Annamalai Subhasria,
Chinnadurai Anbuselvan*a and
Nagarajen Rajendraprasadb
aDepartment of Chemistry, Annamalai University, Annamalainagar-608 002, India. E-mail: cas_amu@yahoo.co.in
bDepartment of Biochemistry and Biotechnology, Annamalai University, Annamalai Nagar-608 002, India
First published on 23rd October 2014
Fluorescent chemosensors based on 2-(2-phenylhydrazono)-1H-indene-1,3(2H)-dione (1) and 2-(2-(4-methoxyphenyl)hydrazono)-1H-indene-1,3(2H)-dione (2) with high sensitivity and selectivity toward paramagnetic Cu2+ were developed over other cations. The Cu2+-induced fluorescence turn-on mechanism was revealed to be mediated by intramolecular charge transfer from the ligand to the metal. The compounds were characterized by UV-vis, FT-IR, 1H and 13C NMR techniques, scanning electron microscopy (SEM), energy dispersive spectroscopy (EDS), elemental color mapping and fluorescence spectroscopy. Morphology changes of compounds (1) and (2) to complex with Cu2+ were also investigated by SEM. EDS analysis exposed the occurrence of copper in complexes. Elemental color mapping also supported the copper present in the complexes of corresponding compounds (1) and (2). The metal sensing (chemosensing) properties of compounds (1) and (2) were also examined via fluorescence spectroscopy. The sensor showed excellent selectivity with fluorescence enhancement to copper over other cations in ethanolic solution. The chelating functionality of compounds (1) and (2) was evaluated for inhibitory properties against various human cancer proteins like 4LRH, 4EKD, 4GIW and 4L9K using an online docking server. The cytotoxicity study demonstrated the ability of compounds (1) and (2) to inhibit the growth of KB cell lines. Combined experimental and theoretical studies were carried out on the molecular structure using density functional methods (B3LYP) invoking the 6-31G basis set. The density functional theory (DFT) calculation was carried out for both compounds. Energy of the highest occupied molecular (HOMO) orbital and lowest unoccupied (LUMO) molecular orbitals were predicted.
Expensive and complicated methods, such as atomic absorption,6 inductively coupled plasma, and atomic emission spectroscopy7 have been employed in detection of toxic metal ions. Fluorescent chemosensors are often used to detect many ions due to their simplicity and sensitivity, low cost, high selectivity and quick response.8,9 The classical design of a fluorescent indicator includes two moieties, a receptor responsible for the molecular recognition of the analyte and a fluorophore responsible of signaling the recognition event. Copper is a vital element in the human body, a co-factor in various enzymes and copper-based pigments. Despite its significant roles in organisms, the accumulation of excess amounts of copper ions or their misregulation can cause a number of cruel diseases. It is believed that the disruption of copper homeostasis is implicated in certain neurodegenerative disorders, including Alzheimer's and Parkinson's diseases.10,11 Interest in the design and synthesis of fluorescent chemosensors with a Cu2+-induced “turn-on” fluorescence signal has increased in recent years. However, only a few sensors are currently available because implementation of sensing probes in functional devices without loss of sensitivity remains a major challenge.12 Lu et al. have reported a functional oligonucleotide-based “turn-on” fluorescent probe for detection of Cu2+ in aqueous solution with satisfying sensitivity and selectivity.13 Swamy et al. have proposed a monoboronic acid-conjugated, rhodamine-based probe for Cu2+ with reversible fluorescence “turn-on” response and high selectivity, and they successfully applied it to image Cu2+ in living cells and organisms.14
Aryl hydrazones possess desirable biological and pharmacological properties, such as antibacterial, antiviral, antineoplastic, and antimalarial behaviors.15 The resonance-assisted hydrogen bond systems engage a synergistic reinforcement of hydrogen bonds by delocalization of a p-conjugated chain connecting donor and acceptor atoms; they have been applied for activation of a carbon in α-position to a carbonyl, induced enolization in keto–enol tautomerism, controlled crystal packing, formation of H-bonds in functional molecular materials, activation of dinitriles towards formation of amidines, carboxamides and iminoesters, and so on. Special attention should be focused on the nature of the strong intramolecular O⋯H–N resonance-assisted hydrogen bond and its influence on the enol–azo hydrazone transformation.16–19 The rich tautomerism and isomerism of aryl hydrazones together with the intramolecular resonance-assisted hydrogen bond system can be applied for regulation of tautomerization–isomerization; activation of the carbon in α-position to a carbonyl, antiferroelectric paraelectric transition, regioselective activation of dinitriles, catalysis, ligand liberation, and so on.20 In some cases hydrogen bonding acts as an active site for initiation of chemical reactions. These results led to the observation that the system exists in a conformation with a preferred orientation where the stereoelectronic constraints and stabilizing effect of hydrogen bonding are competing.
In this article, we demonstrate the use of compounds (1) and (2) as a highly efficient fluorescent chemosensor for Cu2+. In terms of sensitivity concerns, chemosensors exhibiting fluorescence enhancement (fluorescence “turn-on”) upon Cu2+ ion complexation are favoured. Among these detection approaches for the Cu2+ ion, fluorescence spectroscopy was used because of its high sensitivity, high selectivity and low cost. Our sensor shows simple and good selectivity compared to recently developed Cu2+ sensors21–28 attributed to the very high association constants for binding Cu2+.
The arylhydrazones of 2-(2-phenyl)hydrazono-2H-indene-1,3-dione (1) and 2-(2-(4-methoxyphenyl)hydrazono)-2H-indene-1,3-dione (2) have been synthesized and their IR spectra for 1 and 2 show ν(NH) vibration at 3421–3431 cm−1, while ν(C
O), ν(C
O⋯H) and ν(C
N) are observed at 1718–1711, 1664–1671 and 1539–1592 cm−1, respectively (Fig. S1 and S2†). These bands can be related to the intramolecular hydrogen-bonded hydrazone fragment. This conclusion was further supported by the 1H and 13C NMR spectrum; the 1H NMR spectra of (1) and (2) showed signals at 13.48–13.63 ppm, which can be assigned to the proton of the NH moiety adjacent to the aryl unit. In compound 2 the methoxy proton appeared at 3.85 ppm. In compound 1 and 2, the aromatic proton signals appear in the downfield region of 6.95–7.96 ppm with the expected splitting patterns. Carbonyl sp2 carbon atoms appear as separate signals in the low field region of about 186 and 188 ppm for C-(1) and C-(2), respectively. Strong intramolecular N–H⋯O–C hydrogen bonding deshields C-(1) with respect to C-(2) to the extent of about 2 ppm.
O) due to intramolecular hydrogen bonding between them. This intramolecular hydrogen bond is responsible for stabilizing the charge transfer band. A shoulder at a shorter wavelength around 480 nm is due to the π–π* transition. In particular, for hydazone derivatives charge transfer between a donor part, N–H (D) and an acceptor part, indene-1,3-dione (C
O), has been considered. Both donor and acceptor appear to be localized in the quasi-aromatic rings (Fig. 9) containing the intramolecular hydrogen bonds. Thus, these hydrogen-bonded, quasi-aromatic rings contribute to stabilization of the charge transfer band.
The emission spectra of compounds (1) and (2) were studied in various solvents. The results are summarized in Table S1† and the spectrum of compounds (1) and (2) in various solvents are depicted in Fig. S8A and B.† Compound (1) showed two emission bands: one band occurred at around 460 nm and another band at around 480 nm. It is assumed that the longer one pertains to charge transfer (CT) and the shorter one to local excitation (LE).31–33 In both compounds, the photoinduced electron transfer (PET) from donor to acceptor produces a low-lying anomalous CT state that leads to new longer wavelength fluorescence. The emission maximum for both states was found to be independent of the excitation wavelength. In protic solvents, protonation of the fluorophore increased the oxidation state of the donor group, which resulted in the LE of fluorescence only. The LE state was more enhanced in intensity when compared to the CT state due to the increased charge density on nitrogen.
Similarly, compound (2) had the LE at 485 nm and the CT band at 515 nm. The variations in solvatochromic properties of compounds (1) and (2) depend on the substitution pattern. The red shift of the emission maximum increases in –OCH3 substituted compound (2) when compared to compound (1). This observation reveals the significant influence of H-bonding on the solvatochromic properties, presumably as a result of stabilization of the negative charge at the carbonyl oxygen atoms during solvent relaxation.
To better understand the solvent polarity effect, the Lippert–Mataga relation was applied. This relation has been widely used to correlate the energy difference between absorption (νa) and emission (νf), also known as the Stokes shift, with solvent polarity represented by Δf. This relation, given in eqn (1), involves both the dielectric constant and the refractive index (n) of the solvents:34,35
![]() | (1) |
In eqn (1), νa and νf are the wavenumbers (cm−1) corresponding to absorption and emission, respectively; h is Planck's constant, c is the speed of light; and a is the radius of the solvent cavity in which the fluorophore resides. For an elongated molecule shape, a is usually estimated as 40% of its longest axis.36 The term (Δf) involving ε and n is called the orientation polarizability, which only accounts for spectral shifts due to reorientation of the solvent molecules. Therefore, the Lippert–Mataga relation is based on the assumption that the energy difference is only proportional to the solvent orientation polarizability (known as the general solvent effect). Inability of the Stokes shift to increase linearly with Δf usually implies that specific solvent effects are involved.
![]() | (2) |
Fig. S9† shows the Lippert–Mataga plots of compounds (1) and (2) in 13 organic solvents. The estimation from the Lippert–Mataga equation is based on the assumption that the photophysical properties of compounds (1) and (2) can be described by the general solvent effect; hence, it may not hold if specific solvent effects are involved. The Stokes shift slightly increases with increasing polarity of the solvent (as expected, based on the CT character of the excited state) and does not follow linear dependence on f(ε, n) in all solvents used, that is, solvent-specific interactions were observed. The linear variation of the Stokes shift with ET (30) (with the solvent polarity parameter) is shown in Fig. S10.† The compounds (1) and (2) exhibited a slight increase or no increase in Stokes shift from non-polar to polar aprotic solvents, mainly due to the combined effect of increasing polarity of the medium and intramolecular charge transfer (CT) state, and it confirmed that solvent polarity cannot affect the Stokes shift of compounds (1) and (2). The νa − νf value is unusually small in hexane, and remarkably, deviates from the linear correlation. This deviation was particularly connected with the decrease of λF in hexane and indicated a different charge redistribution of the excited state of 1 than in other solvents. This suggested that the solvent polarity might not be the only factor affecting the spectral shifts. Specific solvent effects including hydrogen bonding, acid–base chemistry, and charge transfer interactions can also result in nonlinear Lippert–Mataga plots.
:
1, v/v) solution [sensor (1) and (2)] = 1 and 2 μM, [Cu2+] = 1 μM, as expected, sensors (1) and (2) showed weak fluorescence in aqueous solution. Upon addition of different metal ions, sensors (1) and (2) showed remarkable fluorescence enhancement only with Cu2+, due to coordination to a paramagnetic Cu2+ centre (Fig. 1A and 2A).
Based on earlier studies37–39 and our own observations, we proposed the following mechanism. Cu2+ is a well-known paramagnetic ion with an unfilled d shell and could strongly quench the fluorescence of the fluorophore near it via electron or energy transfer. Thus it was of interest that the Cu2+ by compounds (1) and (2) did not quench the fluorescence. It is known that the nitrogen lone pair electrons in compounds (1) and (2) can quench the fluorescence of the indene 1,3-dione moiety through photo-induced electron transfer (PET).40–44 Masking of the nitrogen lone-pair electrons caused suppression of their fluorescence quenching, and therefore, results increased in fluorescence intensity. It is evident from the emission spectra that in the absence of Cu2+ the sensor exhibited weak fluorescence owing to quenching of compounds (1) and (2) fluorescence through PET. Also, in the presence of the quenching metal ions, the relative contribution of two opposing factors, the metal-ion receptor binding that led to enhancement of the fluorescence signal and the fluorophore–metal ion interaction that led to fluorophore quenching determined the net effect. Therefore, it was possible to observe fluorescence enhancement in the presence of quenching ions when metal binding-induced enhancement of fluorescence intensity was greater than the quenching-induced reduction in fluorescence intensity. So, it is clear that for an already PET-quenched system, otherwise strong-quenching metal ions act as poor quenchers. Obviously, in the absence of quenching interaction, metal ions binding with receptors will only lead to fluorescence enhancement. Finally, we conclude that unavailability of the lone-pair electrons of N for PET caused fluorescence enhancement.
The emission at around 500 nm in (1) and 510 nm in (2) did not change when the concentration of Cu2+ increased from 1 μmol L−1 of compounds (1) and (2) in 1
:
1 (volume ratio) CH3CH2OH–H2O (10 mmol L−1 10 μM HEPES buffer, pH = 7.0) binding to Cu2+ with different ion concentrations.
Implied here is a 1
:
1 complex of compounds (1) and (2) with Cu2+ ions (Fig. 1B and 2B). The maximum fluorescence enhancement was observed in the presence of 2 μm (1) and 1.8 μm in (2). When the concentration of Cu2+ was less or equal to around 2 μm, the binding of Cu2+ made the polarization of the molecule larger, and it allowed only partial electron transfer; therefore the fluorescence intensity was enhanced but the emission maximum did not change with the increase of Cu2+ concentration. When the concentration of Cu2+ was more than 2 μm, excess Cu2+ resulted in a sudden drop of relative intensity, no doubt due to concentration quenching.45
Furthermore, the relationship of fluorescence enhancement ratio (defined as I/I0, where I and I0 are the fluorescence intensities of compound (1) and (2) in solution when binding with and without Cu2+, respectively) with the binding concentration of Cu2+ was also investigated. Apparently, the ratio was nearly proportional to the molar binding concentration of Cu2+ (R2 = 0.9941 (1) and R2 = 0.9986 (2)) when the binding concentration of copper was in a range of 0–5.0 μmol L−1 (Fig. 1c and 2c).
The pH emission spectra of sensors (1) and (2) and its ability to detect the Cu2+ cations might be influenced by the solution pH because of the nitrogen atom in hydrazone moiety with lone-pair electrons in the sensors (1) and (2). To verify this hypothesis, the pH effect on the fluorescence intensity of sensors (1) and (2) in unbound and bound Cu2+ was studied. As shown in Fig. S12,† in the unbound Cu2+ (curve b), sensors (1) and (2) presented a low fluorescence intensity that remained stable and did not change with the pH. Upon addition of Cu2+ (curve a), the emission intensity of compounds (1) and (2) with Cu2+ increased dramatically from pH 3 to 5, resulting from competition between the N–H proton and Cu2+ ion.46,47 In particular, slight significant changes in fluorescence spectra were observed in the pH 5–8 range and decreased under alkaline conditions with a color change of yellow to blue. The quenching at higher pH could be well explained by formation of Cu(OH)− or Cu(OH)2, thus reducing the concentration of Cu2+–(1)/(2);48 therefore, there were fewer ions as Cu2+ in solution that were able to be complexed with compounds (1) and (2), which explains the decrease in fluorescence intensity. The results demonstrated that sensors (1) and (2) could be used in environments where pH ranges from 5 to 8, which cover physiological pH conditions.
Elemental mapping FE-SEM was used to confirm the presence of C, O, N, and Cu in the surface of complexes (corresponding to compounds (1) and (2)). Fig. 3a and 4a exhibit FE-SEM color images of complexes (compounds (1)–Cu2+ and (2)–Cu2+); Fig. 3b–e and 4b–e show the elemental mapping for carbon, nitrogen, oxygen, and copper, respectively. It is evident from Fig. 3b–d and 4b–d that carbon, nitrogen, and oxygen are higher in density. There is a homogenous distribution of carbon, nitrogen, oxygen, and copper (Fig. 3a and 4a). Thus, elemental mapping shows that the complex of compounds (1) and (2) contains Cu2+. This also indicates the purity of the complexes (compounds (1)–Cu2+ and (2)–Cu2+). EDS is generally accurate up to trace amounts of metal present in the surface of complexes. The EDS recorded from the selected area is shown in Fig. 3f and 4f, which reveals the presence of C, O, N, and Cu in the complexes.
![]() | ||
| Fig. 3 FE-SEM elemental colour mapping image of compound (1)–Cu2+. (a) Survey spectrum, (b) C, (c) N, (d) O, (e) Cu, and (f) EDS spectra. | ||
![]() | ||
| Fig. 4 FE-SEM elemental colour mapping image of compound (2)-Cu2+. (a) Survey spectrum, (b) C, (c) N, (d) O, (e) Cu, and (f) EDS spectra. | ||
![]() | ||
| Fig. 5 Molecular docking studies of compound (1) with 4LRH. | ||
![]() | ||
| Fig. 6 Molecular docking studies of compound (2) with 4LRH. | ||
Fig. 5c shows the protein residues in the near vicinity (within 4 Å) of the probe, which shows the presence of both hydrophobic residues (TYR-60, TRP-171, and TYR-85), as well as charged/polar residues (TRP-102 (cation-pi), TRP-140, and HIS-135) for compound (1) and hydrophobic residues (TRP-171, TYR-85, PHE-62, TRP-140, and TRP-138), as well as charged/polar residues (TYR-60 (cation-pi), HIS-135 (hydrogen bond), ARG-103) for compound (2) (Fig. 6c). These results indicate that the binding phenomenon was mainly governed by hydrophobic forces with a significant contribution from electrostatic interactions. Asymmetric charge distribution over the molecule due to the presence of heteroatoms might have been responsible for such binding. Based on molecular docking results, we conclude that the alkyl group has been more responsible for biological activity than the aromatic ring.
![]() | ||
| Fig. 7 Live cell images of compound (1): (a) before and (b and c) after treatment with compound (1), examined by fluorescence microscopy. | ||
![]() | ||
| Fig. 8 Live cell images of compound (2): (a) before and (b and c) after treatment with compound (1), examined by fluorescence microscopy. | ||
:
1 complex with Cu2+ ion are given, which display the effective binding sites, namely –NH and C
O for the Cu2+ ion. To obtain insight into electronic behavior in the presence and absence of Cu2+ with compound (1) and compound (2), TD-DFT calculations were carried out. HOMO and LUMO plots for compound (2) showed that the π-electrons on the ligand HOMO focused on one of the ketone groups of the dione moiety and the ligand LUMO focused on the N–NH group of the hydrazone moiety. When Cu2+ was added, it showed a significant change in the distribution of the π-electrons on the HOMO and LUMO ligands. More significantly, the HOMO and LUMO densities in compounds (1) and (2) with Cu2+ were delocalized over the entire molecule except for the phenyl ring of the dione moiety in HOMO and the methoxy phenyl ring in LUMO. The energies of HOMO and LUMO of compounds (1) and (2)–Cu2+ complex were slightly lower than that of compounds (1) and (2). More importantly, the decreasing energy in the LUMO was more obvious than that of the HOMO. It suggests a smaller HOMO–LUMO gap and a bathochromic shift compared with compounds in emission, which may result from the increased conjugation of the complex.
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: 10.1039/c4ra11006d |
| This journal is © The Royal Society of Chemistry 2014 |