A novel 2-(2′-aminophenyl)benzothiazole derivative displays ESIPT and permits selective detection of Zn2+ ions: experimental and theoretical studies

Subramaniyan Janakipriyaa, Selvaraj Tamilmanib and Sathiah Thennarasu*a
aOrganic and Bioorganic Chemistry Laboratory, CSIR-Central Leather Research Institute, Adyar, Chennai-600 020, India. E-mail: thennarasu@gmail.com
bSchool of Chemistry, Bharathidasan University, Tiruchirappalli-24, India

Received 14th April 2016 , Accepted 22nd July 2016

First published on 22nd July 2016


Abstract

Synthesis of a novel 2-(2′-aminophenyl)benzothiazole based probe (1) and demonstration of excited state intramolecular proton transfer (ESIPT) with a large Stokes shift (∼246 nm) are presented. Selective interactions of probe 1 with DMSO and Zn2+ result in the inhibition of ESIPT operating in 1. Inhibition of ESIPT by only Zn2+ and not by other common metal ions permits the selective detection of Zn2+ at concentrations as low as 4.5 × 10−9 M in aqueous medium.


Introduction

Over the past few decades, excited state intramolecular proton transfer (ESIPT) has been exploited for the development of molecular logic gates,1 bioimaging,2 ultraviolet stabilizers,3 light-emitting materials4 and molecular probes.1,4 Generally, ESIPT occurs in molecules containing intramolecular hydrogen bonds through tautomerization during excitation. The excited state is typically characterized by very short lifetimes and low fluorescence quantum yields. However, the large Stokes shift associated with the ESIPT process enables fluorescence detection at longer wavelengths without self-absorption and interference from auto-fluorescence.5

Molecules containing 2-(2′-hydroxyphenyl)benzothiazole (HBT), 2-(2′-hydroxyphenyl)benzoxazole (HBO), 2-(2′-hydroxyphenyl)benzimidazole or 2-(2′-aminophenyl)benzimidazole (APBI) moiety display ESIPT process, and they have been applied to elucidate different analytes.6,7 Recently, the ESIPT process has been studied in a number of interesting Schiff bases that interact with different analytes.8–10 To our knowledge, 2-(2′-aminophenyl)benzothiazole (APBT) derivatives, which display both ESIPT and –CH[double bond, length as m-dash]N– isomerization processes simultaneously, have not been explored for the detection and estimation of trace amounts of analytes.

We report, herein, a new ESIPT based 2-(2′-aminophenyl)benzothiazole (APBT) probe 1 – (E)-N-(2-benzo[d]thiazol-2-yl)phenyl)-2-((2-hydroxybenzilidene)amino)acetamide – with intramolecular hydrogen bond and isomerizable imine moiety. The synthesis of 1 was readily achieved in four steps as shown in Scheme 1. The structure of 1 was determined using 1H, 13C-NMR, HMBC and ESI-MS data (Fig. S1–S4). Incorporation of a phenol-Schiff base with the APBT through an amide linkage conferred N–N–N–O combination of coordinating atoms for chelation with specific metal ions. As with the APBI probes,11 the intramolecularly hydrogen bonded form of 1 is likely to be the stable form in the ground state, and the tautomer form could become dominant in the excited-state (Scheme 1). Interaction of 1 with metal ions can be expected to inhibit ESIPT and –CH[double bond, length as m-dash]N– isomerization processes in 1 and, thereby, permit the detection of specific metal ions.


image file: c6ra09713h-s1.tif
Scheme 1 Synthesis of probe 1.

Results and discussion

We first examined the effect of different solvents on the absorption and fluorescence profiles of 1 to assess the ESIPT phenomenon. The absorption spectrum of 1 in all the solvents chosen for the study was characterized by a broad band in the range from 275 to 375 nm (Fig. 1), except in methanol in which an additional absorption band at ∼400 nm was observed (inset to Fig. 1). The band at ∼400 nm is presumably due to deprotonation of amide hydrogen.7c Another interesting feature of 1 was that the absorption maximum was only slightly sensitive to solvent polarity. These results indicate that the normal form of 1 is the major species at the ground state and that probe 1 is almost insensitive to the surrounding medium.
image file: c6ra09713h-f1.tif
Fig. 1 The absorption (a) and fluorescence (b) spectra of 1 (100 μM) in different solvents.

When excited at the respective absorption maximum of 1 in different solvents, the fluorescence spectra showed comparable emission patterns (Fig. 1b) with a prominent emission band at ∼550, displaying a large Stokes shift (∼240 nm). The huge Stokes shift can only be attributed to the ESIPT process at the excited state and tautomerization to the other form of 1. Strikingly, the fluorescence spectrum of 1 in dimethyl sulfoxide displayed a strong emission at ∼451 nm with a Stokes shift of only 139 nm, presumably due to disruption of intramolecular hydrogen bond needed for ESIPT process.

To validate the solvent dependent variation in ESIPT process, 1H NMR spectra of 1 in CDCl3, DMSO-d6 and CD3CN were examined (Fig. 2). Interestingly, in DMSO-d6, the resonance positions of H1 and H9 are shifted downfield (to 8.78 and 12.71 ppm, respectively), while those of H2 and H8 are shifted upfield (to 8.57 and 11.93 ppm, respectively) compared to their resonance positions in CD3CN and CDCl3. These results are consistent with the observed absorption and fluorescence emission profiles (Fig. 1), and revealed the existence of the intramolecular hydrogen bond and the influence of DMSO on ESIPT process in 1.


image file: c6ra09713h-f2.tif
Fig. 2 1H NMR spectrum of 1 in (a) CDCl3, (b) DMSO-d6 and (c) CD3CN-d3.

Interruption of ESIPT process in 1 by DMSO prompted us to explore common di- and trivalent metal ions for their ability to disrupt the intramolecular hydrogen bond in 1 and inhibit the ESIPT process. Consequently, the absorbance and fluorescence profiles of 1 (10 μM) in aqueous medium (H2O–CH3CN, 3[thin space (1/6-em)]:[thin space (1/6-em)]7, v/v) in the absence and presence of different metal ions (10 equiv.) such as Na+, K+, Zn2+, Co2+, Ca2+, Ni2+, Ba2+, Mg2+, Pb2+, Cu2+, Mn2+, Hg2+, Cd2+, Fe2+, Fe3+, Cr3+ and Al3+ were analysed. To our delight, only Zn2+ induced a significant change in the absorption and fluorescence profiles of 1, while no significant changes were observed upon addition of other metal ions (Fig. 3). Upon addition of Zn2+ ions, the absorption value at ∼320 nm decreased with the appearance of a new band at ∼355 nm (Fig. 3a), presumably due to deprotonation of the amide moiety in 1. An analogous addition of Zn2+ ions produced a strong emission band at ∼459 nm in the fluorescence spectrum, corresponding to the normal form of probe 1 without ESIPT and –CH[double bond, length as m-dash]N– isomerization processes (Fig. 3b). Since the probe 1 shows weak emission at ∼556 nm associated with ESIPT and –CH[double bond, length as m-dash]N– isomerization process, the observed Zn2+ induced emission at ∼459 nm could only be ascribed to the formation of 1-Zn2+ complex and absence of ESIPT and –CH[double bond, length as m-dash]N– isomerization processes in the complex.


image file: c6ra09713h-f3.tif
Fig. 3 Change in the absorption (a) and fluorescence (b) spectra of 1 (10 μM) upon addition of 10 equiv. of various metal ions in H2O[thin space (1/6-em)]:[thin space (1/6-em)]CH3CN (3[thin space (1/6-em)]:[thin space (1/6-em)]7, v/v) medium.

To support the experimental results, time dependent density functional theory (TDDFT) CAM-B3LYP/6-31++G and CAM-B3LYP/6-31G** basis sets12 were used for calculating electronic excitation energy, absorption wavelength and oscillator strength of 1 in different solvents. The calculated absorption maxima for 1 in different solvents13 (ESI, Table S1) are in agreement with the experimentally observed values for 1 (Fig. 1a), and confirm the normal form of 1 as the predominant species in the ground state (Scheme 1).11

To establish the involvement of –NH, –CH[double bond, length as m-dash]N– and –OH groups in ESIPT and –CH[double bond, length as m-dash]N– isomerization processes, FT-IR and 1H-NMR studies were carried out. In the FT-IR spectrum of 1, the stretching vibrations of the –OH and –NH groups occur at 3203.50 and 3180.43 cm−1 (Fig. S5), and they disappear upon complexation with Zn2+ ions. In addition, the vibrational stretching band of C[double bond, length as m-dash]O observed at 1689.23 cm−1 and that of C[double bond, length as m-dash]N observed at 1632.06 cm−1 are shifted to 1644.67 and 1592.09 cm−1, respectively. These observations are convincing proof to indicate the coordination of amide and imine N-atoms. Further, the 1-Zn2+ complex formation was evidenced from 1H-NMR spectra of 1 and 1-Zn2+ complex in CDCl3 (Fig. S6). While the amide (H8) and hydroxyl (H9) proton resonances of 1 disappeared completely upon addition of 1 equiv. of Zn2+, the imine (H1) proton resonance at 8.54 ppm was shifted upfield significantly (Δδ = 0.74 ppm). In addition, the resonance positions of methylene (H10) and aromatic (H2, H3, H6 and H7) protons showed distinct differences (Fig. S7). These observations clearly show the involvement of –CH[double bond, length as m-dash]N– (H1), –OH (H9) and –NH (H8) groups, as well as the N-atom of benzothiazole ring, in the formation of 1-Zn2+ complex.

To gain more insight about the inhibition of ESIPT process in 1-Zn2+ complex, we used the B3LYP/6-31++G and B3LYP/6-31G** basis sets14 embedded in Gaussian 09 to calculate the electronic structures of 1 and 1-Zn2+ complex. Fig. 4 shows the optimized ground state geometries of probe 1 and its 1[thin space (1/6-em)]:[thin space (1/6-em)]1 complex with Zn2+. The LUMO of 1 shows large orbital interactions between the amide and benzothiazole groups in APBT moiety. The electron density localized at the amide group is lesser than that localized at the benzothiazole group (Fig. S7). This difference in electron density indicates that the acidity of the amide N-atom and the basicity of N-atom of benzothiazole are increased upon excitation, which could facilitate proton transfer in the excited state. This proposition is consistent with the experimental finding that 1 displays only ESIPT emission at ∼550 nm.


image file: c6ra09713h-f4.tif
Fig. 4 Optimized geometries of 1 (front view) and 1–Zn2+ complex as predicted by quantum-chemical calculations.

The optimized structure of 1-Zn2+ complex shows the participation of four donor atoms, viz., –CH[double bond, length as m-dash]N– (H1), –OH (H9) –NH– (H8) and N-atom of benzothiazole as shown in Fig. 4. The electron density localized at the N-atom of benzothiazole group in the LUMO becomes weaker in the LUMO+1 of 1-Zn2+ complex (Fig. S7). As the N-atom of benzothiazole becomes less basic, the ESIPT process is inhibited in 1-Zn2+ complex. Such transitions of the HOMO/LUMO orbitals in Zn2+ complexes have been reported.9h Consequently, 1-Zn2+ complex shows transitions from HOMO to LUMO+1 and HOMO−1 to LUMO, contributing mainly to the absorption bands at ∼321 and ∼350 nm, respectively (ESI, Table S2).

The observed selective interaction of 1 with Zn2+ assumes importance as zinc is the second most abundant transition metal ion present in human body (2–3 g) and plays critical roles in gene expression, regulation of metalloenzymes, apoptosis, DNA binding proteins, neurotransmission and pathological diseases.15 Zinc(II) is also a harmful environmental pollutant.16 Therefore, selective sensing of Zn2+ and non-interference of other common metal ions were assessed (Fig. S8).

Spectrophotometric and fluorometric titrations of 1 with serial concentrations of Zn2+ (0–10 equiv.) were carried out in aqueous medium (H2O–CH3CN, 3[thin space (1/6-em)]:[thin space (1/6-em)]7, v/v). As shown in Fig. 5a, the absorption value at ∼320 nm gradually decreased with a concomitant increase in the absorption at ∼355 nm (absorption of deprotonated form of 1). Under similar experimental conditions, the fluorescence emission intensity at 459 nm (Fig. 5b) showed a linear change in the concentration range from 0–20 μM (R2 = 0.9973, Fig. S9). The limit of detection was found to be 4.5 × 10−9 M (detection limit, DL = 3σ/S, where σ is the standard deviation of the blank solution, and S is the slope of the calibration curve).17 The association constant calculated using Benesi–Hildebrand method was 2.4 × 104 M−1 (R2 = 0.9978, Fig. S10).18 The 1[thin space (1/6-em)]:[thin space (1/6-em)]1 stoichiometry of the 1-Zn2+ complex formed was confirmed using Job plot (Fig. S11). Further, the ESI-MS spectrum of 1-Zn2+ complex (Fig. S12) provided additional evidence for the 1[thin space (1/6-em)]:[thin space (1/6-em)]1 binding stoichiometry (m/z = 449.20).


image file: c6ra09713h-f5.tif
Fig. 5 Changes in the absorption (a) and fluorescence spectra (b) of 1 (10 μM) upon serial addition of Zn2+ ions (0–10 equiv.) in H2O[thin space (1/6-em)]:[thin space (1/6-em)]CH3CN (3[thin space (1/6-em)]:[thin space (1/6-em)]7, v/v) medium.

The thermal stability of 1 and 1-Zn2+ complex was also evaluated by thermogravimetric analysis (TGA). The TGA curve of 1 exhibited good thermal stability displaying the onset of decomposition at 175.49 °C (corresponding to 0.89% weight loss), while decomposition of 1-Zn2+ complex was observed at 259.38 °C with a weight loss of 1.15% (Fig. S13). The good thermal stability and minimal weight loss demonstrates the suitability of 1 for environmental applications.

Conclusions

Synthesis of a novel APBT based probe 1 with a tunable ESIPT process is reported. We have also provided experimental and theoretical evidence for the ESIPT process operating in 1. Selective binding of 1 with Zn2+ and formation of 1-Zn2+ complex have been demonstrated using fluorescence, UV-Vis, IR and NMR spectroscopic methods. Binding of Zn2+ with active coordination sites in 1 prevents –CH[double bond, length as m-dash]N– isomerization and inhibits the ESIPT process, as evidenced by the spectroscopic studies and DFT/TDDFT calculations. The observed very low limit of detection of Zn2+ (4.5 × 10−9 M in aqueous medium) and good thermal stability of 1 would be useful for designing new APBT derivatives for tuneable fluorescent sensors.

Experimental

Material and methods

All the materials for synthesis were purchased from Sigma Aldrich, USA, and used as received. Column chromatography was performed on silica gel (200–400 mesh). 1H and 13C NMR spectra were obtained using BRUKER 400 MHz NMR spectrometer. ESI-MS spectra were obtained on a Thermo Finnigan LCQ Advantage MAX 6000 ESI spectrometer. IR spectra (4000–400 cm−1) were recorded using KBr pellets on a Perkin-Elmer Fourier transform infrared spectrophotometer. Analytical grade solvents and double distilled water were used in all experiments. The solutions of different metal ions (Na+, K+, Zn2+, Co2+, Ca2+, Ni2+, Ba2+, Mn2+, Cd2+, Pb2+, Cu2+, Hg2+, Fe2+, Fe3+, Cr3+, and Al3+) were prepared using their chloride salts, and for Mg2+ solution, magnesium sulfate was used. Absorption spectra were recorded on a SPECORD 200 PLUS UV-Visible spectrophotometer. Fluorescence measurements were performed on a Cary Eclipse spectrofluorometer (excitation wavelength: 310 nm, excitation and emission slit width: 5 nm). All measurements were carried out at room temperature.

The geometries of 1 and 1-Zn2+ complex were optimized using DFT B3LYP/6-31++G basis set.14 The vibrational frequency analysis of the optimized geometry confirmed that the optimized geometry corresponded to the minima on the potential energy surface by exhibiting all real frequencies. Based on the optimized ground state geometries, the vertical excitation energies and associated oscillator strengths were obtained by performing TDDFT calculations using the long-range corrected Coulomb attenuated method (CAM-B3LYP functional). The CAM-B3LYP functional was used to get more accurate results for the excited state properties than the B3LYP functional as the compounds 1 and 1-Zn2+ complex show significant charge transfer characteristics.12 The solvent effects on absorption properties was included using the polarizable continuum model (PCM).13 The TDDFT CAM-B3LYP/6-31++G and CAM-B3LYP/6-31G** basis sets were also used to predict the absorption properties in different solvents. All calculations were performed using Gaussian 09 package.

Synthesis of ESIPT probe 1

The probe 1 was synthesized in a four step procedure. To the DMF solution of N-(2-(benzothiazol-2-yl)phenyl)-2-bromoacetamide obtained from the reaction of 2-(2′-aminophenyl)benzothiazole with bromoacetyl bromide,19 sodium azide (0.43 g, 2.6 mmol) was added, and the mixture was stirred at room temperature. After completion of the reaction, the reaction mixture was extracted using DCM, concentrated and purified through column chromatography to obtain 2-azido-N-(2-(benzothiazol-2-yl)phenyl)acetamide (A). To the solution of A (0.63 g) in 10 mL THF[thin space (1/6-em)]:[thin space (1/6-em)]water (7[thin space (1/6-em)]:[thin space (1/6-em)]3), triphenylphosphine (0.73 g, 2.8 mmol) was added, and the mixture was refluxed at 80 °C for 24 h. After completion of the reaction, the reaction mixture was concentrated and subjected to silica gel column chromatography to obtain compound B as pale yellow solid. Salicylaldehyde (0.14 mL, 1.34 mmol) was added to a solution of B (0.32 g, 1.12 mmol) in ethanol at room temperature, with constant stirring. The reaction was monitored through TLC. After completion of the reaction, the pale green solid obtained was filtered, washed with ethanol and dried under vacuum to obtain 1 in 60% (0.26 g) yield. 1H NMR (400 MHz, CDCl3): δ (ppm) 12.70 (s, 1H), 12.31 (s, 1H), 8.75 (d, J = 8 Hz, 1H), 8.54 (s, 1H), 7.83 (d, J = 8 Hz, 1H), 7.51–7.46 (m, 2H), 7.40–7.30 (m, 3H), 7.22–7.16 (m, 2H), 6.99–6.93 (dd, J = 14, 8 Hz, 1H), 4.57 (s, 2H). 13C NMR (100 MHz, CDCl3): δ (ppm) 168.89, 168.26, 168.03, 160.92, 153.07, 137.01, 133.32, 133.21, 132.13, 131.70, 130.14, 126.32, 125.59, 123.95, 123.03, 121.75, 121.16, 120.35, 119.12, 118.79, 117.42, 64.47. IR (KBr, cm−1): 3204, 3180, 3118, 3055, 2922, 2852, 1689, 1632, 1583, 1530, 1489, 1456, 1441, 1408, 1298, 1277, 1213, 974, 845, 754, 739, 729. Mass calculated for [C22H17N3O2S + H]+: 388.46; found: 388.33.

Acknowledgements

One of the authors, S. J., thanks the DST, New Delhi, India for the INSPIRE research fellowship. Financial support from CSIR 12th plan project CSC0201 is acknowledged.

Notes and references

  1. (a) C. Hsieh, C. Jiang and P. Chou, Acc. Chem. Res., 2010, 43, 1364 CrossRef CAS PubMed; (b) V. Luxami and S. Kumar, New J. Chem., 2008, 32, 2074 RSC; (c) P. Singh, H. Singh, G. Bhargava and S. Kumar, J. Mater. Chem. C, 2015, 3, 5524 RSC.
  2. (a) V. V. Shynkar, A. S. Klymchenko, C. Kunzelmann, G. Duportail, C. D. Muller, A. P. Demchenko, J. Freyssinet and Y. Mely, J. Am. Chem. Soc., 2007, 129, 2187 CrossRef CAS PubMed; (b) B. K. Paul and N. Guchhait, J. Phys. Chem. B, 2010, 114, 12528 CrossRef CAS PubMed; (c) O. M. Zamotaiev, V. Y. Postupalenko, V. V. Shvadchak, V. G. Pivovarenko, A. S. Klymchenko and Y. Mély, Bioconjugate Chem., 2011, 22, 101 CrossRef CAS PubMed.
  3. J. Catalan, F. Fabero, M. S. Guijarro, R. M. Claramunt, M. D. S. Maria, M. C. Foces-Foces, F. H. Cano, J. Elguero and R. Sastre, J. Am. Chem. Soc., 1990, 112, 747 CrossRef CAS.
  4. J. Zhao, S. Ji, Y. Chen, H. Guo and P. Yang, Phys. Chem. Chem. Phys., 2012, 14, 8803 RSC.
  5. (a) B. Liu, J. Wang, G. Zhang, R. Bai and Y. Pang, ACS Appl. Mater. Interfaces, 2014, 6, 4402 CrossRef CAS PubMed; (b) S. Goswami, A. Manna, S. Paul, A. K. Maity, P. Saha, C. K. Quah and H. K. Fun, RSC Adv., 2014, 4, 34572 RSC; (c) S. Goswami, A. Manna, M. Mondal and D. Sarkar, RSC Adv., 2014, 4, 62639 RSC; (d) S. Goswami, A. Manna, S. Paul, A. K. Das, P. K. Nandi, A. K. Maity and P. Saha, Tetrahedron Lett., 2014, 55, 490 CrossRef CAS.
  6. (a) B. Liu, H. Wang, T. Wang, Y. Bao, F. Du, J. Tian, Q. Li and R. Bai, Chem. Commun., 2012, 48, 2867 RSC; (b) J. Wang, Q. Chu, X. Liu, C. Wesdemiotis and Y. Pang, J. Phys. Chem. B, 2013, 117, 4127 CrossRef CAS PubMed; (c) Y. Shiraishi, Y. Matsunaga and T. Hirai, J. Phys. Chem. A, 2013, 117, 3387 CrossRef CAS PubMed; (d) Y. Xu and Y. Pang, Chem. Commun., 2010, 46, 4070 RSC; (e) J. Wang, X. Liu and Y. Pang, J. Mater. Chem. B, 2014, 2, 6634 RSC; (f) B. Gu, L. Huang, N. Mi, P. Yin, Y. Zhang, X. Tu and S. Yao, Analyst, 2015, 140, 2778 RSC; (g) M. Santra, B. Roy and K. H. Ahn, Org. Lett., 2011, 13, 3422 CrossRef CAS PubMed; (h) D. Maity, V. Kumar and T. Govindaraju, Org. Lett., 2012, 14, 6008 CrossRef CAS PubMed; (i) L. Tang, M. Cai, P. Zhou, J. Zhao, K. Zhong, S. Hou and Y. Bian, RSC Adv., 2013, 3, 16802 RSC; (j) Y. Xu and Y. Pang, Dalton Trans., 2011, 40, 1503 RSC.
  7. (a) L. Xiong, J. Feng, R. Hu, S. Wang, S. Li, Y. Li and G. Q. Yang, Anal. Chem., 2013, 85, 4113 CrossRef CAS PubMed; (b) J. Wang, W. Chen, X. Liu, C. Wesdemiotis and Y. Pang, J. Mater. Chem. B, 2014, 2, 3349 RSC; (c) Y. Wu, X. Peng, J. Fan, S. Gao, M. Tian, J. Zhao and S. Sun, J. Org. Chem., 2007, 72, 62 CrossRef CAS PubMed; (d) X. Yang, Y. Guo and R. M. Strongin, Angew. Chem., Int. Ed., 2011, 50, 10690 CrossRef CAS PubMed; (e) S. Goswami, A. Manna, S. Paul, A. K. Das, K. Aich and P. K. Nandi, Chem. Commun., 2013, 49, 2912 RSC; (f) S. Goswami, S. Das, K. Aich, B. Pakhira, S. Panja, S. K. Mukherjee and S. Sarkar, Org. Lett., 2013, 15, 5412 CrossRef CAS PubMed; (g) Y. Zhang, J. H. Wang, W. Zheng, T. Chen, Q. X. Tong and D. Li, J. Mater. Chem. B, 2014, 2, 4159 RSC.
  8. (a) J. C. Qin, Z. Y. Yang, L. Fan, X. Y. Cheng, T. R. Li and B. D. Wang, Anal. Methods, 2014, 6, 7343 RSC; (b) L. N. Wang, W. W. Qin, X. L. Tang, W. Dou and W. S. Liu, J. Phys. Chem. A, 2011, 115, 1609 CrossRef CAS PubMed; (c) J. Wang and Y. Pang, RSC Adv., 2014, 4, 5845 RSC; (d) F. Gao, X. Wang, H. Li and X. Ye, Tetrahedron, 2013, 69, 5355 CrossRef CAS; (e) R. Azadbakht, T. Almasi, H. Keypour and M. Rezaeivala, Inorg. Chem. Commun., 2013, 33, 63 CrossRef CAS; (f) F. Gao, X. Ye, H. Li, X. Zhong and Q. Wang, ChemPhysChem, 2012, 13, 1313 CrossRef CAS PubMed; (g) D. Udhayakumari, S. Saravanamoorthy, M. Ashok and S. Velmathi, Tetrahedron Lett., 2011, 52, 4631 CrossRef CAS; (h) V. Kumar, A. Kumar, U. Diwan, S. K. Srivastava and K. K. Upadhyay, Sens. Actuators, B, 2015, 207, 650 CrossRef CAS.
  9. (a) J. Wang, Y. Li, N. G. Patel, G. Zhang, D. Zhou and Y. Pang, Chem. Commun., 2014, 50, 12258 RSC; (b) H. Y. Lin, P. Y. Cheng, C. F. Wan and A. T. Wu, Analyst, 2012, 137, 4415 RSC; (c) A. Kuwar, R. Patil, A. Singh, S. K. Sahoo, J. Marek and N. Singh, J. Mater. Chem. C, 2015, 3, 453 RSC; (d) S. A. Lee, G. R. You, Y. W. Choi, H. Y. Jo, A. R. Kim, I. Noh and C. Kim, Dalton Trans., 2014, 43, 6650 RSC; (e) S. Sharma, M. S. Hundal, A. Walia, V. Vanita and G. Hundal, Org. Biomol. Chem., 2014, 12, 4445 RSC; (f) G. J. Park, Y. J. Na, H. Y. Jo, S. A. Lee, A. R. Kim, I. Noh and C. Kim, New J. Chem., 2014, 38, 2587 RSC; (g) J. Wu, W. Liu, J. Ge, H. Zhang and P. Wang, Chem. Soc. Rev., 2011, 40, 3483 RSC; (h) J. Wang, Y. Li, E. Duah, S. Paruchuri, D. Zhou and Y. Pang, J. Mater. Chem. B, 2014, 2, 2008 RSC.
  10. (a) X. Zhou, B. Yu, Y. Guo, X. Tang, H. Zhang and W. Liu, Inorg. Chem., 2010, 49, 4002 CrossRef CAS PubMed; (b) A. Jiménez-Sánchez, B. Ortíz, V. O. Navarrete, N. Farfán and R. Santillan, Analyst, 2015, 140, 6031 RSC; (c) L. Wang, W. Qin and W. A. Liu, Inorg. Chem. Commun., 2010, 13, 1122 CrossRef CAS; (d) W. H. Hsieh, C. F. Wan, D. J. Liao and A. T. Wu, Tetrahedron Lett., 2012, 53, 5848 CrossRef CAS; (e) L. Li, Y.-Q. Dang, H. W. Li, B. Wang and Y. Wu, Tetrahedron Lett., 2010, 51, 618 CrossRef CAS; (f) J. Cao, C. Zhao, X. Wang, Y. Zhang and W. Zhu, Chem. Commun., 2012, 48, 9897 RSC; (g) Z. Dong, X. Le, P. Zhou, C. Dong and J. Ma, New J. Chem., 2014, 38, 1802 RSC.
  11. S. Santra, G. Krishnamoorthy and S. K. Dogra, J. Phys. Chem. A, 2000, 104, 476 CrossRef CAS.
  12. T. Yanai, D. P. Tew and N. C. Handy, Chem. Phys. Lett., 2004, 393, 51 CrossRef CAS.
  13. (a) J. Tomasi, B. Mennucci and R. Cammi, Chem. Rev., 2005, 105, 2999 CrossRef CAS PubMed; (b) G. Scalmani and M. J. Frisch, J. Chem. Phys., 2010, 132, 114110 CrossRef PubMed.
  14. (a) A. D. Becke, Phys. Rev. A, 1988, 38, 3098 CrossRef CAS; (b) A. D. Becke, J. Chem. Phys., 1993, 98, 5648 CrossRef CAS; (c) C. Lee, W. Yang and R. G. Parr, Phys. Rev. B: Condens. Matter Mater. Phys., 1988, 37, 785 CrossRef CAS; (d) J. Catalán, J. Palomar and J. L. G. De Paz, J. Phys. Chem. A, 1997, 101, 7914 CrossRef; (e) J. Wu, Y. Liao, S. X. Wu, H. B. Lia and Z. M. Su, Phys. Chem. Chem. Phys., 2012, 14, 1685 RSC.
  15. (a) L. A. Finney and T. V. O'Halloran, Science, 2003, 300, 931 CrossRef CAS PubMed; (b) E. M. Nolan and S. J. Lippard, Acc. Chem. Res., 2009, 42, 193 CrossRef CAS PubMed; (c) H. Kozlowskia, A. Janicka-Klosb, J. Brasunb, E. Gaggellic, D. Valensinc and G. Valensinc, Coord. Chem. Rev., 2009, 253, 2665 CrossRef; (d) E. L. Que, D. W. Domaille and C. J. Chang, Chem. Rev., 2008, 108, 1517 CrossRef CAS PubMed; (e) C. F. Walker and R. E. Black, Annu. Rev. Nutr., 2004, 24, 255 CrossRef CAS PubMed; (f) C. R. Cole and F. Lifshitz, Pediatr. Endocrinol. Rev., 2008, 5, 889 Search PubMed; (g) H. Kozlowski, M. Luczkowski, M. Remelli and D. Valensin, Coord. Chem. Rev., 2012, 256, 2129 CrossRef CAS.
  16. (a) A. Voegelin, S. Pfister, A. C. Scheinost, M. A. Marcus and R. Kretzschmar, Environ. Sci. Technol., 2005, 39, 6616 CrossRef CAS PubMed; (b) C. Rensing and R. M. Maier, Ecotoxicol. Environ. Saf., 2003, 56, 140 CrossRef CAS PubMed.
  17. (a) K. B. Kim, H. Kim, E. J. Song, S. Kim, I. Noh and C. Kim, Dalton Trans., 2013, 42, 16569 RSC; (b) Y. Tsui, S. Devaraj and Y. Yen, Sens. Actuators, B, 2012, 161, 510 CrossRef CAS.
  18. H. A. Benesi and J. H. Hildebrand, J. Am. Chem. Soc., 1949, 71, 2703 CrossRef CAS.
  19. (a) N. R. Chereddy, S. Thennarasu and A. B. Mandal, Biochem. Anal. Biochem., 2013, 1, 126 Search PubMed; (b) S. Janakipriya, N. R. Chereddy, P. S. Korrapati, S. Thennarasu and A. B. Mandal, Spectrochim. Acta, Part A, 2016, 153, 465 CrossRef CAS PubMed.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra09713h

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.