Skip to content
Publicly Available Published by De Gruyter November 10, 2016

A highly selective fluorescent chemosensor based on naphthalimide and Schiff base units for Cu2+ detection in aqueous medium

  • Yuling Xu , Stephen Opeyemi Aderinto , Huilu Wu EMAIL logo , Hongping Peng , Han Zhang , Jiawen Zhang and Xuyang Fan

Abstract

A new Schiff base, 4-allylamine-N-(N-5-methylsalicylidene)-1,8-naphthalimide (1), has been designed and synthesized by combining a 1,8-naphthalimide moiety as a fluorophore and a Schiff base as a recognition group. Its photophysical properties were investigated by absorption and fluorescence spectroscopy, and this sensor exhibits a high fluorescence quantum yield of 0.75–0.91 inorganic solvents of different polarity. It also shows high selectivity for Cu2+ over other ions with fluorescence quenching in aqueous medium (pH=7.2). The reason for this phenomenon (fluorescence quenching) is attributed to the formation of a 1:1 complex between 1 and Cu2+ according to the Job plot and fluorescence titration. The sensor can be applied to the quantification of Cu2+ in a linear fashion from 0.5 to 5 μM with a detection limit of 0.23 μM. Additionally, the association constant (Ka) between Cu2+ and 1 is 1.328×106 M1 in aqueous media.

1 Introduction

As the third most abundant essential trace element after iron and zinc in the human body, copper performs an important role in many fundamental physiological processes in the human body, such as functional and structural strengthening of protein and gene expression, even in the human body’s nervous system [1], [2]. It is well known that it serves as a catalytic cofactor for a variety of metalloenzymes [3], but unregulated large intake of copper can result in severe neurodegenerative diseases such as Menkes and Wilson diseases, familial amyotrophic lateral sclerosis, Alzheimer’s disease and prion diseases, and can cause infant liver damage [4], [5], [6]. Thus, the selective measurement and monitoring of Cu2+ in environmental and biological samples are of considerable significance for environment protection and human health.

Till date, the available methods for Cu2+ detection include colorimetry [7], [8], [9], [10], [11], [12], electrochemistry [13], [14], room temperature phosphorimetry [15], atomic absorption spectrometry [16], [17], inductive coupled plasma-mass spectrometry [18] and fluorescence sensoring [19], [20], [21], [22], [23]. During the past few decades, the fluorescence chemosensors, which combine a recognition functional moiety with optical transduction, have received much attention as an efficient analytical technique for the detection of a particular species due to their high selectivity and sensitivity, intrinsic specificity, real-time monitoring and fast response time [24], [25], [26]. Thereby, many fluorescence sensors based on quinoline [27], fluorescein [28], pyrene [29], [30], rhodamine [31], [32], azobenzene [33], coumarin [34], [35], [36], 1,8-naphthalimide [37], [38], [39], anthraquinone [40], [41] and other fluorophores [42], [43] for Cu2+ sensing have been synthesized and investigated. However, there is still an intense demand for efficient Cu2+ fluorescence chemosensors, especially those that can work in aqueous solution with high selectivity and sensitivity.

Herein, we developed a new sensor based on a naphthalimide framework. The selectivity of 1 towards various transition metal cations was investigated. It could be used to detect and recognize a Cu2+ ion with high selectivity and a low detection limit in aqueous solution.

2 Results and discussions

The synthesis of compound 1 was achieved in three steps as shown in Scheme 1. It is soluble in dipolar aprotic solvents such as THF, DMF, DMSO, acetonitrile, acetone and dichloromethane, and insoluble in Et2O and water. The structure and purity of the synthesized compound 1 were confirmed by conventional techniques – melting point, thin-layer chromatography (TLC) (Rf values) and UV/Vis spectra – and identified by 1H and 13C NMR spectra, FT-IR and elemental analysis data. The data are presented in Section 3.

Scheme 1: Chemical structure of and synthetic route to compound 1.
Scheme 1:

Chemical structure of and synthetic route to compound 1.

2.1 Photophysical characteristics of 1

The photophysical properties of the substituted 1,8-naphthalimides are known to depend mainly on the polarization of their chromophoric system. Light absorption in this molecule generates a charge transfer interaction between the substituents at C-4 position and the imide carbonyl groups. Therefore, the photophysical characteristics of 1 in organic solvents with different polarity have been investigated. Table 1 exhibits its absorption (λA) and fluorescence (λF) maxima, the extinction coefficient (ε), the Stokes shift (ʋAʋF) and the quantum fluorescence yield (ФF).

Table 1:

Photophysical properties of 1 in organic solvents with different polarity.

Organic solutionλA (nm)ε (M−1 cm−1)λF (nm)ʋAʋF (cm−1)ФF
THF42913 26350836250.91
DMF44018 60352637160.75
Acetonitrile43311 75152239380.76
Acetone432842251738060.80
Dichloromethane42611 22350235540.77

Compound 1 absorbs in the near-UV region with maxima at λA=426–440 nm, and the absorption can be ascribed to π–π* transitions of the naphthalene ring. The respective fluorescence maxima are at 502–526 nm. The molar extinction coefficient (ε) of 8422–18 603 M1 cm1 corresponds to an S0 → S1 transition.

The polarity of the organic solvents is of great importance for the photophysical properties of 1 under study and especially for the quantum fluorescence yield and Stokes shift. The Stokes shift is an important parameter, which indicates the difference in the properties and structure of the fluorophore between the ground state S0 and the first exited state S1. The Stokes shift was calculated by eq. 1.

(1)(υAυF)=(1/λA1/λF)×107cm1

The obtained Stokes shift values are in the 3554–3938 cm1 region. Table 1 reveals that the Stokes shift depends on the media and it is larger in the case of polar solvents where hydrogen bond formation or dipole–dipole interactions are favored in comparison to non-polar media. The ability of 1 to emit absorbed light energy is characterized quantitatively by the quantum yield of fluorescence ФF. The fluorescence quantum yield has been calculated on the basis of the absorption and fluorescence spectra using N-butyl-4-n-butylamino-naphthalimide (ФF=0.81 in ethanol) according to eq. 2.

(2)ΦF=Φref(Ssample/Sref)(Aref/Asample)(nsample/nref)2

Herein, ФF is the emission quantum yield of the sample, Фref is the emission quantum yield of the standard, Aref and Asample represent the absorbance of the standard and sample at the excited wavelength, respectively, whereas Sref and Ssample are the integrated emission band areas of the standard and sample, respectively, and nref and nsample are the solvent refractive indices of the standard and sample, respectively. It is seen that the polarity of the environment has no obvious influence on the ФF values (Table 1).

2.2 Influence of pH on the photophysical properties of 1

The spectroscopic characteristics of compound 1 were performed in the pH range 1.81–11.82 in Britton–Robinson buffer–DMF (v/v = 1:1) solution. The fluorescence emission spectra were recorded at room temperature with an excitation at 430 nm. It is obvious from Fig. 1 that the fluorescence intensity (539 nm) of 1 decreases dramatically when the pH value goes down from 3.78 to 1.81 or goes up from 7.96 to 11.82. However, the fluorescence intensity of 1 is high and relatively stable in the pH range from 3.78 to 7.96, implying that 1 was actually pH-independent between pH 3.78 and 7.96. Therefore, the following experiments were carried out in solution at pH 7.2.

Fig. 1: Photophysical behavior of 1 in Britton–Robinson buffer–DMF (1:1) solution at different pH values at 539 nm.
Fig. 1:

Photophysical behavior of 1 in Britton–Robinson buffer–DMF (1:1) solution at different pH values at 539 nm.

2.3 The fluorescence emission spectroscopy of 1 to Cu2+

High selectivity for a specific analyte in the presence of a competing species is an important feature of a chemosensor. To examine the selectivity of 1 (5 μM) towards various transition metal ions [Na+, K+, Ca2+, Mg2+, Al3+, Pb2+, Fe3+, Ni2+, Zn2+, Cu2+, Hg2+, Ag+, Co2+, Cr3+, Mn2+ and Cd2+ (10 μM, 2 eq)], the fluorescence spectroscopic properties were investigated in Tris-HCl (pH=7.2) buffer–DMF (1:1, v/v) solution. The fluorescence emission spectra were collected at room temperature with an excitation at 430 nm. Both the excitation and emission slit widths were 1.0 nm. As seen from Fig. 2a, we found that only Cu2+ caused a noteworthy fluorescence decrease at 539 nm, whereas other metal ions induced only negligible fluorescence variations. The selectivity difference can be assigned to the different binding abilities of these metal ions to this sensor. These results indicate that 1 could be used as a Cu2+ selective fluorescent sensor.

Fig. 2: (a) Fluorescence spectra of 1 (5×10− 6 M) in Tris-HCl (pH=7.2) buffer–DMF (1:1, v/v) solution towards Na+, K+, Ca2+, Mg2+, Al3+, Pb2+, Fe3+, Ni2+, Zn2+, Cu2+, Hg2+, Ag+, Co2+, Cr3+, Mn2+ or Cd2+ (2 eq) at 539 nm (lambdaex=430 nm); (b) Competitive experiments in the compound 1 + Cu2+ system with interfering metal ions at 539 nm (lambdaex=430 nm); (c) Fluorescence titration spectra of 1 with Cu2+ (0–2 eq) in Tris-HCl (pH=7.2) buffer–DMF (1:1, v/v) solution at 539 nm (lambdaex=430 nm); (d) Job’s plot at 539 nm (lambdaex=430 nm).
Fig. 2:

(a) Fluorescence spectra of 1 (5×10− 6 M) in Tris-HCl (pH=7.2) buffer–DMF (1:1, v/v) solution towards Na+, K+, Ca2+, Mg2+, Al3+, Pb2+, Fe3+, Ni2+, Zn2+, Cu2+, Hg2+, Ag+, Co2+, Cr3+, Mn2+ or Cd2+ (2 eq) at 539 nm (lambdaex=430 nm); (b) Competitive experiments in the compound 1 + Cu2+ system with interfering metal ions at 539 nm (lambdaex=430 nm); (c) Fluorescence titration spectra of 1 with Cu2+ (0–2 eq) in Tris-HCl (pH=7.2) buffer–DMF (1:1, v/v) solution at 539 nm (lambdaex=430 nm); (d) Job’s plot at 539 nm (lambdaex=430 nm).

To further examine the affinity of 1 to Cu2+, competitive binding between Cu2+ and other metals to 1 was measured by using a mixed solution containing Cu2+ (10 μM) and each of the other metal cations (Na+, K+, Ca2+, Mg2+, Al3+, Pb2+, Fe3+, Ni2+, Zn2+, Hg2+, Ag+, Co2+, Cr3+, Mn2+ and Cd2+) at a concentration of 10 μM in Tris-HCl (pH=7.2) buffer–DMF (1:1, v/v) solution. From Fig. 2b, we can find that all competitive metal ions had no obvious interference with the detection of a Cu2+ ion, which indicated that the compound 1–Cu2+ system was hardly affected by these ions.

In order to get further insight into the binding interaction between 1 and Cu2+, a fluorescence titration of 1 with Cu2+ was carried out and showed an emission maximum at 539 nm. It can be seen from Fig. 2c that on the addition of a Cu2+ ion, the fluorescence intensity of 1 gradually decreased, and when the amount of Cu2+ ion added was about 5×10− 6 M, the fluorescence intensity reached a minimum. When more Cu2+ was added, the fluorescence intensity showed negligible changes. The nonlinear curve fitting of the fluorescence titration gives a 1:1 stoichiometric ratio between compound 1 and Cu2+. Moreover, a Job’s plot [44], which exhibits a maximum at 0.5 M fraction of Cu2+, further indicates that only a 1:1 complex is formed (Fig. 2d).

Based on the fluorescence titration of 1 with Cu2+, the association constant has been calculated to be 1.328×106 M− 1 (error limits ≤ 10%) by a Benesi–Hildebrand equation [45], [46], [47] (Fig. 3a and eq. 3).

Fig. 3: (a) Benesi–Hildebrand linear analysis plots of 1 at different Cu2+ concentrations; (b) curve of fluorescence intensity at 539 nm of 1 (5×10− 6 M) versus increasing concentrations of Cu2+ (0.5–5 μM).
Fig. 3:

(a) Benesi–Hildebrand linear analysis plots of 1 at different Cu2+ concentrations; (b) curve of fluorescence intensity at 539 nm of 1 (5×10− 6 M) versus increasing concentrations of Cu2+ (0.5–5 μM).

(3)1/(FF0)=1/{Ka(FmF0)[Cu2+]n}+1/(FmF0)

Herein, F is the fluorescence intensity at 539 nm at any given Cu2+ concentration, F0 is the fluorescence intensity at 539 nm in the absence of Cu2+ and Fm is the minimum fluorescence intensity at 539 nm in the presence of Cu2+ in solution. The association constant Ka was evaluated graphically by plotting log[(FF0)/(FmF)] against log[Cu2+].

Figure 3b also shows good linearity relating the emission at 539 nm and the concentrations of Cu2+ in the range from 0.5 to 5 μM, indicating that 1 can detect quantitatively relevant concentrations of Cu2+.

The detection limit based on the definition by IUPAC was calculated by the Stern–Volmer plot [48]:

CDL=3σ/k=0.23μM

where σ is the standard deviation of the blank solution, and k is the slope between F0/F and [Cu2+]. In order to determine σ, the emission intensity of 1 in Tris-HCl (pH=7.2) buffer–DMF (1:1, v/v) solution without any metal ions was measured 10 times.

2.4 The UV/Vis absorption spectroscopic response of 1 to Cu2+

Additionally, a UV/Vis titration experiment of 1 was performed with Cu2+ in Tris-HCl (pH=7.2) buffer–DMF (1:1, v/v) solution. Figure 4 shows the UV/Vis absorption spectra of 1 (5×10− 6 M) in the presence of various concentrations of Cu2+ ion (0–1×10− 5 M, 0–2 eq), and the inset shows the plots of changes in the 334 and 376 nm maxima as a function of increasing concentrations of Cu2+. The absorbance of 1 at 334 nm gradually decreases with an increasing concentration of Cu2+ ions; this change is clear evidence of C=N coordination to a Cu2+ ion. Owing to the coordination of 1 with a Cu2+ ion, the absorbance at 376 nm increases with an increasing concentration of Cu2+ ions. As Cu2+ was gradually titrated, the absorbance of 1 at 334 and 376 nm reached extrema when the amount of Cu2+ ion added was about 5×10− 6 M. When more Cu2+ was titrated, the absorbance showed negligible changes, implying that a 1:1 complex was formed. Moreover, two isosbestic points appear at 307 and 360 nm. The absorptions at 274 and 449 nm can be attributed to π–π* transitions of the benzene and naphthalene rings, respectively.

Fig. 4: UV/Vis titration of 1 (5×10− 6 M) in Tris-HCl (pH=7.2) buffer–DMF (1:1, v/v) solution with increasing amount of Cu2+ (0–1×10− 5 M, 0–2 eq); the absorbance plot of 1 against increasing [Cu2+] at λ334nm and λ376nm is shown in the inset.
Fig. 4:

UV/Vis titration of 1 (5×10− 6 M) in Tris-HCl (pH=7.2) buffer–DMF (1:1, v/v) solution with increasing amount of Cu2+ (0–1×10− 5 M, 0–2 eq); the absorbance plot of 1 against increasing [Cu2+] at λ334nm and λ376nm is shown in the inset.

Coupling fluorescence titration, Job’s plot and UV/Vis titration analysis together, we could primarily confirm that the 1:1 stoichiometry for the 1–Cu(II) complex was formed as shown in Scheme 1. The results obtained were consistent with the reported reference [49].

3 Experimental section

3.1 Apparatus and reagents

Absorption spectra were scanned using a Lab-Tech UV Bluestar spectrophotometer with one pair of 10-mm quartz cells at room temperature. The IR spectra were recorded in the 4000–400 cm− 1 region with a Nicolet FT-VERTEX 70 spectrometer using KBr pellets. Mass spectra were recorded on a Mass Spectrometer micrOTOF. 1H NMR and 13C NMR spectra were acquired on a Mercury plus 400 MHz NMR spectrometer with tetramethylsilane (TMS) as an internal standard and [D6]DMSO as solvent. The corrected excitation and fluorescence spectra were taken on an F97Pro fluorescence spectrophotometer. A quartz cuvette (1×1 cm2) was used for the spectroscopic analysis. TLC was performed on silica gel, Fluka F60 254, 20×20, 0.2 mm. The melting points were determined by means of a Kofler melting point microscope.

All chemicals and solvents used in the synthesis were analytical reagent grade and used without further purification. N-Butyl-4-n-butylamino-naphthalimide (ФF=0.81 in ethanol) was synthesized according to the literature [50]. Britton–Robinson buffer was prepared with 40 mM acetic acid, boric acid and phosphoric acid. Dilute hydrochloric acid or sodium hydroxide was used to adjust pH values. Tris-HCl buffer (pH=7.2) was prepared using bidistilled water. The metal cation sources, NaNO3, KNO3, CaCl2, MgSO4, AlCl3, Pb(NO3)2, Fe(NO3)3⋅9H2O, Ni(NO3)2⋅6H2O, Zn(NO3)2⋅6H2O, Cu(NO3)2⋅5H2O, Hg(NO3)2⋅H2O, AgNO3, Co(NO3)2⋅6H2O, Cr(NO3)3⋅9H2O, 50% Mn(NO3)2 and Cd(NO3)2⋅H2O, were analytical reagent grade and were dissolved in bidistilled water.

3.2 Synthesis

The solution of 4-bromo-1,8-naphthalic anhydride (13.85 g, 50 mmol) and 80% hydrazine hydrate (6.25 g, 100 mmol) in 150 mL of ethanol was refluxed with stirring for 4 h [51]. The reaction’s progress was monitored by TLC (with CH2Cl2 as the eluent). The mixture was cooled and the precipitated solids were filtered, recrystallized from ethanol and dried to give 14.5 g (85.6%) of N-amido-4-bromo-1,8-naphthalimide (2) as yellow-brown crystals. – 1H NMR (CDCl3): δ (ppm)=8.68 (d, 1H, J=7.2 Hz), 8.60 (d, 1H, J = 8.4 Hz); 8.44 (d, 1H, J = 7.6 Hz); 8.06 (d, 1H, J = 8.06 Hz), 7.87 (t, 1H, J = 7.6 Hz), 5.30 (s, 2H).

N-Amido-4-allylamino-1,8-naphthalimide (3) was prepared from 7.35 g (25 mmol) of N-amido-4-bromo-1,8-naphthalimide (2), 7.13 g (75 mmol) of allylamine and 100 mL of 2-methoxyethanol. The resulting mixture was refluxed and stirred for 64 h and then poured into 500 mL of water. The precipitate was collected by filtration, washed with water and dried to give 6.25 g (93%) of N-amido-4-allylamino-1,8-naphthalimide (3); m.p.: 180–182°C. – Elemental analysis for C15H13N3O2: calcd. C 67.41, H 4.90, N 11.72; found C 67.93, H 5.34, N 12.34.

N-Amido-4-allylamino-1,8-naphthalimide (3) (5.00 g, 18.75 mmol) was dissolved in absolute ethanol (40 mL). An excess of 5-methylsalicylaldehyde (3.9 g, 28.7 mmol) was added and the mixture was refluxed for 6 h. After the mixture was cooled to room temperature, the precipitate produced was filtered, washed with water (3×30 mL) and dried to give a yellow crude product. The crude product was purified by column chromatography on flash silica gel using dichloromethane–acetone (v:v = 20:1, Rf=0.73) as the eluent to give 2.1 g (29%) of 4-allylamino-N-(N-5-methylsalicylidene)-1,8-naphthalimide (1); m.p.: 210–211°C. – IR (KBr; v/cm− 1): 1701, 1653, 1637, 1582, 1534, 1335, 914, 826. – UV/Vis (in DMF, nm): 273, 327, 440. – MS: m/z=386.1857 ([C23H19N3O3]+1). – 1H NMR ([D6]DMSO, 400 MHz): δ (ppm)=8.91(s, 1H), 8.73 (d, 1H, J = 8 Hz), 8.48 (d, 1H, J=7.2 Hz); 8.28 (d, 1H, J = 8 Hz), 7.73–7.75 (m, 1H), 7.57–7.70 (m, 1H), 7.29–7.30 (m, 1H), 6.95 (d, 1H, J = 8.4 Hz), 6.75(d, 1H, J = 8.8 Hz), 5.96–6.03 (m, 1H), 5.19–5.31 (m, 2H), 4.09 (s, 2H), 2.29 (s, 3H). – 13C NMR ([D6]DMSO, 400 MHz): δ (ppm)=169.47, 160.73, 160.11, 156.62, 150.86, 134.70, 134.47, 134.11, 131.23, 131.02, 130.68, 130.51, 128.79, 128.35, 124.56, 124.34, 121.88, 120.10, 117.41, 116.76, 107.50, 44.97, 19.96. – Elemental analysis for C23H19N3O3: calcd. C 71.67, H 4.97, N 10.90; found: C 71.63, H 4.98, N 10.93.

4 Conclusions

In conclusion, we have developed a highly selective fluorescence sensor for a Cu2+ ion based on the naphthalimide framework. Compound 1 exhibits a highly selective response to Cu2+ over other metal ions in Tris-HCl (pH=7.2) buffer–DMF (1:1, v/v) solution with a detection limit of 0.23 μM and an association constant of 1.328×106 M− 1. UV/Vis titration and fluorescence titration experiments and Job plot analysis of 1 to Cu2+ indicate the formation of a 1:1 stoichiometric complex. The good selectivity and sensitivity in aqueous medium effectively enhance the application value of the chemosensor for Cu2+.

4.1 Supplementary information

Figures of the IR, UV/Vis, NMR and mass spectra of compound 1 are given as Supplementary Information available online (DOI: 10.1515/znb-2016-0138).

Acknowledgments

The present research was supported by the National Natural Science Foundation of China (Grant No. 21367017), Natural Science Foundation of Gansu Province (Grant No. 1212RJZA037) and Graduate Student Innovation Projects of Lanzhou Jiaotong University.

References

[1] R. Uauy, M. Olivares, M. Gonzalez, Am. J. Clin. Nutr.1998, 67, 952S.10.1093/ajcn/67.5.952SSearch in Google Scholar

[2] H. Tapiero, D. M. Townsend, K. D. Tew, Biomed. Pharmacother.2003, 57, 386.10.1016/S0753-3322(03)00012-XSearch in Google Scholar

[3] K. M. K. Swamy, S. K. Ko, S. K. Kwon, H. N. Lee, C. Mao, J. M. Kim, K. H. Lee, J. Kim, I. Shin, J. Yoon, Chem. Commun.2008, 5915.10.1039/b814167cSearch in Google Scholar

[4] G. Muthaup, A. Schlicksupp, L. Hess, D. Beher, T. Ruppert, C. L. Masters, K. Beyreuther, Science1996, 271, 1406.10.1126/science.271.5254.1406Search in Google Scholar

[5] S. H. Hahn, M. S. Tanner, D. M. Danke, W. A. Gahl, Biochem. Mol. Med.1995, 54, 142.10.1006/bmme.1995.1021Search in Google Scholar

[6] B. P. Zietz, J. Dassel de Vergara, H. Dunkelberg, Environ. Res.2003, 92, 129.10.1016/S0013-9351(03)00037-9Search in Google Scholar

[7] J. M. Liu, H. F. Wang, X. P. Yan, Analyst2011, 136, 3904.10.1039/c1an15460eSearch in Google Scholar PubMed

[8] Y. J. Song, K. G. Qu, C. Xu, J. S. Ren, X. G. Qu, Chem. Commun.2010, 46, 6572.10.1039/c0cc01593hSearch in Google Scholar PubMed

[9] Y. Zhou, S. X. Wang, K. Zhang, X. Y. Jiang, Angew. Chem. Int. Ed.2008, 47, 7454.10.1002/anie.200802317Search in Google Scholar PubMed

[10] B. C. Yin, B. C. Ye, W. H. Tan, H. Wang, C. C. Xie, J. Am. Chem. Soc.2009, 131, 14624.10.1021/ja9062426Search in Google Scholar PubMed PubMed Central

[11] T. Gunnlaugsson, J. P. Leonard, N. S. Murray, Org. Lett.2004, 6, 1557.10.1021/ol0498951Search in Google Scholar

[12] Y. Zhou, H. Zhao, Y. J. He, N. Ding, Q. Cao, Colloids Surf. A2011, 391, 179.10.1016/j.colsurfa.2011.07.026Search in Google Scholar

[13] P. Salaün, C. M. G. van den Berg, Anal. Chem.2006, 78, 5052.10.1021/ac060231+Search in Google Scholar

[14] B. K. Jena, C. R. Raj, Anal. Chem.2008, 80, 4836.10.1021/ac071064wSearch in Google Scholar

[15] Z. M. Li, J. M. Liu, Z. B. Liu, Q. Y. Liu, X. Lin, F. M. Li, M. L. Yang, G. H. Zhu, X. M. Huang, Anal. Chim. Acta2007, 589, 44.10.1016/j.aca.2007.02.044Search in Google Scholar

[16] M. S. Chan, S. D. Huang, Talanta2000, 51, 373.10.1016/S0039-9140(99)00283-0Search in Google Scholar

[17] İ. Durukan, Ç. A. Şahin, S. Bektaş, Microchem. J.2011, 98, 215.10.1016/j.microc.2011.02.001Search in Google Scholar

[18] J. F. Wu, E. A. Boyle, Anal. Chem.1997, 69, 2464.10.1021/ac961204uSearch in Google Scholar PubMed

[19] Z. C. Wen, R. Yang, H. He, Y. B. Jiang, Chem. Commun.2006, 106.10.1039/B510546CSearch in Google Scholar

[20] X. Ma, Z. W. Tan, G. H. Wei, D. B. Wei, Y. G. Du, Analyst2012, 137, 1436.10.1039/c2an16155aSearch in Google Scholar PubMed

[21] C. V. Durgadas, C. P. Sharma, K. Sreenivasan, Analyst2011, 136, 933.10.1039/C0AN00424CSearch in Google Scholar

[22] J. M. Liu, L. P. Lin, X. X. Wang, S. Q. Lin, W. L. Cai, L. H. Zhang, Z. Y. Zheng, Analyst2012, 137, 2637.10.1039/c2an35130gSearch in Google Scholar

[23] Y. Q. Dong, R. X. Wang, G. L. Li, C. Q. Chen, Y. W. Chi, G. N. Chen, Anal. Chem.2012, 84, 6220.10.1021/ac3012126Search in Google Scholar

[24] X. Chen, T. Pradhan, F. Wang, J. S. Kim, J. Yoon, Chem. Rev.2011, 112, 1910.10.1021/cr200201zSearch in Google Scholar

[25] L. Fabbrizzi, M. Licchelli, P. Pallavicini, D. Sacchi, A. Taglietti, Analyst1996, 121, 1763.10.1039/an9962101763Search in Google Scholar

[26] R. Krämer, Angew. Chem. Int. Ed.1998, 37, 772.10.1002/(SICI)1521-3773(19980403)37:6<772::AID-ANIE772>3.0.CO;2-ZSearch in Google Scholar

[27] J. S. Park, S. Jeong, S. Dho, M. Lee, C. Song, Dyes Pigm.2010, 87, 49.10.1016/j.dyepig.2010.02.003Search in Google Scholar

[28] M. H. Kim, J. H. Noh, S. Kim, S. Ahn, S. K. Chang, Dyes Pigm.2009, 82, 341.10.1016/j.dyepig.2009.02.004Search in Google Scholar

[29] R. Martinez, F. Zapata, A. Caballero, A. Espinosa, A. Tarraga, P. Molina, Org. Lett.2006, 8, 3235.10.1021/ol0610791Search in Google Scholar

[30] W. C. Lin, C. Y. Wu, Z. H. Liu, C. Y. Lin, Y. P. Yen, Talanta2010, 81, 1209.10.1016/j.talanta.2010.02.012Search in Google Scholar

[31] X. Zeng, L. Dong, C. Wu, L. Mu, S. F. Xue, Z. Tao, Sens. Actuators, B2009, 141, 506.10.1016/j.snb.2009.07.013Search in Google Scholar

[32] Y. Zhou, F. Wang, Y. Kim, S. J. Kim, J. Yoon, Org. Lett.2009, 11, 4442.10.1021/ol901804nSearch in Google Scholar PubMed

[33] S. J. Lee, J. H. Jung, J. Seo, I. I. Yoon, K. M. Park, L. F. Lindoy, Org. Lett.2006, 8, 1641.10.1021/ol0602405Search in Google Scholar PubMed

[34] W. Lin, L. Yuan, X. Cao, W. Tan, Y. Feng, Eur. J. Org. Chem.2008, 2008, 4981.10.1002/ejoc.200800667Search in Google Scholar

[35] H. S. Jung, P. S. Kwon, J. W. Lee, J. I. Kim, C. S. Hong, J. W. Kim, J. Am. Chem. Soc.2009, 131, 2008.10.1021/ja808611dSearch in Google Scholar PubMed

[36] A. Helal, M. H. O. Rashid, C. H. Choi, H. S. Kim, Tetrahedron2011, 67, 2794.10.1016/j.tet.2011.01.093Search in Google Scholar

[37] Z. Xu, Y. Xiao, X. Qian, J. Cui, D. Cui, Org. Lett.2005, 7, 889.10.1021/ol0473445Search in Google Scholar PubMed

[38] J. Huang, Y. Xu, X. Qian, Dalton Trans.2009, 10, 1761.10.1039/b816999cSearch in Google Scholar PubMed

[39] J. F. Zhang, Y. Zhou, J. Yoon, Y. Kim, S. J. Kim, J. S. Kim, Org. Lett.2010, 12, 3852.10.1021/ol101535sSearch in Google Scholar PubMed

[40] N. Kaur, S. Kumar, Chem. Commun.2007, 29, 3069.10.1039/b703529bSearch in Google Scholar PubMed

[41] S. P. Wu, K. J. Du, Y. M. Sung, Dalton Trans.2010, 39, 4363.10.1039/b925898aSearch in Google Scholar PubMed

[42] E. Ballesteros, D. Moreno, T. Gomez, T. Rodriguez, J. Rojo, M. Garcia-Valverde, Org. Lett.2009, 11, 1269.10.1021/ol900050zSearch in Google Scholar PubMed

[43] Z. Q. Guo, W. Q. Chen, X. M. Duan, Org. Lett.2010, 12, 2202.10.1021/ol100381gSearch in Google Scholar PubMed

[44] P. Job, Ann. Chim.1928, 9, 113.Search in Google Scholar

[45] H. A. Benesi, J. H. Hildebrand, J. Am. Chem. Soc.1949, 71, 2703.10.1021/ja01176a030Search in Google Scholar

[46] K. A. Connors, Binding Constants, John Wiley & Sons, New York, 1987, p. 339.Search in Google Scholar

[47] B. Valeur, Molecular Fluorescence, Wiley-VCH, Weinheim, 2000, p. 339.Search in Google Scholar

[48] B. P. Joshi, J. Park, W. I. Lee, K. H. Lee, Talanta2009, 78, 903.10.1016/j.talanta.2008.12.062Search in Google Scholar PubMed

[49] Y. G. Shi, J. H. Yao, Y. L. Duan, Q. L. Mi, J. H. Chen, Q. Q. Xu, G. Z. Gou, Y. Zhou, J. F. Zhang, Bioorg. Med. Chem. Lett. 2013, 23, 2538.10.1016/j.bmcl.2013.03.004Search in Google Scholar PubMed

[50] M. S. Alexiou, V. Tychopoulos, S. Ghorbanian, J. H. P. Tyman, R. G. Brown, P. Brittain, J. Chem. Soc. Perkin Trans.1990, 2, 837.10.1039/P29900000837Search in Google Scholar

[51] Z. Chen, L. Wang, G. Zou, X. Cao, Y. Wu, P. Hu, Spectrochim. Acta, Part A2013, 114, 323.10.1016/j.saa.2013.05.072Search in Google Scholar PubMed


Supplemental Material:

The online version of this article (DOI: 10.1515/znb-2016-0138) offers supplementary material, available to authorized users.


Received: 2016-6-3
Accepted: 2016-7-6
Published Online: 2016-11-10
Published in Print: 2017-1-1

©2017 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 4.5.2024 from https://www.degruyter.com/document/doi/10.1515/znb-2016-0138/html
Scroll to top button