Azaindole-1,2,3-triazole conjugate in a tripod for selective sensing of Cl, H2PO4 and ATP under different conditions

Kumaresh Ghosh*a, Debasis Kara, Soumen Joardara, Asmita Samadderb and Anisur Rahaman Khuda-Bukhshb
aDepartment of Chemistry, University of Kalyani, Kalyani-741235, India. E-mail: ghosh_k2003@yahoo.co.in; Fax: +91 3325828282; Tel: +91 3325828750 ext. 305
bDepartment of Zoology, University of Kalyani, Kalyani-741235, India

Received 11th September 2013 , Accepted 13th January 2014

First published on 14th January 2014


Abstract

A new tripodal sensor 1 has been designed and synthesized. The cavity of the tripod selectively recognizes Cl and H2PO4 over a series of other anions in CH3CN containing 0.01% DMSO by exhibiting a significant change in emission. Between H2PO4 and Cl ions, H2PO4 is distinguished by 1 through a ratiometric change in emission. In comparison, the indole-based tripod 2 did not show any binding-selectivity with the same anions. Compound 1, furthermore, selectivity recognizes phosphate based biomolecule ATP over ADP and AMP in semi aqueous solvent (CH3CN containing 0.01% DMSO[thin space (1/6-em)]:[thin space (1/6-em)]H2O, 1[thin space (1/6-em)]:[thin space (1/6-em)]1, v/v) at pH 7.3. The tripod 1 is cell permeable and detects ATP by showing quenching of emission.


Introduction

Anion complexation by design-based synthetic receptors has become an important topic in supramolecular chemistry due to the amazing impact of anions in many chemical and biological processes.1 Among various design-based synthetic receptors, tripodal receptors are of special interest for their preorganized C3-symmetry anion-binding geometry. The tripodal molecular platform with three arms allows the rational control of binding properties such as complex stability and selectivity. Compared to a rigid cyclic system, they can show rapid complexation/decomplexation kinetics and may undergo significant conformational changes upon binding. Therefore, the tripodal receptors are hypothesized to be between cyclic and acyclic ligands with regard to preorganization and are thus believed to be able to complex an ion more effectively than analogous acyclic one.2 A number of synthetic receptors of this class with different functional groups are known in the literature.3 In this context, the use of 7-azaindole as an anion binder around the tripodal core is completely unknown. Recently, for the first time, we have used 7-azaindole in collaboration with a 1,2,3-triazole motif in the construction of an anion binding receptor.4 The promising results have inspired us further to undertake the design of a tripodal-shaped receptor 1 that recognizes different anions such as dihydrogenphosphate, chloride and ATP under different conditions by exhibiting a considerable change in emission in different modes. To establish the binding role of the ring nitrogen atom in 1, a model compound 2 was prepared. Structure 2 did not give any measurable selectivity under identical conditions.
image file: c3ra45018j-u1.tif

It is worth mentioning that the recognition and sensing of inorganic phosphates (Pi), dihydrogenphosphate (H2PO4) and phosphate-based biomolecules such as ATP, ADP and AMP draws attention. Till now, a plethora of receptors that are capable of fluorimetric sensing of H2PO4 anions are known in the literature.5 Among the nucleotides, ATP is very important as it is deeply involved in intercellular energy transfer, DNA duplication and transcription.6 Its concentration in the cell is indicative of cell metabolic rate. So, its selective detection by synthetic receptors is desirable. There are only a few synthetic receptors that are found to recognize these phosphate-based biomolecules.7

On the other hand, chloride ion recognition by an abiotic host is important due to its biological significance. It is associated with a chloride channel that is critically linked to respiration with the exchange of Cl for HCO3 ion in erythrocytes. Disorder of the chloride channel causes the genetic disease cystic fibrosis.8 Therefore, synthetic receptors for chloride recognition is also crucial. It is mentionable that a single structure that recognizes multiple anions by exhibiting different photo physical features is demanding and along this direction the receptor structure 1 in this report draws attention.

Results and discussion

Compounds 1 and 2 were obtained according to Scheme 1. For 1, the tris azide 3 obtained from 1,3,5-tris(bromomethyl)-2,4,6-trimethylbenzene was treated with 5 equiv. 2-alkynyl-7-azaindole 8a4 to couple via click reaction. Compound 8a was obtained from 7-azaindole after performing a series of reactions as mentioned in Scheme 1b. In Scheme 1b, the intermediate compound 6a was synthesized according to the literature procedure.9 Indole-based compound 2 was synthesized following the same reaction pathway as maintained for the synthesis of 1. Use of 5 equiv. 8b4 in the click reaction with 3 afforded compound 2 in appreciable yield. Like 8a, compound 8b was synthesized from indole according to Scheme 1b. In this regard, the intermediate compound 6b, used for the synthesis of 2, was prepared according to the literature procedure.9 Both compounds 1 and 2 were fully characterized spectroscopically.
image file: c3ra45018j-s1.tif
Scheme 1 (a) (i) Acetone–water (4[thin space (1/6-em)]:[thin space (1/6-em)]1, v/v) NaN3, reflux, 3 h; (ii) 8a (5 equiv.), sodium ascorbate, CuSO4, EtOH–H2O (1[thin space (1/6-em)]:[thin space (1/6-em)]1, v/v); (iii) 8b (5 equiv.), sodium ascorbate, CuSO4, EtOH–H2O (1[thin space (1/6-em)]:[thin space (1/6-em)]1, v/v); (iv) PhSO2Cl, K2CO3, dry acetone, reflux, 16 h; (v) LDA, TMEDA, I2, THF, −30 °C; (vi) NaOtBu, dioxane, sealed tube, 80 °C, 3 h; (vii) TMS–acetylene, Pd(PPh3)2Cl2, CuI, Et3N, THF, rt; (viii) TBAF, THF, rt.

Anion recognition studies

Before investigating the anion binding behavior, the absorption and emission spectra of both 1 (Fig. 1) and 2 (ESI) were recorded in DMSO and CH3CN containing 0.01% DMSO. We were unable to record the spectra in other solvents due to insolubility. As can be seen from Fig. 1, the emission of 1 is much changed in DMSO rather than CH3CN. This is attributed to the coordination role of DMSO with the azaindole-1,2,3-triazole conjugate. This is in accordance with our previous observation on dipodal structure.4
image file: c3ra45018j-f1.tif
Fig. 1 Absorbance spectra (a) and emission spectra (b) of 1 (c = 5.08 × 10−5 M) in different solvents.

The anion binding ability (F, Cl, Br, I, AcO, NO3, HSO4, H2PO4 and HP2O73− as their tetrabutylammonium salts) of 1 was evaluated by fluorescence titration in CH3CN containing 0.01% DMSO. In fluorescence, the emission of 1 (c = 5.02 × 10−5 M; λexc = 310 nm) at 363 nm for the azaindole motif underwent change to different extents in the presence of the anions (ESI). Fig. 2 highlights the change in fluorescence ratio of 1 in the presence of 15 equiv. amounts of a particular anion.


image file: c3ra45018j-f2.tif
Fig. 2 Change in fluorescence ratio of 1 (c = 5.02 × 10−5 M) at 369 nm upon addition of 15 equiv. amounts of different anions in CH3CN containing 0.01% DMSO.

Among the anions, the Cl ion brought about a considerable increase in monomer emission of 1 without producing any other change in the spectrum (Fig. 3). The preorganised scaffold of the tripodal core with the triazole ring leads to the formation of a perfect cavity for accommodating the Cl ion. On the other hand, a ratiometric change in emission with H2PO4, which was convenient to distinguish it from the other tested anions, was observed (Fig. 4). This finding is in accordance with our previously reported dipodal receptor.4 In comparison to the dipodal receptor,4 the tripodal receptor 1 in the present report is much more attractive in its ratiometric behavior. In relation to this, the change in emission at the two different wavelengths 369 nm and 475 nm is presented in the inset of Fig. 4. Careful scrutiny reveals that ratiometric chemosensors for H2PO4 are rare in the literature.10,11 The ratiometric chemosensors offer advantages over the usual monitoring of fluorescence intensity at a single wavelength. A dual emission system can minimize the measurement errors because of factors such as phototransformation, receptor concentrations, and environmental effects.10


image file: c3ra45018j-f3.tif
Fig. 3 Change in emission of 1 (c = 5.02 × 10−5 M) upon addition of 15 equiv. Cl (c = 1 × 10−3 M) in CH3CN containing 0.01% DMSO.

image file: c3ra45018j-f4.tif
Fig. 4 Emission titration spectra of 1 (c = 5.02 × 10−5 M) in CH3CN containing 0.01% DMSO upon addition of 15 equiv. H2PO4. Inset: change in emission at the monomer and excimer wavelengths with guest concentration.

Dihydrogenphosphate binding induced quenching of the monomer emission is explained to be due to the perturbation of nπ* state of the azaindole moiety in a stabilizing manner, for which presumably the lowest energy singlet excited state (nπ*) comes closer to the ground singlet state.11 Detailed information supported by theoretical calculation in this aspect has been given in our earlier report on a dipodal system.4 The growth of the emission at the longer wavelength upon complexation is supposed to be due to excimer formation by closely spaced azaindole rings or a strong conjugation established between the azaindole and 1,2,3-triazole motifs. DFT calculation12 using the b3lyp function and 6-31(g) basis set on the complex of 1 with H2PO4 as displayed in Fig. 5A clearly interprets the hydrogen bonding features of the pseudo cavity and the orientation of the azaindole motifs around the tripodal core. In 1, C–H, N–H and ring nitrogens are found to participate actively in making a strong complex with H2PO4. DFT calculation on the complex 1 with Cl ion was also performed. The hydrogen bonding arrangements of the complex are depicted in Fig. 5B. A lower number of hydrogen bonds are noted in the complex and thus corroborates a weak complex.


image file: c3ra45018j-f5.tif
Fig. 5 DFT optimized geometries of the complexes of 1 with (i) H2PO4 (a = 1.70 Å, b = 1.64 Å, c = 1.58 Å, d = 2.82 Å, e = 2.45 Å, f = 2.03 Å, g = 1.54 Å, h = 2.87 Å and i = 1.84 Å) and (ii) Cl (a = 1.83 Å, b = 2.32 Å, c = 3.01 Å, d = 2.88 Å) ions.

Fig. 6 corroborates the comparative views on the emission changes upon addition of 15 equiv. F, Cl, Br, I, NO3, AcO, ClO4, HSO4 and H2PO4 to the solution of 1 in CH3CN containing 0.01% DMSO. The stoichiometries of the complexes of 1 with Cl and H2PO4 were determined to be 1[thin space (1/6-em)]:[thin space (1/6-em)]1 as confirmed by the Job plot13 (ESI). Non linear curve fitting gave binding constant values14 (ESI) of (1.12 ± 0.14) × 104 M−1 with H2PO4 and (2.04 ± 0.29) × 103 M−1 with Cl ions. Under similar conditions, gradual addition of different anions to the solution of 2 in CH3CN containing 0.01% DMSO brought no selectivity in emission. This is easily understandable from Fig. 7. This observation is also similar to the case of our earlier reported dipodal receptor.4


image file: c3ra45018j-f6.tif
Fig. 6 Change in emission of 1 (c = 5.02 × 10−5 M) upon addition of 15 equiv. amounts of different guests (c = 1 × 10−3 M) in CH3CN containing 0.01% DMSO.

image file: c3ra45018j-f7.tif
Fig. 7 Change in fluorescence ratio of 2 (c = 5.07 × 10−5 M) at 369 nm upon addition of 15 equiv. amounts of different guests in CH3CN containing 0.01% DMSO.

To be acquainted with the selectivity of the functional receptor 1 towards Cl and H2PO4 ions, the fluorescence sensing of a particular ion by 1 was understood by recording the emission spectra of 1 in the presence and absence of other anions in CH3CN containing 0.01% DMSO. As shown in Fig. 8, the anions except H2PO4 in the study show negligible interference in the binding of the Cl ion. On the other hand, Fig. 9 displays the selectivity profile where the interference of other anions in the selective sensing of H2PO4 is noted to be negligible. Even the Cl ion did not interfere (Fig. 9). This is in contrast to the case of Fig. 8. This happens due to a stronger affinity of the H2PO4 ion than Cl ion.


image file: c3ra45018j-f8.tif
Fig. 8 Change in fluorescence ratio of 1 (c = 5.08 × 10−5 M) upon addition of 15 equiv. Cl to the receptor solution containing other anions in CH3CN containing 0.01% DMSO.

image file: c3ra45018j-f9.tif
Fig. 9 Change in fluorescence ratio of 1 (c = 5.08 × 10−5 M) upon addition of 15 equiv. H2PO4 to the receptor solution containing other anions in CH3CN containing 0.01% DMSO.

The anion binding of the tripod 1 in the ground state was understood from UV-vis titration. In all cases the absorption centered at 306 nm in 1 decreased upon complexation (ESI). The change was only appreciable for 1 with H2PO4 and Cl ions (Fig. 10).


image file: c3ra45018j-f10.tif
Fig. 10 UV-vis titration spectra for 1 (c = 2.51 × 10−5 M) with (a) Cl and (b) H2PO4 (c = 1 × 10−3 M) in CH3CN containing 0.01% DMSO.

Without having much information in UV, we performed NMR study to realize the actual binding features in the ground state. 1H NMR of 1 in the presence of 1 equiv. Cl was recorded in CD3CN containing 2% d6-DMSO (Fig. 11A). Upon interaction, while the NH proton of the azaindole motif moved downfield slightly (Δδ = 0.07 ppm), the signal for the CH proton of the 1,2,3-triazole ring suffered much downfield chemical shift (Δδ = 0.31 ppm) and thereby confirmed the participation of both the azaindole and triazole motifs in complexation. Unfortunately, we failed to record the 1H NMR of 1 with H2PO4 in the same solvent system due to precipitation. Thus we had to check the interaction of H2PO4 with 1 in d6-DMSO (Fig. 11B). In this solvent, the signal for the triazole ring proton (Fig. 11B) moved downfield by 0.12 ppm upon interaction with H2PO4. The NH proton that appeared at 12.10 ppm gave a downfield chemical shift (Δδ = 0.18 ppm). The upfield chemical shift of the rest aromatic protons is assumed to be due to partial displacement of DMSO by the H2PO4 ion from the interfering zone of 1. DMSO is a hydrogen bond interfering solvent in such systems, as established in the earlier case.4 It severely interferes with the 7-azaindole-1,2,3-triazole conjugate and blocks the cavity from allowing complexation of anions. Comparison of 1H NMR of 1 in CD3CN containing 2% d6-DMSO and pure d6-DMSO (ESI) indicated the change in chemical shifts of the signals and thereby corroborated the interaction role of DMSO. Indeed, the emission of 1 in pure DMSO was also checked with the addition of all anions considered in the study. No characteristic selectivity in the sensing process was observed (ESI).


image file: c3ra45018j-f11.tif
Fig. 11 Partial 1H NMR (400 MHz) of 1 (c = 3.27 × 10−3 M) in the absence (a) and presence of 1 equiv. (b) tetrabutylammonium chloride in CD3CN containing 2% d6-DMSO; (B) partial 1H NMR (400 MHz) of 1 (c = 5.83 × 10−3 M) in the absence (a) and presence of 1 equiv. (b) tetrabutylammonium dihydrogenphosphate in d6-DMSO (for labeling of protons see structure 1).

Furthermore, 31P NMR was used to check the interaction. The signal for the P-atom in H2PO4 that appeared at 1.93 ppm suffered an upfield chemical shift by 0.5 ppm in the presence of 1 in CD3CN containing 4% d6-DMSO (ESI) and indicated a measurable interaction. In d6-DMSO the shift was noted to be less (Δδ = 0.1 ppm). This occurred due to less interaction of H2PO4 for the severe interference of DMSO.

For the application of the tripod 1 in an aqueous system, we studied the complexation of 1 with different phosphate salts, ATP, ADP, AMP and Cl as its potassium salt in CH3CN containing 0.01% DMSO[thin space (1/6-em)]:[thin space (1/6-em)]H2O (1[thin space (1/6-em)]:[thin space (1/6-em)]1, v/v) at pH 7.3 containing 10 mM HEPES buffer. The change in emission of 1 under physiological condition was found to be considerable in the presence of ATP (Fig. 12a) only. Other inorganic phosphate salts including sodium triphosphate brought about minor change in emission (Fig. 12b). Receptor 1 showed 1[thin space (1/6-em)]:[thin space (1/6-em)]1 binding with ATP and gave a binding constant value14 of (8.27 ± 1.89) × 104 M−1 (ESI). The preferential binding of ATP over ADP and AMP is presumably attributed to the pseudo cavity of 1 that accommodates ATP comfortably in comparison to our previously reported dipodal receptor4 involving a greater number of hydrogen bonds. The selective recognition of ATP by 1 in a semi aqueous system was ascertained in the presence and absence of all the phosphate salts undertaken in the study. Fig. 13A represents this feature.


image file: c3ra45018j-f12.tif
Fig. 12 (a) Change in emission of 1 (c = 4.25 × 10−5 M) upon addition of 15 equiv. ATP in CH3CN containing 0.01% DMSO[thin space (1/6-em)]:[thin space (1/6-em)]H2O (1[thin space (1/6-em)]:[thin space (1/6-em)]1, v/v) at pH 7.3 containing 10 mM HEPES buffer; (b) change in fluorescence ratio of 1 (c = 4.25 × 10−5 M) upon addition of 15 equiv. amounts of different guests in CH3CN containing 0.01% DMSO[thin space (1/6-em)]:[thin space (1/6-em)]H2O (1[thin space (1/6-em)]:[thin space (1/6-em)]1, v/v) at pH 7.3 containing 10 mM HEPES buffer.

image file: c3ra45018j-f13.tif
Fig. 13 (A) Change in fluorescence ratio of 1 (c = 4.25 × 10−5 M) at 369 nm upon addition of 15 equiv. ATP in the presence and absence of other anions in CH3CN[thin space (1/6-em)]:[thin space (1/6-em)]H2O (1[thin space (1/6-em)]:[thin space (1/6-em)]1, v/v, 10 mM HEPES buffer) at pH 7.3; (B) partial 31P NMR (400 MHz) of 1 (c = 5.02 × 10−3 M) in the absence (a) and presence (b) of 1 equiv. ATP in CD3CN containing 0.01% d6-DMSO[thin space (1/6-em)]:[thin space (1/6-em)]D2O (1[thin space (1/6-em)]:[thin space (1/6-em)]1, v/v).

The substantial interaction of 1 with ATP was confirmed by recording 31P NMR of 1 in the presence and absence of ATP in CD3CN containing 0.01% d6-DMSO[thin space (1/6-em)]:[thin space (1/6-em)]D2O (1[thin space (1/6-em)]:[thin space (1/6-em)]1, v/v). Considerable shift of the different P-atoms of the phosphate chain in ATP was observed (Fig. 13B) upon interaction. In 1H NMR (taken in d6-DMSO[thin space (1/6-em)]:[thin space (1/6-em)]D2O (1[thin space (1/6-em)]:[thin space (1/6-em)]1, v/v)) upon interaction the signals of aromatic protons were too broad to calculate the chemical shift values exactly (ESI).

Due to good anion sensing properties of 1 with H2PO4 ion in CH3CN containing 0.01% DMSO and ATP in aq. CH3CN containing 0.01% DMSO (CH3CN containing 0.01% DMSO[thin space (1/6-em)]:[thin space (1/6-em)]H2O = 1[thin space (1/6-em)]:[thin space (1/6-em)]1, v/v, pH 7.3 using 10 mM HEPES buffer), we further applied the compound 1 to evaluate its ATP sensing potential in cancer cell lines.

Human cervical cancer cells (HeLa) were incubated with different concentrations of 1 (10.0 mM and 20.0 mM in H2O–CH3CN containing 0.01% DMSO (100[thin space (1/6-em)]:[thin space (1/6-em)]1, v/v; buffered with HEPES, pH 7.0) in a DMEM medium for 20 min at 37 °C and washed with a phosphate-buffered saline (PBS) (pH = 7.4) to remove excess of receptor 1. Microscopic images showed mild fluorescence due to the accumulation of 1 within the cells (Fig. 14b). Compound 1 was non-toxic as confirmed by MTT assay (ESI). However, the subsequent incubation of cells pre-incubated cells with 1 exhibited no fluorescence after further incubation with ATP (5 mM) for 1 h at 37 °C. These results suggest that receptor 1 is highly cell membrane permeable and can be used as a bio-sensor probe to detect/sensitize the intracellular ATP concentration in living cells. ADP and AMP remained silent in the similar study. Further we investigated the potentiality/ability of this biosensor to detect the intercellular cleavage of ATP by ALP. The experiment was carried out with cells pre-incubated with 1 and then further incubated with an equimolar mixture of ATP and ALP. The cells were then observed under microscope after 3 h. Un-interruption of the fluorescence intensity were observed similar to that of 1 with ADP and AMP which confirms the selective detection of ALP mediated cleavage of ATP with 1.


image file: c3ra45018j-f14.tif
Fig. 14 Bright field and fluorescence microscopic images of HeLa cell: (a) bright field image of HeLa cells incubated with receptor 1 (10 mM) for 30 min and subsequently until 3 h; (b) fluorescence image of HeLa cells incubated with receptor 1 (10 mM) for 30 min and subsequently until 3 h; (c) fluorescence image of HeLa cells incubated with receptor 1 (10 mM) for 30 min and subsequently treated with 5 mM ATP for 3 h; (d) fluorescence image of HeLa cells incubated with receptor 1 (10 mM) for 30 min and subsequently treated with 5 mM ADP for 3 h; (e) fluorescence image of HeLa cells incubated with receptor 1 (10 mM) for 30 min and subsequently treated with 5 mM AMP for 3 h; (f) fluorescence image of HeLa cells incubated with receptor 1 (10 mM) for 30 min and after addition of ALP and ATP subsequently until 3 h.

Conclusions

In continuation of our recently reported dipodal 7-azaindole-1,2,3-triazole conjugate,4 the tripodal compound 1 in the present account shows a marked affinity for a variety of anions under different conditions. While the tripod 1 prefers to detect H2PO4 in fluorescence via a sharp ratiometric fashion, Cl ions are sensed through chelation induced fluorescence enhancement in CH3CN containing 0.01% DMSO. On the contrary, the tripod in a semi-aqueous environment [CH3CN containing 0.01% DMSO[thin space (1/6-em)]:[thin space (1/6-em)]H2O (1[thin space (1/6-em)]:[thin space (1/6-em)]1, v/v), pH = 7.3, 10 mM HEPES buffer] exhibits a significant preference for ATP over ADP, AMP and other inorganic phosphate salts. The dimension of the cleft and the hydrogen bonding features of 7-azaindole-1,2,3-triazole conjugate play a pivotal role in the recognition process as supported by theoretical observation also. Experimentation on the model structure 2 gave no meaningful result in emission and thus proved the role of the ring nitrogens at the 7-positions of azaindole motifs of 1 in the recognition process. The tripod 1 is cell permeable and can sense ATP inside the cell. ALP mediated cleavage of ATP can also be assessed by 1 in the cell line.

Experimental section

For the syntheses of compounds 1 and 2, the intermediates 7–8 were obtained according to our recently reported methods.4 The detailed procedures have been cited in the report.

2-((Trimethylsilyl)ethynyl)-1H-pyrrolo[2,3-b]pyridine (7a)

1H NMR (400 MHz, CDCl3) δ 10.14 (1H, s), 8.34 (1H, d, J = 4 Hz), 7.88 (1H, d, J = 8 Hz), 7.08 (1H, dd, J1 = 8 Hz, J2 = 4 Hz), 6.70 (1H, s), 0.28 (9H, s); 13C NMR (100 MHz, CDCl3) δ 148.3, 143.2, 129.4, 120.6, 120.0, 116.2, 106.5, 99.1, 97.0, −0.09; mass (LCMS): 215.0 (M + 1)+.

2-Ethynyl-1H-pyrrolo[2, 3-b]pyridine (8a)

1H NMR (400 MHz, CDCl3) δ 10.2 (1H, s), 8.37 (1H, d, J = 4 Hz), 7.90 (1H, d, J = 8 Hz), 7.09 (1H, dd, J1 = 8 Hz, J2 = 4 Hz), 6.75 (1H, s), 3.38 (1H, s); FT-IR: ν cm−1 (KBr): 3436, 3286, 2897, 2817, 2355, 1694, 1583; mass (LCMS): 144.2 (M + 2)+, 142.8 (M)+.

2-Iodo-7-azaindole (6a)

Compound 6a9a,b was prepared according to the reported procedure. 1H NMR (400 MHz, d6-DMSO) δ 12.20 (1H, s), 8.13 (1H, d, J = 4 Hz), 7.85 (1H, d, J = 8 Hz), 7.01 (1H, dd, J1 = 8 Hz, J2 = 4 Hz), 6.69 (1H, s); mass (LCMS): 245.0 (M + 2)+.

2-((Trimethylsilyl)ethynyl)-1H-indole (7b)

1H NMR (400 MHz, CDCl3) δ 8.16 (1H, s), 7.56 (1H, d, J = 8 Hz), 7.28 (1H, d, J = 8 Hz), 7.21 (1H, t, J = 8 Hz), 7.10 (1H, t, J = 8 Hz), 6.76 (1H, s), 0.26 (9H, s); 13C NMR (100 MHz, CDCl3) δ 135.9, 127.5, 123.7, 121.0, 120.5, 118.6, 110.7, 109.3, 98.5, 97.0, −0.09; mass (LCMS): 213.0 (M)+.

2-Ethynyl-1H-indole (8b)

1H NMR (400 MHz, d6-DMSO) δ 11.6 (1H, s), 7.51 (1H, d, J = 8 Hz), 7.31 (1H, d, J = 8 Hz), 7.15 (1H, t, J = 8 Hz), 7.02 (1H, t, J = 8 Hz), 6.73 (1H, s), 4.46 (1H, s); FT-IR: ν cm−1 (KBr): 3387, 3268, 2923, 2857, 2365, 1607; mass (LCMS): 141.0 (M)+, 140.0 (M − 1)+.

2-Iodoindole (6b)

Compound 6b9c was prepared according to the reported procedure. 1H NMR (400 MHz, CDCl3) δ 8.06 (1H, s), 7.52 (1H, d, J = 8 Hz), 7.32 (1H, d, J = 8 Hz), 7.14–7.05 (2H, m), 6.71 (1H, s); mass (LCMS): 243.0 (M)+.

1,3,5-Tris(azidomethyl)-2,4,6-trimethylbenzene (3)

A mixture of 1,3,5-tris(bromomethyl)-2,4,6-trimethylbenzene (1.5 g, 3.76 mmol) and NaN3 (1.22 g, 18.80 mmol) in acetone–water (4[thin space (1/6-em)]:[thin space (1/6-em)]1 v/v, 30 mL) was refluxed for 3 h. Acetone was evaporated and the mass was extracted with ethyl acetate (20 mL × 3). Evaporation of ethyl acetate afforded compound 3 as a white solid (1 g, yield: 93%). 1H NMR (400 MHz, CDCl3) δ 7.40 (1H, t, J = 8 Hz), 7.29–7.27 (3H, m), 4.36 (4H, s).

Compound 1

To a solution of 3 (300 mg, 1.05 mmol) in ethanol–water (1[thin space (1/6-em)]:[thin space (1/6-em)]1 v/v, 15 mL), alkyne 8a (747 mg, 5.26 mmol) was added. Then a fresh solution of sodium ascorbate (126 mg, 0.6 mmol) and CuSO4·5H2O (78 mg, 0.3 mmol) in distilled water (3 mL) was added and the mixture was stirred under air for 16 h. After completion of the reaction, ethanol was evaporated; the brown solid was washed with distilled water and recrystallized with ethyl acetate to afford product 1 (600 mg, yield: 80%, decomposition temp. 270 °C).

1H NMR (400 MHz, d6-DMSO) δ 12.11 (3H, s), 8.36 (6H, brs), 7.91 (3H, d, J = 8 Hz), 7.07 (3H, brt), 6.84 (3H, s), 5.86 (6H, s), 2.56 (9H, s); 13C NMR (100 MHz, d6-DMSO) δ 148.6, 142.4, 140.5, 140.1, 139.1, 131.3, 130.3, 128.2 (2C unresolved), 121.7, 97.4, 49.1, 16.9; HRMS (TOF MS ES+): calcd for C39H33N15 (M + 1)+ 712.3043: found 712.3094; FT-IR: ν cm−1 (KBr): 3270, 1612, 1489, 1453, 1320, 1220; anal. calcd for C39H33N15: C, 65.81; H, 4.67; N, 29.52; found: C, 65.83; H, 4.70; N, 29.50%.

Compound 2

Following the same procedure as mentioned for compound 1, compound 2 was prepared. The amounts taken were: compound 3 (300 mg, 1.05 mmol), alkyne 8b (742 mg, 5.26 mmol), sodium ascorbate (126 mg, 0.6 mmol), CuSO4·5H2O (78 mg, 0.3 mmol). After completion of reaction, ethanol was evaporated, water was added, extracted with CH2Cl2, organic part was washed with brine solution, and dried over Na2SO4. The solvent was removed under vacuum and the product 2 (610 mg, yield: 81%, mp 165 °C) was isolated by column chromatography using 5% MeOH in CH2Cl2 as eluent.

1H NMR (400 MHz, d6-DMSO) δ 11.51 (3H, s), 8.29 (3H, s), 7.48 (3H, d, J = 8 Hz), 7.36 (3H, d, J = 8 Hz), 7.06 (3H, t, J = 8 Hz), 6.97 (3H, t, J = 8 Hz), 6.78 (3H, s), 5.84 (6H, s), 2.54 (9H, s); 13C NMR (100 MHz, d6-DMSO) δ 141.1, 140.1, 136.9, 131.3, 129.6, 128.6, 121.9, 121.2, 120.4, 119.8, 111.8, 99.1, 49.1, 16.9; LCMS (EI+): 709.2 (M + H)+; FT-IR: ν cm−1 (KBr): 3398, 3305, 1660, 1616, 1453, 1316; anal. calcd for C42H36N12: C, 71.17; H, 5.12; N, 23.71; found: C, 71.21; H, 5.10; N, 23.68%.

Method for MTT assay

Reagents. MTT [3-(4,5-dimethyl-thiazol-2-yl)-2,S-diphenyltetrazolium bromide] and all other reagents were purchased from Sigma-Aldrich Inc. (St-Louis, MO, USA); Dulbecco's modified Eagle's medium (DMEM), fetal bovine serum, penicillin, streptomycin, neomycin (PSN) antibiotics were purchased from Gibco BRL (Grand Island, NY, USA).
Cell culture. Human cervical cancer cell line HeLa was procured from NCCS Pune, India. The cells were cultured at 5 × 105 cells per mL in DMEM supplemented with 10% fetal bovine serum and 1% PSN antibiotic at 37 °C, 5% CO2 for experimental purpose.
Assessment of percentage of viable cells. The percentage viability of HeLa cells, after being exposed to the receptors, was evaluated by MTT assay.15 The cells were incubated in 96-well microplates for 24 hours along with the receptors at different concentrations. A series of normal cells (without any exposure) and a series of positive control cells (exposed to similar concentrations of acetonitrile, the “vehicle solvent” of the receptors as that of the exposure of the receptors) were taken. The intracellular formazan crystals formed were solubilized with dimethyl sulfoxide (DMSO) and the absorbance of the solution was measured at 595 nm by using a microplate reader (Thermo scientific, Multiskan ELISA, USA). The calculated cell survival rate = percentage of MTT inhibition as follows: percentage of survival = (mean experimental absorbance/mean control absorbance) × 100%.

Acknowledgements

D.K. thanks CSIR, New Delhi, India for fellowship.

References

  1. (a) R. Martinez-Manez and F. Sancenon, Chem. Rev., 2003, 103, 4419 CrossRef CAS PubMed; (b) C. Caltagirone and P. A. Gale, Chem. Soc. Rev., 2009, 38, 520 RSC; (c) P. A. Gale, S. E. García-Garrido and J. Garric, Chem. Soc. Rev., 2008, 37, 151 RSC; (d) Z. Xu, N. J. Singh, J. Lim, J. Pan, H. N. Kim, S. Park, K. S. Kim and J. Yoon, J. Am. Chem. Soc., 2009, 131, 15528 CrossRef CAS PubMed; (e) Y. Zhou, Z. Xu and J. Yoon, Chem. Soc. Rev., 2011, 40, 2222 RSC; (f) Z. Xu, N. J. Singh, S. K. Kim, D. R. Spring, K. S. Kim and J. Yoon, Chem.–Eur. J., 2011, 17, 1163 CrossRef CAS PubMed; (g) Z. Xu, S. K. Kim and J. Yoon, Chem. Soc. Rev., 2010, 39, 1457 RSC; (h) M. Wenzel, J. R. Hiscock and P. A. Gale, Chem. Soc. Rev., 2012, 41, 480 RSC.
  2. M. J. Berocal, A. Cruz, I. H. A. Badr and L. G. Bachas, Anal. Chem., 2000, 72, 5295 CrossRef.
  3. (a) G. Hennrich and E. V. Anslyn, Chem. – Eur. J., 2002, 8, 2219 Search PubMed; (b) G. Hennrich, V. M. Lynch and E. V. Anslyn, Chem.–Eur. J., 2002, 8, 2274 CrossRef CAS; (c) C. Schmuck and M. Schwegmann, J. Am. Chem. Soc., 2005, 127, 3373 CrossRef CAS PubMed; (d) B. D. Smith and T. N. Lambert, Chem. Commun., 2003, 2261 RSC; (e) K. Ghosh and I. Saha, Supramol. Chem., 2011, 23, 518 CrossRef CAS.
  4. K. Ghosh, D. Kar, S. Joardar, D. Sahu and B. Ganguly, RSC Adv., 2013, 3, 16144 RSC.
  5. (a) K. H. Chen, J. H. Liao, H. Y. Chang and J. M. Fang, J. Org. Chem., 2009, 74, 895 CrossRef CAS PubMed; (b) A. E. Hargrove, S. Nieto, T. Zhang, J. L. Sessler and E. V. Anslyn, Chem. Rev., 2011, 111, 6603 CrossRef CAS PubMed; (c) A. E. Hargrove, S. Nieto, T. Zhang, J. L. Sessler and E. V. Anslyn, Chem. Rev., 2011, 111, 6603 CrossRef CAS PubMed; (d) V. K. Khatri, S. Upreti and P. S. Pandey, J. Org. Chem., 2007, 72, 10224 CrossRef CAS PubMed; (e) K. Ghosh and D. Kar, Beilstein J. Org. Chem., 2011, 7, 254 CrossRef CAS PubMed; (f) K. Ghosh, D. Kar and P. Ray Chowdhry, Tetrahedron Lett., 2011, 52, 5098 CrossRef CAS PubMed; (g) K. Shin-Ichi, H. Yuichi, K. Namiko and Y. Yumihiko, Chem. Commun., 2005, 1720 Search PubMed; (h) H. Ihm, S. Yun, H. G. Kim, J. K. Kim and K. S. Kim, Org. Lett., 2002, 4, 2897 CrossRef CAS PubMed; (i) K. Choi and A. D. Hamilton, Angew. Chem., Int. Ed., 2001, 40, 3912 CrossRef CAS; (j) Q.-Y. Cao, T. Pradhan, S. Kim and J. S. Kim, Org. Lett., 2011, 13, 4386 CrossRef CAS PubMed; (k) T. H. Kwon and K.-S. Jeong, Tetrahedron Lett., 2006, 47, 8539 CrossRef CAS PubMed; (l) H. Xie, S. Yi, X. Yang and S. Wu, New J. Chem., 1999, 23, 1105 RSC; (m) X.-H. Huang, Y.-B. He, C.-G. Hu and Z.-H. Chen, Eur. J. Org. Chem., 2009, 1549 CrossRef CAS; (n) H. N. Lee, N. J. Singh, S. K. Kim, J. Y. Kwon, Y. Y. Kim, K. S. Kim and J. Yoon, Tetrahedron Lett., 2007, 48, 169 CrossRef CAS PubMed; (o) K. Ghosh, A. R. Sarkar, A. Ghorai and U. Ghosh, New J. Chem., 2012, 36, 1231 RSC; (p) K. Ghosh and I. Saha, Org. Biomol. Chem., 2012, 10, 9383 RSC.
  6. (a) P. Ritter, Biochemistry, a Foundation, Creation Research Society Quarterly June 1999, Brooks/Cole, Pacific Grove, CA, 1996, vol. 36 Search PubMed; (b) W. N. Lipscomb and N. Strater, Chem. Rev., 1996, 96, 2375 CrossRef CAS PubMed; (c) P. Nyron, Anal. Biochem., 1987, 167, 235 CrossRef; (d) T. Tabary and L. Ju, J. Immunol. Methods, 1992, 156, 55 CrossRef CAS; (e) D. Voet, J. G. Voet and C. W. Pratt, Fundamentals of Biochemistry, Wiley-VCH, Weinheim, 2002 Search PubMed.
  7. (a) C. Bazzicalupi, A. Bencini, S. Biagini, E. Faggi, S. Meini, C. Giorgi, A. Spepi and B. Valtancoli, J. Org. Chem., 2009, 74, 7349 CrossRef CAS PubMed; (b) A. Bencini, S. Biagini, C. Giorgi, H. Handel, M. Le Baccon, P. Mariani, P. Paoletti, P. Paoli, P. Rossi, R. Tripier and B. Valtanocli, Eur. J. Org. Chem., 2009, 5610 CrossRef CAS; (c) S. Menuel, R. E. Duval, D. Cue, P. Mutzenhardt and a. Marsura, New J. Chem., 2007, 31, 995 RSC; (d) P. Neelakandan, M. Hariharan and D. Ramaiah, Org. Lett., 2005, 7, 5765 CrossRef CAS PubMed; (e) D. Amilan Jose, S. Mishra, A. Ghosh, A. Shrivastav, S. K. Mishra and A. Das, Org. Lett., 2007, 9, 1979 CrossRef PubMed; (f) H. N. Lee, K. M. K. Swamy, S. K. Kim, J.-Y. Kwon, Y. Kim, S.-J. Kim, Y. J. Yoon and J. Yoon, Org. Lett., 2007, 9, 243 CrossRef CAS PubMed; (g) S. M. Butterfield and M. L. Waters, J. Am. Chem. Soc., 2003, 125, 9580 CrossRef CAS PubMed; (h) Y. Zhou and J. Yoon, Chem. Soc. Rev., 2011, 40, 2222 RSC; (i) N. Ryu and H. Hachisako, Org. Biomol. Chem., 2011, 9, 2000 RSC; (j) H. Wang and W.-H. Chan, Org. Biomol. Chem., 2008, 6, 162 RSC; (k) Y. Kanekiyo, R. Naganawa and H. Tao, Chem. Commun., 2004, 1006 RSC; (l) A. J. Moro, P. J. Cywinski, S. Korsten and G. J. Mohr, Chem. Commun., 2010, 46, 1085 RSC.
  8. (a) C. Huber, T. Werner, C. Krause, I. Klimant and O. S. Wolfbeis, Anal. Chim. Acta, 1998, 364, 143 CrossRef CAS; (b) R. Krapf, C. A. Berry and A. S. Verkman, Biophys. J., 1988, 53, 955 CrossRef CAS; (c) C. D. Geddes, Sens. Actuators, B, 2001, 72, 188 CrossRef CAS; (d) B. J. Rosenstein and P. L. Zeitlin, Lancet, 1998, 351, 277 CrossRef CAS; (e) M. J. Welsh, B. W. Ramsey, R. Accurso and G. R. Cutting, The Metabolic and Molecular Basis of Inherited Disease, ed. C. R. Scriver, A. L. Beaudet, W. S. Sly and D. Valle, McGraw-Hill, New York, 2001, p. 5121 Search PubMed.
  9. (a) B. Joseph, H. D. Costa, J.-Y. Merour and S. Leonce, Tetrahedron, 2000, 56, 3189 CrossRef CAS; (b) C. Chaulrt, C. Croix, J. Basset, M.-D. Pujol and M.-C. Viaud-Massuard, Synlett, 2010, 10, 1481 Search PubMed; (c) S. Hamri, J. Rodriguez, J. Basset, G. Guillaumet and M. Dolors Pujol, Tetrahedron, 2012, 68, 6269 CrossRef CAS PubMed.
  10. G. Grynkiewicz, M. Poenie and R. Y. Tsien, J. Biol. Chem., 1985, 260, 3440 CAS.
  11. A. P. de Silva, H. Q. N. Gunaratne, T. Gunnlaugsson, A. J. M. Huxley, C. P. McCoy, J. T. Rademacher and T. E. Rice, Chem. Rev., 1997, 97, 1515 CrossRef CAS PubMed.
  12. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, J. A. Montgomery, Jr. T. Vreven, K. N. Kudin, J. C. Burant, J. M. Millam, S. S. Iyengar, J. Tomasi, V. Barone, B. Mennucci, M. Cossi, G. Scalmani, N. Rega, G. A. Petersson, H. Nakatsuji, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, M. Klene, X. Li, J. E. Knox, H. P. Hratchian, J. B. Cross, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, P. Y. Ayala, K. Morokuma, G. A. Voth, P. Salvador, J. J. Dannenberg, V. G. Zakrzewski, S. Dapprich, A. D. Daniels, M. C. Strain, O. Farkas, D. K. Malick, A. D. Rabuck, K. Raghavachari, J. B. Foresman, J. V. Ortiz, Q. Cui, A. G. Baboul, S. Clifford, J. Cioslowski, B. B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R. L. Martin, D. J. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng, A. Nanayakkara, M. Challacombe, P. M. W. Gill, B. Johnson, W. Chen, M. W. Wong, C. Gonzalez and J. A. Pople, Gaussian 09 (Revision A.02), Gaussian, Inc., Wallingford CT, 2004 Search PubMed.
  13. P. Job, Ann. Chim., 1928, 9, 113 CAS.
  14. B. Valeur, J. Pouget, J. Bourson, M. Kaschke and N. P. Enesting, J. Phys. Chem., 1992, 96, 6545 CrossRef CAS.
  15. T. Mossam, J. Immunol. Methods, 1983, 65, 55 CrossRef.

Footnote

Electronic supplementary information (ESI) available: Figures showing the fluorescence and UV-vis titrations of receptor 1 with various anions, absorption and emission spectra of 2, Job plot, binding curves, fluorescence titration of 1 in DMSO, 1H and 31P NMR, other spectral data for characterization. See DOI: 10.1039/c3ra45018j

This journal is © The Royal Society of Chemistry 2014