A FRET-based rhodamine–benzimidazole conjugate as a Cu2+-selective colorimetric and ratiometric fluorescence probe that functions as a cytoplasm marker

Shyamaprosad Goswami*a, Sibaprasad Maityab, Annada C. Maitya, Anup Kumar Maityc, Avijit Kumar Dasa and Partha Sahac
aDepartment of Chemistry, Bengal Engineering and Science University, Shibpur, Howrah 711103, West Bengal, India. E-mail: spgoswamical@yahoo.com; Fax: +91-3326682916
bDepartment of Applied Sciences, Haldia Institute of Technology, Hatiberia, Haldia, West Bengal-721657, India. E-mail: spmaity2003@gmail.com
cCrystallography and Molecular Biology Division, Saha Institute of Nuclear Physics, Kolkata 700064, West Bengal, India. E-mail: partha.saha@saha.ac.in

Received 31st October 2013 , Accepted 18th November 2013

First published on 19th November 2013


Abstract

On the basis of fluorescence resonance energy transfer (FRET) from benzimidazole to a rhodamine moiety, a rhodamine–benzimidazole conjugate (RBC) ratiometric fluorescent probe has been designed and synthesized. The RBC selectively binds to Cu2+, showing visually observable changes in absorption and emission behavior, and demonstrates an effective intracellular Cu2+ imaging ability, allowing it to function as a cytoplasm marker.


Introduction

Fluorescent chemosensors which display a significant change in optical signal when they selectively sense a specific guest analyte, especially for metal ions of biological interest, have several advantages over other methods such as their sensitivity, specificity and real-time monitoring with fast response times.1,2 However, variation in the fluorescence intensity without much shift in either the excitation or emission wavelengths, can be influenced by many environmental aspects, the emission collection efficiency, the effective cell thickness in the optical beam and changes in the excitation intensity.3 To reduce the influence of such factors, ratiometric measurement is utilized, where simultaneous recording of the fluorescence intensities at two wavelengths and computation of their ratio is carried out.4 This technique provides greater precision than measurement at a single wavelength, and is suitable for cellular imaging studies.

The design of fluorescent probes is generally based on intra-molecular charge transfer (ICT), photo-induced electron transfer (PET), chelation-enhanced fluorescence (CHEF), metal–ligand charge transfer (MLCT), excimer/exciplex formation, imine isomerization, intermolecular hydrogen bonding, excited-state intra-molecular proton transfer, and fluorescence resonance energy transfer (FRET).5–12 FRET, which is our present interest, is defined as an excited-state energy interaction between two fluorophores in which the excited donor energy is transferred non-radiatively to an acceptor unit. Among these mechanisms using a single fluorophore to obtain ratiometric changes, the use of multi-fluorophores with energy donor–acceptor architectures can achieve large pseudo-Stokes shifts. This approach can provide simultaneous recording ratio signals for two emission intensities at different wavelengths, which could afford a built-in correction for environmental effects13 and supply a facile method for DNA detection, the labeling of proteins and other biomarkers, and for visualizing complex biological processes at the molecular level, such as nucleic acid regulation.14 Thus, FRET is one of the most commonly used chemical principles for ratiometric imaging, whether in vitro or in vivo.

Copper(II) plays a significant role in various fields.15,16 However, exposure to high concentration levels of copper, even for a short period of time, can cause gastrointestinal disturbances, while long-term exposure can cause liver or kidney damage.17–20 Furthermore, Cu2+ can act as a significant environmental pollutant because of its extensive use in industry and agriculture. Thus, the fast detection of Cu2+ in environmental and biological samples has become increasingly critical, not only due to its important role in biological processes, but also, because of its high toxicity to organisms at increased concentration levels.21

Considerable efforts have been made to synthesize fluorescent chemosensors that are selective, sensitive and suited to high-resolution imaging for monitoring biological processes.22 Even though great achievements in the field of colorimetric and/or fluorescent chemosensors for Cu2+ have been reported,23–31 there is still a demand to develop new indicators with improved properties. Due to the significant physiological relevance and associated biomedical implications, the design and development of selective and sensitive sensors directed toward the detection and measurement of divalent Cu2+, the third most abundant essential trace element after iron and zinc in the human body, is highly desirable. There are only a few reports on FRET-based ion chemosensors.32,33 In particular, the design of ratiometric probes for Cu2+ ions is a challenge due to their inherent paramagnetic nature, and hence, complexation generally results in quenching of the fluorescence intensity of the probe.

Among the fluorophores developed, rhodamine is highly favored because of its photophysical properties, which include a high extinction coefficient, a high quantum yield and remarkable photostability. In general, spirolactam ring opening of a rhodamine derivative on complexation with a metal ion gives rise to a color change and strong fluorescence.34–39

Motivated by the biological importance of Cu2+, herein we report the ratiometric metal ion sensing capability of a benzimidazole–rhodamine conjugate (RBC) and its absorption and emission activities upon metal complexation for detecting Cu2+ and its effective bioimaging. In the present study, our design strategy for the detection of Cu2+ is based on modulating the FRET process in a fluorophore dyad comprising the rhodamine acceptor and the benzimidazole derivative as a donor, linked by a multi-chelating site.

Results and discussion

The fluorescent RBC probe was obtained from the reaction of rhodamine B hydrazide40 (1) with 2-chloromethyl-1-methyl-1H-benzimidazole (2) in refluxing THF for two days as shown in Scheme 1. The molecular structure of the RBC was confirmed using 1H NMR and 13C NMR spectroscopy, and HRMS (ESI).
image file: c3ra46280c-s1.tif
Scheme 1 Synthesis of the RBC.

As the RBC bears two different fluorophore units, we considered it appropriate to study the metal binding event at two different excitation wavelengths corresponding to the benzimidazole unit (315 nm) and the xanthene unit (495 nm). In the absence of Cu2+, the rhodamine moiety adopts a closed, non-fluorescent spirolactam form, corresponding to a weak spectral overlap between benzimidazole emission and rhodamine absorption.

The binding of Cu2+ induces opening of the spirolactam ring in the RBC with an associated switch to a UV-vis spectral response in the 500–580 nm range, which has a significant spectral overlap with the emission spectrum of the benzimidazole fragment (shaded black area in Fig. 1), which could provide a plausible route for the non-radiative transfer of excitation energy from the donor benzimidazole to the acceptor rhodamine moiety, thereby initiating an intramolecular FRET process.


image file: c3ra46280c-f1.tif
Fig. 1 Spectral overlap between benzimidazolyl (10 μM) emission (blue line) and ring-opened rhodamine B (10 μM) absorption (pink line).

The photophysical properties of the RBC were investigated in an acetonitrile–HEPES buffer solution (1 mM, pH 7.4; 8[thin space (1/6-em)]:[thin space (1/6-em)]2 v/v). As expected, the RBC barely displayed absorbance at 450–650 nm, indicating that the RBC exists in a spiro-cycle closed form, and instead, shows an absorption band centered at 315 nm. The gradual addition of Cu2+ to a solution of the RBC caused a significant enhancement in absorbance (Fig. 2) located at 556 nm, while the absorbance at 315 nm diminished slowly.


image file: c3ra46280c-f2.tif
Fig. 2 UV-vis titration spectra of the RBC (10 μM) with varying [Cu2+] from 0 to 3 equiv in a CH3CN–HEPES buffer solution (1 mM, pH 7.4; 8[thin space (1/6-em)]:[thin space (1/6-em)]2 v/v).

These results indicate that chelation with Cu2+ induced the development of a pink color as a result of ring opening of the spirolactam form of the RBC.

In the presence of competing metal ions such as Mn2+,Cr3+, Pb2+, Fe2+, Co2+, Ni2+, Zn2+, Cd2+ and Hg2+ (as perchlorate, nitrate or chloride salts), in turn, no significant absorbance (Fig. 3) was observed at 556 nm that could interfere with the selectivity of the RBC for the detection of Cu2+. These results show that the RBC is a Cu2+-specific probe that allows the visual detection of Cu2+.


image file: c3ra46280c-f3.tif
Fig. 3 Changes in the absorption spectra of the RBC–Cu2+ complex in the presence of different metal ions.

The selectivity of the RBC towards Cu2+ binding was also observed through fluorescence titration. The RBC was excited at 315 nm, which is the excitation wavelength of the benzimidazole moiety. In the absence of Cu2+, the fluorescence emission peak appeared at 490 nm. Upon the gradual addition of Cu2+, a significant decrease in the emission intensity at 490 nm and a new fluorescence emission band centered at 580 nm were observed, with a clear iso-emissive point at 553 nm (Fig. 4), which resulted in an intense reddish orange color.


image file: c3ra46280c-f4.tif
Fig. 4 Fluorescence titration spectra (λex = 315 nm) of the RBC (10 μM) upon the incremental addition of 0 to 3 equiv of Cu2+, in an CH3CN–HEPES buffer solution (1 mM, pH 7.4; 8[thin space (1/6-em)]:[thin space (1/6-em)]2 v/v).

This is due to the fact that binding of Cu2+ with the RBC induces spirolactam ring opening of the rhodamine derivative, whose absorption spectrum shows a significant spectral overlap with the emission spectrum of the benzimidazole moiety. This facilitates the resonance energy transfer process (Scheme 2), and hence, a pseudo-large Stokes shift based on the FRET mechanism is observed . In the presence of Cu2+, the ratio of the emission intensities for rhodamine to the benzimidazole moiety at 580 nm and 490 nm (I580/I490) increases substantially from 0.28 in the absence of Cu2+ to 15.1 in the presence of Cu2+ (up to 3 equiv.).


image file: c3ra46280c-s2.tif
Scheme 2 Cu2+-induced FRET OFF → ON of the RBC.

A competition experiment was also performed by adding Cu2+ (3.0 equiv.) to a RBC solution in the presence of commonly employed interfering metal ions (3.0 equiv.). The selectivity profile diagram (Fig. 6) reveals that Cu2+-induced fluorescence enhancement (I580/I490) remains unaffected by the co-existence of other metal ions, and interference does not occur.

Thus, the RBC exhibits excellent selectivity for Cu2+ and is insensitive to interference from other metal ions (Fig. 5). Only Fe2+ and Co2+ showed some emission at around 580, but they could not compete with Cu2+. Fluorescence titration and Job plot analysis (Fig. S1) confirmed a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 stoichiometry for Cu2+ and the RBC, with an association constant of 1.5 × 104 M−1 (Fig. S2) and a detection limit of 3.1 μM (Fig. S3).


image file: c3ra46280c-f5.tif
Fig. 5 Changes in the fluorescence emission of the RBC (10 μM) observed upon addition of different metal ions in a CH3CN–HEPES buffer solution (1 mM, pH 7.4; 8[thin space (1/6-em)]:[thin space (1/6-em)]2 v/v).

The mass spectrum (ESI) of the RBC–Cu2+ complex (expected M+) shows a molecular-ion peak at m/z 425.27 [C48H51CuN9O2 + 2H]2+ = 425 corresponding to [RBC + Cu + CH3CN]2+(ESI) [C48H51CuN9O2 + 2H]2+ = 425, which suggests a probable mode of binding, as proposed in Fig. 7.


image file: c3ra46280c-f6.tif
Fig. 6 Metal ion selectivity profile of the RBC sensor (10 μM): change in emission intensity of the RBC + 3.0 equiv. Mn+ (black bars); change in emission intensity of the RBC + 3.0 equiv. Mn+, followed by 3.0 equiv. Cu2+ (pink bars) at I580/I490.

image file: c3ra46280c-f7.tif
Fig. 7 Probable mode of binding of the RBC–Cu2+ complex in acetonitrile.

Cell imaging

Due to the favorable binding properties of the RBC with respect to Cu2+, and its intense emission in the visible region, it was conceived that it could be exploited for practical bioimaging, particularly for the sensitive detection of intracellular Cu2+. In order to determine the membrane permeability of the RBC receptor and its ability to specifically bind Cu2+ ions in living cells, HeLa cells were first incubated with CuCl2 followed by the addition of the RBC. A control experiment, i.e., incubation with the RBC only, was also carried out. As shown in Fig. 8, the cytoplasm of the cells showed intense red fluorescence in the red channel when they were treated with CuCl2 followed by the RBC, but no fluorescence was observed in cells that were treated with the RBC only in the same channel. As expected, a strong blue fluorescence could be obtained from the nucleus due to DAPI treatment.
image file: c3ra46280c-f8.tif
Fig. 8 Fluorescence images of HeLa cells incubated with 50 μM of the RBC in the presence (a and b) and in the absence (e and f) of 50 μM of CuCl2. The corresponding bright field images (c and g) and merge images (d and h) of the cells are shown.

The results clearly indicate that the RBC receptor not only permeates the plasma membrane of the cells, but also brings about a specific cytoplasmic fluorescence in the presence of Cu2+ ions. Therefore, the RBC receptor could act as a good cytoplasmic marker in the presence of Cu2+ ions.

Conclusions

In summary, we have developed a sensitive RBC probe which selectively binds to Cu2+ ions and induces a switch ON response in fluorescence spectra in the visible region. In addition to this, the FRET-based fluorescence response makes it a dual probe for naked eye detection through changes in color and fluorescence. The detection limit for Cu2+ was found to be much lower than the standard permissible Cu2+ concentrations in drinking water. Furthermore, the RBC demonstrates good bioimaging ability as a cytoplasmic marker in the presence of intracellular Cu2+.

Experimental section

General

Unless otherwise mentioned, chemicals and solvents were purchased from Sigma-Aldrich, and were used without further purification. 1H-NMR spectra were recorded on a Bruker 400 MHz instrument. For NMR spectra, CDCl3 was used as a solvent with TMS as an internal standard. Chemical shifts are expressed in δ units and as 1H–1H and 1H–C coupling constants in Hz. UV-vis titration experiments were performed on a JASCO UV-V530 spectrophotometer and fluorescence experimentation was carried out using a PerkinElmer LS 55 fluorescence spectrophotometer using a fluorescence cell of 10 mm path.

General method for UV-vis and fluorescence titrations

For UV-vis and fluorescence titrations, a stock solution of the RBC was prepared (c = 1 × 10−5 ML−1) in an acetonitrile–HEPES buffer solution (1 mM, pH 7.4; 8[thin space (1/6-em)]:[thin space (1/6-em)]2 v/v). Solutions of the guest metal ions, including Cr3+, Ni2+, Cu2+, Pb2+, Co2+, Mn2+, Cd2+, Fe2+, Zn2+, and Hg2+ (chloride salts), were prepared (c = 2 × 10−4 ML−1) in CH3CN. The original volume of the RBC solution was 2 ml. RBC solutions of various concentrations, and increasing concentrations of the metal ions, were prepared separately. The absorption and fluorescence sensing of the metal ions were then recorded.

Cell culture and imaging

HeLa cells were grown in Dulbecco's modified Eagle's medium (DMEM) supplemented with 10% fetal bovine serum and penicillin–streptomycin (0.5 U/ml of penicillin and 0.5 μg ml−1 streptomycin), on a cover slip in 35 mm dishes at 37 °C, in an atmosphere of air with 5% CO2 and constant humidity. The cells were initially incubated with the addition of 50 μM of CuCl2 in the growth medium for 30 minutes. After washing three times with phosphate-buffered saline (PBS), fresh growth medium containing 50 μM of the of RBC was added and the cells were incubated for another 30 minutes. Following incubation, the cells were washed three times with PBS and cross-linked with 4% HCHO. Finally, the cover slip was placed on a glass slide containing a DAPI solution and imaging was carried out using a Zeiss Axio Observer Fluorescence Microscope equipped with an ApoTome apparatus.

Synthesis of the RBC

2-Amino-3′,6′-bis(diethylamino)spiro[isoindoline-1,9′-xanthen]-3-one (1) (500 mg, 1.1 mmol) and 2-chloromethyl-1-methyl-1H-benzimidazole (2) were refluxed in dry THF in the presence of Et3N for two days. The mixture was extracted with chloroform. The combined organic layer was dried over anhydrous magnesium sulfate and then filtered. The filtrate was concentrated and purified through column chromatography using 10% CH3OH in CHCl3, to give a 72% yield.

1H NMR (CDCl3, 400 MHz): δ (ppm): 7.93 (s, 2H), 7.41 (s, 4H), 7.08 (s, 2H), 6.45 (d, 7H, J = 8.00 Hz), 6.28 (d, 3H, J = 8.00 Hz), 3.62 (s, 4H), 3.32 (s, 14H), 1.16 (s, 12H).

13C NMR (CDCl3, 100 MHz): δ (ppm): 166.05, 153.84, 151.58, 148.84, 132.48, 130.03, 128.04, 123.50, 122.90, 108.01, 104.63, 98.00, 65.86, 44.35, 12.62.

HRMS (M + H+): calcd 745.2978, found 745.2430. Anal calcd for C46H48N8O2: 74.17% C, 6.49% H, 15.05% N, 4.30% O; found: 74.26% C, 6.90% H, 15.18% N, 4.68% O.

Acknowledgements

The authors thank the DST and CSIR (Govt. of India) for financial support.

Notes and references

  1. C.-Y. Lai, B. G. Trewyn, D. M. Jeftinija, K. Jeftinija, S. Xu, S. Jeftinija and V. S.-Y. Lin, J. Am. Chem. Soc., 2003, 125, 4451 CrossRef CAS PubMed.
  2. M. Numata, C. Li, A.-H. Bae, K. Kaneko, K. Sakurai and S. Shinkai, Chem. Commun., 2005, 4655 RSC.
  3. K. Kikuchi, Chem. Soc. Rev., 2010, 39, 2048 RSC.
  4. R. Y. Tsien and A. T. Harootunian, Cell Calcium, 1990, 11, 93 CrossRef CAS.
  5. J. B. Wang, X. F. Qian and J. N. Cui, J. Org. Chem., 2006, 71, 4308 CrossRef CAS PubMed.
  6. S. K. Kim and J. Yoon, Chem. Commun., 2002, 770 RSC.
  7. A. Sahana, A. Banerjee, S. Lohar, S. Panja, S. K. Mukhopadhyay, J. S. Matalobos and D. Das, Chem. Commun., 2013, 49, 7231 RSC.
  8. B. Schazmann, N. Alhashimy and D. Diamond, J. Am. Chem. Soc., 2006, 128, 8607 CrossRef CAS PubMed.
  9. J.-S. Wu, W.-M. Liu, X.-Q. Zhuang, F. Wang, P.-F. Wang, S.-L. Tao, X.-H. Zhang, S.-K. Wu and S.-T. Lee, Org. Lett., 2007, 9, 33 CrossRef CAS PubMed.
  10. A. Sahana, A. Banerjee, S. Guha, S. Lohar, A. Chattopadhyay, S. K. Mukhopadhyay and D. Das, Analyst, 2012, 137, 1544 RSC.
  11. X. Peng, Y. Wu, J. Fan, M. Tian and K. Han, J. Org. Chem., 2005, 70, 10524 CrossRef CAS PubMed.
  12. J. M. Serin, D. W. Brousmiche and J. M. J. Frechet, J. Am. Chem. Soc., 2002, 124, 1184 CrossRef PubMed.
  13. G. Chen, F. Song, J. Wang, Z. Yang, S. Sun, J. Fan, X. Qiang, X. Wang, B. Doub and X. Peng, Chem. Commun., 2012, 48, 2949 RSC.
  14. J. Fan, M. Hu, P. Zhan and X. Peng, Chem. Soc. Rev., 2013, 42, 29 RSC.
  15. K. Rurack, M. Kollmannsberger, U. Resch-Genger and J. Daub, J. Am. Chem. Soc., 2000, 122, 968 CrossRef CAS.
  16. S. Y. Moon, N. R. Cha, Y. H. Kim and S.-K. Chang, J. Org. Chem., 2004, 69, 181 CrossRef CAS PubMed.
  17. D. G. Barceloux and D. Barceloux, Clin. Toxicol., 1999, 37, 217 CrossRef CAS PubMed.
  18. X. B. Zhang, J. Peng, C.-L. He, G.-L. Shen and R.-Q. Yu, Anal. Chim. Acta, 2006, 567, 189 CrossRef CAS PubMed.
  19. B. Sarkar, in Metal Ions in Biological Systems, ed. H. Siegel and A. Siegel, Marcel Dekker, New York, 1981, 12, 233 Search PubMed.
  20. E. L. Que, D. W. Domaille and C. J. Chang, Chem. Rev., 2008, 108, 1517 CrossRef CAS PubMed.
  21. W. Wang, A. Fu, J. You, G. Gao, J. Lan and L. Chen, Tetrahedron, 2010, 66, 3695 CrossRef CAS PubMed.
  22. H. S. Jung, P. S. Kwon, J. W. Lee, J.-Il. Kim, C. S. Hong, J. W. Kim, S. Yan, J. Y. Lee, J. H. Lee, T. Joo and J. S. Kim, J. Am. Chem. Soc., 2009, 131, 2008 CrossRef CAS PubMed.
  23. J. Huang, Y. Xu and X. Qian, Org. Biomol. Chem., 2009, 7, 1299 CAS.
  24. Y. Zhou, F. Wang, Y. Kim, S.-J. Kim and J. Yoon, Org. Lett., 2009, 11, 4442 CrossRef CAS PubMed.
  25. G. Sivaraman, T. Anand and D. Chellappa, RSC Adv., 2013, 3, 17029 RSC.
  26. N. R. Chereddy, P. S. Korrapati, S. Thennarasu and A. B. Mandal., Dalton Trans., 2013, 42, 12873 RSC.
  27. Y. Zhao, X.-B. Zhang, Z.-X. Han, L. Qiao, C.-Y. Li, L.-X. Jian, G.-L. Shen and R.-Q. Yu, Anal. Chem., 2009, 81, 7022 CrossRef CAS PubMed.
  28. Z. Guo, W. Zhu and H. Tian, Macromolecules, 2010, 43, 739 CrossRef CAS.
  29. Z. Xu, J. Pan, D. R. Spring, J. Cui and J. Yoon, Tetrahedron, 2010, 66, 1678 CrossRef CAS PubMed.
  30. S. Goswami, D. Sen, A. K. Das, N. K. Das, K. Aich, H.-K. Fun, C. K. Quah, A. K. Maity and P. Saha, Sens. Actuators, B, 2013, 183, 518 CrossRef CAS PubMed.
  31. S. Goswami, D. Sen and N. K. Das, Org. Lett., 2010, 12, 856 CrossRef CAS PubMed.
  32. M. H. Lee, H. J. Kim, S. Yoon, N. Park and J. S. Kim, Org. Lett., 2008, 10, 213 CrossRef CAS PubMed.
  33. A. B. Othman, J. W. Lee, J. S. Wu, J. S. Kim, R. Abidi, P. J. Thuéry, M. Strub, A. V. Dorsselaer and J. Vicens, J. Org. Chem., 2007, 72, 7634 CrossRef PubMed.
  34. X. Chen, T. Pradhan, F. Wang, J. S. Kim and J. Yoon, Chem. Rev., 2012, 112, 1910 CrossRef CAS PubMed.
  35. G. Sivaraman and D. Chellappa, J. Mater. Chem. B, 2013, 1, 5768 RSC.
  36. G. Sivaraman, T. Anand and D. Chellappa, Analyst, 2012, 137, 5881 RSC.
  37. G. Sivaraman, V. Sathiyaraja and D. Chellappa, J. Lumin., 2014, 145, 480 CrossRef CAS PubMed.
  38. H. N. Kim, M. H. Lee, H. J. Kim, J. S. Kim and J. Yoon, Chem. Soc. Rev., 2008, 37, 1465 RSC.
  39. V. Luxami, R. K. Paul and S. Kumar., RSC Adv., 2013, 3, 9189 RSC.
  40. W. Y. Liu, H. Y. Li, B. X. Zhao and J. Y. Miao, Org. Biomol. Chem., 2011, 9, 4802 CAS.

Footnote

Electronic supplementary information (ESI) available: Details of synthetic procedure and spectral data available. See DOI: 10.1039/c3ra46280c

This journal is © The Royal Society of Chemistry 2014