Vertical electronic excitations based on B3LYP optimized geometries were computed using the time-dependent density functional theory (TDDFT) formalism32 in ethanol using a conductor-like polarizable continuum model (CPCM).33
Fluorimetric analysis
Fluorescence quantum yields (Φ) were estimated by integrating the area under the fluorescence curves with the following equation:
Φsample = (ODstandard/ODsample) × (Asample/Astandard) × Φstandard |
where, A is the area under the fluorescence spectral curve and OD is the optical density of the compound at the excitation wavelength. The standard used for the measurement of the fluorescence quantum yield was quinine sulphate (Φ = 0.54 in water).
Synthesis of the chemosensor PydDmen
The chemosensor molecule PydDmen was synthesized by following procedure. Pyridoxal hydrochloride (0.406 g, 2 mmol) was dissolved in absolute ethanol (15 mL) in the presence of KOH (0.112 g, 2 mmol) with stirring. After 1 h of stirring, the separated white solid (KCl) was filtered and the obtained clear solution was added to a solution of N,N-dimethylethylenediamine (0.176 g, 2 mmol) in ethanol (15 mL) with stirring and the resulting reaction mixture was refluxed for 4 h. The completeness of the condensation reaction was checked by performing thin layer chromatography. The solution was evaporated by rotary evaporator and sticky mass obtained was washed by cold ether and dried under vacuum (yield: 0.355 g, 0.75%). 1H NMR (300 MHz, DMSO-d6): δ 8.797 (s, Ha), 7.762 (s, Hd), 4.548 (s, Hb), 3.672 (t, Hg), 2.471 (t, Hh), 2.271 (s, He), 2.10 (s, Hi). Anal. calc. for C12H19N3O2: C, 60.74; H, 8.07; N, 17.71. Found: C, 59.97; H, 7.79; N, 17.05%.
Synthesis of PydDmen-Zn2+
Pyridoxal hydrochloride (0.203 g, 1 mmol) was dissolved in absolute ethanol. To it, ethanolic solution of Zn(NO3)2·2H2O (0.297 g, 1 mmol) was added dropwise under stirring and the solution was stirred for 15 min. Then to this solution, N,N-dimethylethylenediamine (0.088 g, 1 mmol) was added slowly and the resulting yellowish-orange solution was stirred for 2 h. Then it was evaporated by rotary evaporator and resulting sticky mass was washed by cold ether and dried under vacuum (yield: 0.211 g, 0.53%). Anal. calc. for [Zn(PydDmen)(NO3)(H2O)2]: C, 36.06; H, 5.55; N, 14.02. Found: C, 35.54; H, 5.14; N, 13.87%.
Acknowledgements
Financial support from the University Grants Commission for senior research fellowship to S. Mandal [Sanction No. UGC/847/Jr. Fellow (Upgradation)] and from the DST to Y. Sikdar (Sanction no. SR/FT/CS-107/2011) is gratefully acknowledged. Dr Guru Prasad Maiti, Research Associate, is grateful to DST-PURSE Programme at University of Kalyani for financial support to complete a part of this work. Dr Sushil Kumar Mandal is grateful to DST-PURSE PROGRAMME for partial financial support of the biological work. The authors gratefully acknowledge Saikat Khamarui and Deborin Ghosh for helpful discussions.
References
-
(a) A. I. Bush, W. H. Pettingell, G. Multhaup, M. Paradis, J. P. Vonsattel, J. F. Gusella, K. Beyreuther, C. L. Masters and R. E. Tanzi, Science, 1994, 265, 1464–1465 CAS;
(b) C. J. Frederickson, J. Y. Koh and A. I. Bush, Nat. Rev. Neurosci., 2005, 6, 449–452 CrossRef CAS PubMed;
(c) D. D. Mott, M. Benveniste and R. J. Dingledine, J. Neurosci., 2008, 28, 1659–1671 CrossRef CAS PubMed;
(d) J. M. Berg and Y. Shi, Science, 1996, 271, 1081–1085 CAS.
- A. Krężel and W. Maret, J. Biol. Inorg. Chem., 2006, 11, 1049–1062 CrossRef PubMed.
-
(a) A. I. Bush, Trends Neurosci., 2003, 26, 207–214 CrossRef CAS;
(b) D. Noy, I. Solomonov, O. Sinkevich, T. Arad, K. Kjaer and I. Sagi, J. Am. Chem. Soc., 2008, 130, 1376–1383 CrossRef CAS PubMed;
(c) C. V. Banks and R. E. Bisque, Anal. Chem., 1957, 29, 522–526 CrossRef CAS;
(d) A. R. Fakhari, M. Shamsipur and K. H. Ghanbari, Anal. Chim. Acta, 2002, 460, 177–183 CrossRef CAS;
(e) Q. Li, X. H. Zhao, Q. Z. Lv and G. G. Liu, Sep. Purif. Technol., 2007, 55, 76–81 CrossRef CAS PubMed.
-
(a) R. Y. Tsien, Fluorescent and Photochemical Probes of Dynamic Biochemical Signals inside Living Cells, ed. A. W. Czarnik, American Chemical Society, Washington, DC, 1993, pp. 130–146 Search PubMed;
(b) Y. Xiang, A. J. Tong, P. Y. Jin and Y. Ju, Org. Lett., 2006, 8, 2863–2866 CrossRef CAS PubMed.
-
(a) K. Kikuchi, K. Komatsu and T. Nagano, Curr. Opin. Chem. Biol., 2004, 8, 182–191 CrossRef CAS PubMed;
(b) E. L. Que, D. W. Domaille and C. J. Chang, Chem. Rev., 2008, 108, 1517–1549 CrossRef CAS PubMed;
(c) N. C. Lim, H. C. Freake and C. Bruckner, Chem.–Eur. J., 2005, 11, 38–49 CrossRef PubMed;
(d) Z. C. Xu, J. Y. Yoon and D. R. Spring, Chem. Soc. Rev., 2010, 39, 1996–2006 RSC.
-
(a) Y. H. Lau, P. J. Rutledge, M. Watkinson and M. H. Todd, Chem. Soc. Rev., 2011, 40, 2848–2866 RSC;
(b) D. Maity and T. Govindaraju, Chem. Commun., 2010, 46, 4499–4501 RSC;
(c) K. Jobe, C. H. Brennan, M. Motevalli, S. M. Goldup and M. Wakinson, Chem. Commun., 2011, 47, 6036–6038 RSC;
(d) T. L. Mindt, H. Struthers, L. Brans, T. Anguelov, C. Schweinsberg, V. Maes, D. Tourwe and B. Schibli, J. Am. Chem. Soc., 2006, 128, 15096–15097 CrossRef CAS PubMed;
(e) C. Ornelas, J. R. Aranzaes, E. Cloutet, S. Alves and D. Astruc, Angew. Chem., Int. Ed., 2007, 46, 872–877 CrossRef CAS PubMed;
(f) K.-C. Chang, I.-H. Su, A. Senthilvelan and W.-S. Chung, Org. Lett., 2007, 9, 3363–3366 CrossRef CAS PubMed;
(g) B. Colasson, M. Save, P. Milko, J. Roithova, D. Schroder and O. Reinaud, Org. Lett., 2007, 9, 4987–4990 CrossRef CAS PubMed;
(h) S. Huang, R. J. Clark and L. Zhu, Org. Lett., 2007, 9, 4999–5002 CrossRef CAS PubMed;
(i) K.-C. Chang, I.-H. Su, G.-H. Lee and W.-S. Chung, Tetrahedron Lett., 2007, 48, 7274–7278 CrossRef CAS PubMed;
(j) R. M. Meudtner, M. Ostermeier, R. Goddard, C. Limberg and S. Hecht, Chem.–Eur. J., 2007, 13, 9834–9840 CrossRef CAS PubMed;
(k) S. Y. Park, J. H. Yoon, C. S. Hong, R. Souane, J. S. Kim, S. E. Matthews and J. Vicens, J. Org. Chem., 2008, 73, 8212–8218 CrossRef CAS PubMed;
(l) D. Schweinfurth, K. I. Hardcastle and U. H. F. Bunz, Chem. Commun., 2008, 2203–2205 RSC;
(m) M. Juricek, P. H. J. Kouwer, J. Rehak, J. Sly and A. E. Rowan, J. Org. Chem., 2009, 74, 21–25 CrossRef CAS PubMed;
(n) J. Camponovo, J. Ruiz, E. Cloutet and D. Astruc, Chem.–Eur. J., 2009, 15, 2990–3002 CrossRef CAS PubMed;
(o) R. K. Pathak, V. K. Hinge, M. Mondal and C. P. Rao, J. Org. Chem., 2011, 76, 10039–10049 CrossRef CAS PubMed.
-
(a) K. K. Upadhyay, A. Kumar, J. Zhao and R. K. Mishra, Talanta, 2010, 81, 714–721 CrossRef CAS PubMed;
(b) J. F. Zhang, S. Kim, J. H. Han, S. J. Lee and J. S. Kim, Org. Lett., 2011, 13, 5294–5297 CrossRef CAS PubMed;
(c) Y. Xu, J. Meng, L. X. Meng, Y. Dong, Y. X. Cheng and C. J. Zhu, Chem.–Eur. J., 2010, 16, 12898–12903 CrossRef CAS PubMed;
(d) Q. H. You, P. S. Chan, W. H. Chan, N. K. Mak and R. N. S. Wong, RSC Adv., 2012, 2, 11078–11083 RSC;
(e) H. Y. Lin, P. Y. Cheng, C. F. Wan and A. T. Wu, Analyst, 2012, 137, 4415–4417 RSC.
-
(a) S. Comby, S. A. Tuck, L. K. Truman, O. Kotova and T. Gunnlaugsson, Inorg. Chem., 2012, 51, 10158–10168 CrossRef CAS PubMed;
(b) J. Jia, Q. C. Xu, R. C. Li, X. Tang, Y. F. He, M. Y. Zhang, Y. Zhang and G. W. Xing, Org. Biomol. Chem., 2012, 10, 6279–6286 RSC;
(c) X. Meng, S. Wang, Y. Li, M. Zhu and Q. Guo, Chem. Commun., 2012, 48, 4196–4198 RSC;
(d) T. Mukherjee, J. C. Pessoa, A. Kumar and A. R. Sarkar, Dalton Trans., 2012, 5260–5271 RSC;
(e) S. H. Mashraqui, R. Betkar, S. Ghorpade, S. Tripathi and S. Britto, Sens. Actuators, B, 2012, 174, 299–305 CrossRef CAS PubMed.
-
(a) G. Mandal, M. Darragh, Y. A. Wang and C. D. Heyes, Chem. Commun., 2013, 49, 624–626 RSC;
(b) G. Sivaraman, T. Anand and D. Chellappa, Analyst, 2012, 137, 5881–5884 RSC;
(c) L. J. Liang, S. J. Zhen, X. J. Zhao and C. Z. Huang, Analyst, 2012, 137, 5291–5294 RSC;
(d) P. G. Sutariya, N. R. Modi, A. Pandya, B. K. Joshi, K. V. Joshi and S. K. Menon, Analyst, 2012, 137, 5491–5494 RSC;
(e) Y. W. Choi, G. J. Park, Y. J. Na, H. Y. Jo, S. A. Lee, G. R. You and C. Kim, Sens. Actuators, B, 2014, 194, 343–352 CrossRef CAS PubMed;
(f) E. J. Song, H. Kim, I. H. Hwang, K. B. Kim, A. R. Kim, I. Noh and C. Kim, Sens. Actuators, B, 2014, 195, 36–43 CrossRef CAS PubMed;
(g) G. J. Park, H. Kim, J. J. Lee, Y. S. Kim, S. Y. Lee, S. Lee, I. Noh and C. Kim, Sens. Actuators, B, 2015, 215, 568–576 CrossRef CAS PubMed;
(h) J. J. Lee, S. A. Lee, H. Kim, L. T. Nguyen, I. Noh and C. Kim, RSC Adv., 2015, 5, 41905–41913 RSC;
(i) A. K. Bhanja, C. Patra, S. Mondal, D. Ojha, D. Chattopadhyay and C. Sinha, RSC Adv., 2015, 5, 48997–49005 RSC;
(j) C.-X. Yin, L.-J. Qu and F.-J. Huo, Chin. Chem. Lett., 2014, 25, 1230–1234 CrossRef CAS PubMed.
- J. Wu, W. Liu, X. Zhuang, F. Wang, P. Wang, S. Tao, X. Zhang, S. Wu and S. T. Lee, Org. Lett., 2007, 9, 33–36 CrossRef CAS PubMed.
-
(a) X. Zhang, L. Guo, F. Y. Wu and Y. B. Jiang, Org. Lett., 2003, 5, 2667–2670 CrossRef CAS PubMed;
(b) M. Royzen, A. Durandin, V. G. Young, N. E.Geacintov and J. W. Canary, J. Am. Chem. Soc., 2006, 128, 3854–3855 CrossRef CAS PubMed.
-
(a) Z. Xu, K.-H. Baek, H. N. Kim, J. Cui, X. Qian, D. R. Spring, I. Shin and J. Yoon, J. Am. Chem. Soc., 2010, 132, 601–610 CrossRef CAS PubMed;
(b) Z. Xu, X. Liu, J. Pan and D. R. Spring, Chem. Commun., 2012, 48, 4764–4766 RSC.
-
(a) V. Bhalla, Roopa and M. Kumar, Dalton Trans., 2013, 975–980 RSC;
(b) V. Bhalla, Roopa and M. Kumar, Org. Lett., 2012, 14, 2802–2805 CrossRef CAS PubMed;
(c) A. Satheshkumar, E. H. El-Mossalamy, R. Manivannan, C. Parthiban, L. M. Al-Harbi, S. Kosa and K. P. Elango, Spectrochim. Acta, Part A, 2014, 128, 798–805 CrossRef CAS PubMed;
(d) M. J. Kim, K. Kaur, N. Singh and D. O. Jang, Tetrahedron, 2012, 68, 5429–5433 CrossRef CAS PubMed;
(e) K. Kaur, V. K. Bhardwaj, N. Kaur and N. Singh, Inorg. Chem. Commun., 2012, 26, 31–36 CrossRef CAS PubMed;
(f) B. Babur, N. Seferoğlu and Z. Seferoğlu, Tetrahedron Lett., 2015, 56, 2149–2154 CrossRef CAS PubMed;
(g) A. D. Dubonosov, V. I. Minkin, V. A. Bren, E. N. Shepelenko, A. V. Tsukanov, A. G. Starikov and G. S. Borodkin, Tetrahedron, 2008, 64, 3160–3167 CrossRef CAS PubMed;
(h) A. Misra and M. Shahid, J. Phys. Chem. C, 2010, 114, 16726–16739 CrossRef CAS;
(i) S. Mukherjee, A. K. Paul and H. S. Evans, Sens.
Actuators, B, 2014, 202, 1190–1199 CAS;
(j) A. Samanta, S. Dalapati and N. Guchhait, J. Photochem. Photobiol., A, 2012, 232, 64–72 CrossRef CAS;
(k) Z.-H. Pan, J.-W. Zhou and G.-G. Luo, Phys. Chem. Chem. Phys., 2014, 16, 16290–16301 RSC;
(l) B. Liu, H. Wang, T. Wang, Y. Bao, F. Du, J. Tian, Q. Li and R. Bai, Chem. Commun., 2012, 48, 2867–2869 RSC.
- S. Mandal, R. Modak and S. Goswami, J. Mol. Struct., 2013, 1037, 352–360 CrossRef CAS PubMed.
- T. Mukherjee, J. C. Pessoa, A. Kumar and A. R. Sarkar, Dalton Trans., 2012, 5260–5271 RSC.
- A. C. Eliot and J. F. Kirsch, Annu. Rev. Biochem., 2004, 73, 383–415 CrossRef CAS PubMed.
- R. A. John, Biochim. Biophys. Acta, 1995, 1248, 81–96 CrossRef.
- M. D. Toney, Arch. Biochem. Biophys., 2005, 433, 279–287 CrossRef CAS PubMed.
- P. Christen and D. E. Metzler, Transaminases, Wiley, New York, 1985 Search PubMed.
- H. Hayashi, H. Mizuguchi, I. Miyahara, M. M. Islam, H. Ikushiro, Y. Nakajima, K. Hirotsu and H. Kagamiyama, Biochim. Biophys. Acta, 2003, 1647, 103–109 CrossRef CAS.
- W. Liu, P. E. Peterson, J. A. Langston, X. Jin, X. Zhou, A. J. Fisher and M. D. Toney, Biochemistry, 2005, 44, 2982–2992 CrossRef CAS PubMed.
- O. K. Abou-Ziad and O. I. K. Al-Shini, Phys. Chem. Chem. Phys., 2009, 11, 5377–5383 RSC.
- K. Nakamoto, Infrared Spectra of Inorganic Compounds, Wiley, New York, 1970 Search PubMed.
- L. Wang, W. Qin, X. Tang, W. Dou and W. Liu, J. Phys. Chem. A, 2011, 115, 1609–1616 CrossRef CAS PubMed.
-
(a) L. Long, D. Zhang, X. Li, J. Zhang, C. Zhang and L. Zhou, Anal. Chim. Acta, 2013, 775, 100–105 CrossRef CAS;
(b) M. Zhu, M. Yuan, X. Liu, J. Xu, J. Lv, C. Huang, H. Liu, Y. Li, S. Wang and D. Zhu, Org. Lett., 2008, 10, 1481–1484 CrossRef CAS PubMed;
(c) C. Kar, M. D. Adhikari, A. Ramesh and G. Das, Inorg. Chem., 2013, 52, 743–752 CrossRef CAS PubMed.
- G. L. Long and J. D. Winefordner, Anal. Chem., 1983, 55, 712A–724A CrossRef CAS.
- B. K. Paul, S. Kar and N. Guchhait, J. Photochem. Photobiol., A, 2011, 220, 153–163 CrossRef CAS PubMed.
- B. Valeur, Molecular Fluorescence: Principles and Applications, Wiley-VCH, Weinheim, Germany, 2002 Search PubMed.
-
(a) A. D. Becke, J. Chem. Phys., 1993, 98, 5648–5652 CrossRef CAS PubMed;
(b) C. Lee, W. Yang and R. G. Parr, Phys. Rev. B: Condens. Matter Mater. Phys., 1988, 37, 785–789 CrossRef CAS.
- T. Mossman, J. Immunol. Methods, 1983, 65, 55–63 CrossRef.
- M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery Jr, J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, Ö. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox, Gaussian 09, Revision D.01, Gaussian, Inc., Wallingford CT, 2009 Search PubMed.
-
(a) R. Bauernschmitt and R. Ahlrichs, Chem. Phys. Lett., 1996, 256, 454–464 CrossRef CAS;
(b) R. E. Stratmann, G. E. Scuseria and M. J. Frisch, J. Chem. Phys., 1998, 109, 8218–8224 CrossRef CAS PubMed;
(c) M. E. Casida, C. Jamorski, K. C. Casida and D. R. Salahub, J. Chem. Phys., 1998, 108, 4439–4449 CrossRef CAS PubMed.
-
(a) V. Barone and M. Cossi, J. Phys. Chem. A, 1998, 102, 1995–2001 CrossRef CAS;
(b) M. Cossi and V. Barone, J. Chem. Phys., 2001, 115, 4708–4717 CrossRef CAS PubMed;
(c) M. Cossi, N. Rega, G. Scalmani and V. Barone, J. Comput. Chem., 2003, 24, 669–681 CrossRef CAS PubMed.
Footnote |
† Electronic supplementary information (ESI) available: NMR & IR of the ligand, IR & ESI-MS of the complex, Job's plot, EDTA reversibility plot, detection limit plot, table of fluorescence lifetime of ligand and complex, tale of theoretical bond length and bond angle of complex, table of energy of selected molecular orbitals of complex, tables for calculated vertical electronic transitions of complex. See DOI: 10.1039/c5ra15353k |
|
This journal is © The Royal Society of Chemistry 2015 |
Click here to see how this site uses Cookies. View our privacy policy here.