Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

A Highly Selective and Non-Reaction Based Chemosensor for the Detection of Hg2+ Ions Using a Luminescent Iridium(III) Complex

  • Daniel Shiu-Hin Chan ,

    Contributed equally to this work with: Daniel Shiu-Hin Chan, Wai-Chung Fu

    Affiliation Department of Chemistry, Hong Kong Baptist University, Kowloon Tong, Hong Kong, China

  • Wai-Chung Fu ,

    Contributed equally to this work with: Daniel Shiu-Hin Chan, Wai-Chung Fu

    Affiliation Department of Chemistry, Hong Kong Baptist University, Kowloon Tong, Hong Kong, China

  • Modi Wang,

    Affiliation Department of Chemistry, Hong Kong Baptist University, Kowloon Tong, Hong Kong, China

  • Li-Juan Liu,

    Affiliation State Key Laboratory of Quality Research in Chinese Medicine, Institute of Chinese Medical Sciences, University of Macau, Macao, China

  • Chung-Hang Leung,

    Affiliation State Key Laboratory of Quality Research in Chinese Medicine, Institute of Chinese Medical Sciences, University of Macau, Macao, China

  • Dik-Lung Ma

    edmondma@hkbu.edu.hk

    Affiliation Department of Chemistry, Hong Kong Baptist University, Kowloon Tong, Hong Kong, China

Abstract

We report herein a novel luminescent iridium(III) complex with two hydrophobic carbon chains as a non-reaction based chemosensor for the detection of Hg2+ ions in aqueous solution (<0.002% of organic solvent attributed to the probe solution). Upon the addition of Hg2+ ions, the emission intensity of the complex was significantly enhanced and this change could be monitored by the naked eye under UV irradiation. The iridium(III) complex shows high specificity for Hg2+ ions over eighteen other cations. The system is capable of detecting micromolar levels of Hg2+ ions, which is within the range of many chemical systems.

Introduction

Mercury is a highly toxic heavy metal ion that is harmful to both humans and the environment. Metabolism by marine microorganisms converts mercury ions into methylmercury, a highly toxic and bio-accumulative form [1] that damages the human central nervous and endocrine systems and is associated with sensory, motor and cognitive disorders [2]. Evidence has also suggested that exposure to high levels of mercury ions can damage the lungs and kidneys [3]. Therefore, the development of new methods for the selective detection of mercury ions is of particular importance and remains an active area of research in the scientific community.

Traditional instrumental techniques for detection of Hg2+ ions include atomic absorption/emission spectrometry (AAS/AES) [4], [5], inductively-coupled plasma mass spectrometry or atomic emission spectroscopy (ICP-MS/ICP-AES) [6][8] and X-ray fluorescence (XRF) [9][11]. Despite their widespread usage in industry and the laboratory, these methods are time-consuming and require extensive pre-treatment procedures, and involve the use of complex and expensive instrumentation. Over the past decade, a number of alternative methods for the detection of metal ions have been reported, including luminescent chemosensors [12][19], electrochemical sensors [20], [21] and colorimetric probes [22][24]. However, most luminescent probes for Hg2+ ions only perform well in organic solvents [13], [17], [18], which is not favourable for real sample analysis. Therefore, it is desirable to develop water-soluble luminescent probes for Hg2+ ions that can function effectively in aqueous solution.

Luminescent transition metal complexes have attracted considerable attention in the fabrication of organic optoelectronics [25], [26], luminescent sensors [27][32] and cellular imaging [33][41] by virtue of their salient advantages: (i) the 3MLCT emission of many metal complexes lie in the visible spectral region, (ii) their long-lived phosphorescence emission can be resolved from a fluorescent background by time-resolved spectroscopic techniques, thus enhancing signal imaging stability, (iii) the significant stokes shifts of the complexes allow for easy separation of their excitation and emission wavelengths, thus preventing self-quenching, (iv) their facile colour-tuning ability makes them suitable for different photophysical applications [42][49], and (v) the preparation of metal complexes is highly modular.

While luminescent iridium(III) complexes have been successfully applied in a variety of fields, there are few reports on luminescent iridium(III)-based chemosensors for the detection of Hg2+ ions. Li, Huang and co-workers reported an iridium(III) complex as a chemodosimeter of Hg2+ ions based on the interaction between Hg2+ and the sulfur atom of the cyclometalated ligands [50][52]. Lu and co-workers fabricated a chemosensor for Hg2+ ions based on the dissociation of a dithiocarbamate ligand from the iridium(III) complex [53]. However, these reaction-based iridium(III) chemosensors are strictly dependent upon the quantitative interaction between the metal complex ligands and Hg2+ ions. In this work, we report the application of a novel cyclometalated iridium(III) complex [Ir(dfppy)2(dnbpy)]+ (1, where dfppy = 2,4-difluorophenylpyridine and dnbpy = 4,4′-dinonyl-2,2′-bipyridine) (Figure 1) as a non-reaction based switch-on chemosensor for Hg2+ ions in aqueous solution.

thumbnail
Figure 1. Chemical structure of the iridium(III) complex 1 bearing the 4,4’-dinonyl-2,2’-bipyridine ligand.

https://doi.org/10.1371/journal.pone.0060114.g001

Results and Discussion

The photophysical properties of complex 1 are provided in Table S1, Figure S1 and S2. Complex 1 displays a strong absorption band between 250 and 300 nm which can be attributed to spin-allowed π-π* ligand-centered (1LC) transitions of the dfppy ligand [54]. The absorption band at 303 nm is assigned to an iridium-based spin allowed metal-to-ligand charge transfer (1MLCT) transition, while the low-energy absorption shoulder at approximately 450 nm is assigned to spin-forbidden triplet 3MLCT transitions according to the peak assignment of a similar iridium(III) complex [55]. The emission peak at λ = 490 nm is phosphorescent in nature as revealed by its relatively long emission lifetime (4.53 µs).

The luminescence response of 1 upon addition of different concentrations of Hg2+ ions was first investigated by emission titration experiments. Complex 1 was weakly emissive in aqueous buffered solution. However, the luminescence of 1 was significantly enhanced in the presence of the increasing concentrations of Hg2+ ions. We presumed that the unusual sensing behaviour of the complex towards Hg2+ may be due to the presence of its hydrophobic side chains, which are known to have a tendency to adsorb Hg2+ ions [56]. The possible sensing mechanism of this Hg2+ chemosensor is depicted in Figure 2. Mercury ions may interact with the hydrophobic carbon chains of multiple complexes, inducing aggregation of the iridium(III) complexes into a micelle-like motif. This results in a strong enhancement of the luminescence emission of 1 at λ = 490 nm, presumably due to the partial protection of the complex from non-radiative decay by solvent quenching, thus giving rise to an enhanced 3MLCT emission.

thumbnail
Figure 2. Schematic illustration of the non-reaction based assay for the detection of Hg2+ ions using luminescent iridium(III) complex 1.

The adsorption of Hg2+ ions on the hydrophobic side chains induces aggregation of complex 1. This partially shields the complexes from solvent interactions, hence resulting in an enhanced emission intensity at λ = 490 nm.

https://doi.org/10.1371/journal.pone.0060114.g002

To optimize the performance of the system, we investigated the effect of the concentration of complex 1 on the luminescence response of the system to Hg2+ ions. The results showed that a concentration of 0.5 µM of complex 1 was optimal for this study, as the sensitivity and dynamic range of detection of the system were reduced at higher concentrations (Figure S3). Under the optimal conditions, we investigated the application of the proposed system for the detection of Hg2+ ions in aqueous buffered solution (25 mM Tris, pH 7.0). As shown in Figure 3, a strong increase in the emission intensity of 1 was observed upon addition of Hg2+ ions, with a maximum luminescence enhancement (I/I0 –1) of ca. 1.4 at saturating concentrations of [Hg2+]. A linear relationship was observed between the luminescence intensity of 1 and the Hg2+ concentration (R2 = 0.96) in the range of 0–10 µM of Hg2+ (Figure 3). The detection limit at a signal to noise ratio of 3 was found to be 2.8 µM, which is sufficient for the detection of Hg2+ ions in many chemical systems. The luminescence enhancement of the system upon the addition of micromolar Hg2+ ions can be readily observed by the naked eye under UV-irradiation (Figure 3). These results indicate that with a portable spectrophotometer, complex 1 could possibly be used in field studies as a sensitive “naked-eye” indicator for Hg2+ ions in water samples.

thumbnail
Figure 3. (Upper panel) Emission spectrum of complex 1 (0.5 µM) in the presence of increasing concentrations of Hg2+ ions (0, 1, 2, 5, 10, 15, 20, 25, 30 µM).

(Middle Panel) Luminescence response of the system at λ = 490 nm vs. Hg2+ concentration. Inset: linear plot of the change in luminescence intensity at λ = 490 nm vs. Hg2+ concentration. Error bars represent the standard deviations of the results from three independent experiments. (Lower Panel) Photograph image of 1 (0.5 µM) in Tris buffer (25 mM, pH 7.0) in the absence (left) or presence (right) of 30 µM of Hg2+ ions under UV irradiation.

https://doi.org/10.1371/journal.pone.0060114.g003

Thiol-containing compounds can effectively sequester Hg2+ ions by the formation of the strong Hg(II)–S bond, and this fact has been utilized in the fabrication of assays for detection of both bio-thiols and Hg2+ ions [57], [58]. To validate our hypothesis that the enhanced luminescence of 1 is due to the direct interaction between the metal complex and Hg2+ ions, we investigated the effect of adding cysteine to a solution of 1 and Hg2+ ions (Figure S4). The results showed that the emission intensity of 1 was significantly decreased upon the addition of cysteine, which could be attributed to the extraction of Hg2+ ions by the strong Hg(II)–S interaction and the subsequent dissociation of the metal complex aggregate. The interaction between 1 and Hg2+ ions was further examined by 1H NMR titration experiments in CD3CN solution (Figure S5). The aromatic protons of complex 1 were not significantly perturbed upon the addition of Hg2+ ions, indicating the absence of ligand replacement or covalent binding between Hg2+ ions and metal complex 1, which is unlike the previously reported iridium(III) Hg2+ chemodosimeters reported [51], [53].

The specific response of the system to Hg2+ ions was evaluated by examining the luminescence signal of complex 1 in the presence of various metal ions under the optimal conditions. As shown in Figure 4, only the addition of Hg2+ could induce a prominent increase in the luminescence emission of 1, whereas the addition of 10-fold of eighteen other cations (Li+, Na+, Mg2+, Al3+, K+, Ca2+, Ti3+, Cr3+, Fe3+, Co3+, Ni3+, Cu2+, Zn2+, Sr2+, Ag+, Cd2+, La3+, Pb2+) caused only very slight luminescence changes. The slight decrease in luminescence intensity upon the addition of 10-fold excess of certain metal ions may be presumably attributed to the disruption of pre-aggregation of 1 by those cations [54].

thumbnail
Figure 4. Luminescence response of complex 1 (0.5 µ µM) in the presence of Hg2+ (5 µM) or 10-fold excess of various metal ions (Li+, Na+, Mg2+, Al3+, K+, Ca2+, Ti3+, Cr3+, Fe3+, Co3+, Ni3+, Cu2+, Zn2+, Sr2+, Ag+, Cd2+, La3+, Pb2+) in Tris buffer (25 mM, pH 7.0).

Error bars represent the standard deviations of the results from three independent experiments.

https://doi.org/10.1371/journal.pone.0060114.g004

A competition study on the selectivity of 1 towards Hg2+ was also conducted to investigate the performance of the assay in the presence of interfering metal ions. The luminescence signal of the probe was slightly decreased upon addition of a mixture of five interfering metal ions (150 µM each of Pb2+, Fe3+, Co2+, La3+, Ti3+) (Figure S6). However, the subsequent addition Hg2+ (30 µM) strongly promotes the aggregation of 1 due to the specific binding of Hg2+ ions to the hydrophobic side chains of 1, thereby enhancing its luminescence emission. This result demonstrates that the Hg2+ detection assay is able to function effectively even in the presence of multiple interfering metal ions at excess.

Conclusions

In summary, we have synthesized and characterized a novel non-reaction based luminescent iridium(III) complex 1 for the rapid, selective and direct detection of Hg2+ in aqueous solution. This chemosensor displays a strong luminescence “switch-on” response to Hg2+ with a detection limit in low-micromolar range, which is comparable to existing iridium(III)-based Hg2+ chemosensors, and is highly selective for Hg2+ over eighteen other metal ions. Furthermore, the addition of cysteine to the system can revert the luminescence signal of 1 to the “off” state. We envisage this luminescent iridium(III) complex could be further developed as a reusable Hg2+ chemosensor for the sensitive detection of Hg2+ in aqueous solution.

Materials and Methods

Chemicals and materials

Reagents were purchased from Sigma Aldrich and used as received. Iridium chloride hydrate (IrCl3.xH2O) was purchased from Precious Metals Online.

General experimental

Mass spectrometry was performed at the Mass Spectroscopy Unit at the Department of Chemistry, Hong Kong Baptist University, Hong Kong (China). Melting points were determined using a Gallenkamp melting apparatus and are uncorrected. Deuterated solvents for NMR purposes were obtained from Armar and used as received.

1H and 13C NMR were recorded on a Bruker Avance 400 spectrometer operating at 400 MHz (1H) and 100 MHz (13C). 1H and 13C chemical shifts were referenced internally to solvent shift (CD3CN: 1H, δ1.94, 13C, δ118.7; d6-DMSO: 1H, δ2.50, 13C δ39.5). Chemical shifts (δ) are quoted in ppm, the downfield direction being defined as positive. Uncertainties in chemical shifts are typically ±0.01 ppm for 1H and ±0.05 for 13C. Coupling constants are typically±0.1 Hz for 1H-1H and ±0.5 Hz for 1H-13C couplings. The following abbreviations are used for convenience in reporting the multiplicity of NMR resonances: s, singlet; d, doublet; t, triplet; q, quartet; m, multiplet; br, broad. All NMR data was acquired and processed using standard Bruker software (Topspin).

Absorption spectra were recorded on a Cary 300 UV/Vis spectrometer. Emission spectra were recorded on a PTI QM4 spectrometer. Quantum yields and lifetime measurements were performed on a PTI TimeMaster C720 Spectrometer (Nitrogen laser: pulse output 337 nm) fitted with a 380 nm filter. Error limits were estimated: λ (±1 nm); τ (±10%); φ (±10%). All solvents used for the quantum yield and lifetime measurements were degassed using three cycles of Freeze-Vacuum-Thaw.

Synthesis of [Ir(dfppy)2(dnbpy)]PF6

[Ir(dfppy)2(dnbpy)]PF6.

A suspension of [Ir2(dfppy)4Cl2] [59] (120 mg, 0.1 mmol) and 4,4'-dinonyl-2,2'-bipyridine (89.8 mg, 0.22 mmol) in a mixture of dichloromethane:methanol (1:1, 20 mL) was refluxed overnight under a nitrogen atmosphere. The resulting solution was then allowed to cool to room temperature, and filtered to remove unreacted cyclometalated dimer. To the filtrate, an aqueous solution of ammonium hexafluorophosphate (excess) was added and the filtrate was reduced in volume by rotary evaporation until precipitation of the crude product occurred. The precipitate was then filtered and washed with several portions of water (2×50 mL) followed by diethyl ether (2×50 mL). The product was recrystallized by acetonitrile:diethyl ether vapor diffusion to yield the titled compound as a yellow-green solid.

Yield: 68%. 1H NMR (400 MHz, CD3CN) d 8.38 (s, 2H), 8.31(d, J = 8.0 Hz, 2H), 7.90 (t, J = 8.0 Hz, 2H), 7.82 (d, J = 4.0 Hz, 2H), 7.60 (d, J = 8.0 Hz, 2H), 7.35 (d, J = 8.0 Hz, 2H), 7.08 (t, J = 8.0 Hz, 2H), 6.68 (t, J = 8.0 Hz, 2H), 5.73 (d, J = 8.0 Hz, 2H), 2.81 (t, J = 8.0 Hz, 4H), 1.73–1.66 (m, 4H), 1.33–1.27 (m, 24H), 0.87 (t, J = 4.0 Hz, 6H); 13C NMR (400 MHz, CD3CN) d 166.3, 166.2, 165.3, 165.2, 164.2, 164.0, 163.8, 163.6, 161.6, 161.5, 158.3, 156.8, 156.3, 156.2, 151.6, 150.8, 140.8, 129.8, 129.4, 126.2, 125.2, 125.1, 125.0, 115.2, 115.0, 100.2, 100.0, 99.7, 36.3, 33.0, 31.2, 30.5, 30.4, 30.3, 30.2, 23.8, 14.8; MALDI-TOF-HRMS: Calcd. For C50H56F4IrN4 [M-PF6]+: 981.4068. Found: 981.4089

Photophysical measurement

Emission spectra and lifetime measurements for complex 1 were performed on a PTI TimeMaster C720 Spectrometer (Nitrogen laser: pulse output 337 nm) fitted with a 380 nm filter. Error limits were estimated: λ (±1 nm); τ (±10%); φ (±10%). All solvents used for the lifetime measurements were degassed using three cycles of freeze-vac-thaw.

Luminescence quantum yields were determined using the method of Demas and Crosby [Ru(bpy)3][PF6]2 in degassed acetonitrile as a standard reference solution (Φr = 0.062) and calculated according to the following equation:

where the subscripts s and r refer to sample and reference standard solution respectively, n is the refractive index of the solvents, D is the integrated intensity, and Φ is the luminescence quantum yield. The quantity B was calculated by B = 1 – 10AL, where A is the absorbance at the excitation wavelength and L is the optical path length.

Hg2+ detection in aqueous buffered solution

Complex 1 (0.5 µM) and different concentrations of Hg2+ ions were added into Tris-HCl buffer (25 mM Tris, pH 7.0). Emission spectra were recorded in the 390−620 nm range using an excitation wavelength of 310 nm.

Supporting Information

Figure S1.

Emission and excitation spectrum of complex 1 (20 µM) in acetonitrile solution at 298K.

https://doi.org/10.1371/journal.pone.0060114.s001

(TIF)

Figure S2.

UV/Vis spectrum of complex 1 (20 µM) in acetonitrile solution at 298 K.

https://doi.org/10.1371/journal.pone.0060114.s002

(TIF)

Figure S3.

Relative intensity change at 490 nm of various concentrations of complex 1 in Tris-HCl buffer (25 mM Tris, pH 7.0) with the same concentration of Hg2+ ions (30 µM).

https://doi.org/10.1371/journal.pone.0060114.s003

(TIF)

Figure S4.

Emission spectrum of complex 1 (0.5 µM) upon addition of Hg2+ (30 µM) and upon subsequent addition of cysteine (0–80 µM) in buffered solution (25 mM Tris, pH 7.0).

https://doi.org/10.1371/journal.pone.0060114.s004

(TIF)

Figure S5.

1H NMR spectrum of 1 (5 µM, upper panel) in the absence or in the presence of Hg2+ ions (500 µM, lower panel).

https://doi.org/10.1371/journal.pone.0060114.s005

(TIF)

Figure S6.

Emission spectrum of complex 1 (0.5 µM) upon addition with mix  =  Pb2+, Fe3+, Co2+, La3+, Ti3+ (each 150 µM) and upon subsequent addition of Hg2+ (30 µM) in aqueous buffered solution (25 mM Tris, pH 7.0).

https://doi.org/10.1371/journal.pone.0060114.s006

(TIF)

Table S1.

Photophysical properties of the iridium complex 1.

https://doi.org/10.1371/journal.pone.0060114.s007

(DOCX)

Author Contributions

Directed the research: DLM CHL. Conceived and designed the experiments: WCF DLM. Performed the experiments: WCF MW LJL. Analyzed the data: WCF MW. Contributed reagents/materials/analysis tools: DLM CHL. Wrote the paper: WCF DSHC.

References

  1. 1. Harris HH, Pickering IJ, George GN (2003) The Chemical Form of Mercury in Fish. Science 301: 1203.
  2. 2. Clarkson TW, Magos L, Myers GJ (2003) The Toxicology of Mercury — Current Exposures and Clinical Manifestations. New Engl J Med 349: 1731–1737.
  3. 3. Clifton Ii JC (2007) Mercury Exposure and Public Health. Pediatr Clin North Am 54: 237.e231–237.e245.
  4. 4. Dobrowolski R, Mierzwa J (1992) Determination of mercury in fluorescent lamp cullet by atomic absorption spectrometry. Analyst 117: 1165–1167.
  5. 5. Slevin PJ, Györy-Szebényi E, Svehla G (1972) Application of displacement reactions in flame photometry—II: Emission flame photometric determination of alkaline earth metals in the presence of interfering anions. Talanta 19: 307–315.
  6. 6. Jarzynska G, Falandysz J (2011) The determination of mercury in mushrooms by CV-AAS and ICP-AES techniques. J Environ Sci Health A 46: 569–573.
  7. 7. Piette M, Desmet B, Dams R (1994) Determination of strontium in human whole blood by ICP-AES. Sci Total Environ 141: 269–273.
  8. 8. Usuda K, Kono K, Hayashi S, Kawasaki T, Mitsui G, et al. (2006) Determination of reference concentrations of strontium in urine by inductively coupled plasma atomic emission spectrometry. Environ Health Prev Med 11: 11–16.
  9. 9. Bloch P, Shapiro IM (1981) An x-ray fluorescence technique to measure the mercury burden of dentists in vivo. Med Phys 8: 308–311.
  10. 10. Lucchesi CA (1957) Determination of Strontium by X-Ray Fluorenscence Spectrometry. Anal Chem 29: 370–373.
  11. 11. Pejovic-Milic A, Stronach IM, Gyorffy J, Webber CE, Chettle DR (2004) Quantification of bone strontium levels in humans by in vivo x-ray fluorescence. Med Phys 31: 528–538.
  12. 12. Chen Y, Bai H, Hong W, Shi G (2009) Fluorescence detection of mercury ions in aqueous media with the complex of a cationic oligopyrene derivative and oligothymine. Analyst 134: 2081–2086.
  13. 13. Sivaraman G, Anand T, Chellappa D (2012) Development of a pyrene based "turn on" fluorescent chemosensor for Hg2+. RSC Adv DOI:https://doi.org/10.1039/C2RA21202A.
  14. 14. Zheng H, Zhang XJ, Cai X, Bian QN, Yan M, et al. (2012) Ratiometric Fluorescent Chemosensor for Hg2+ Based on Heptamethine Cyanine Containing a Thymine Moiety. Org Lett 14: 1986–1989.
  15. 15. Xu Z, Yoon J, Spring DR (2010) Fluorescent chemosensors for Zn2+. Chem Soc Rev 39: 1996–2006.
  16. 16. Xu Z, Han SJ, Lee C, Yoon J, Spring DR (2010) Development of off-on fluorescent probes for heavy and transition metal ions. Chem Commun 46: 1679–1681.
  17. 17. Ou D, Qin J, Li Z (2012) A new disubstituted polyacetylene bearing DDTC moieties: Postfunctional synthetic strategy, selective and sensitive chemosensor towards mercury ions. Polymer 53: 5691–5698.
  18. 18. Wang X, Iqbal M, Huskens J, Verboom W (2012) Turn-On Fluorescent Chemosensor for Hg2+ Based on Multivalent Rhodamine Ligands. Int J Mol Sci 13: 16822–16832.
  19. 19. Mandal D, Thakur A, Ghosh S (2012) A triazole tethered triferrocene derivative as a selective chemosensor for mercury(II) in aqueous environment. Polyhedron doi:10.1016/j.poly.2012.06.060
  20. 20. Miao P, Liu L, Li Y, Li G (2009) A novel electrochemical method to detect mercury (II) ions. Electrochem Commun 11: 1904–1907.
  21. 21. Yantasee W, Lin Y, Hongsirikarn K, Fryxell GE, Addleman R, et al. (2007) Electrochemical sensors for the detection of lead and other toxic heavy metals: the next generation of personal exposure biomonitors. Environ Health Persp 115: 1683–1690.
  22. 22. Sancenon F, Martinez-Manez R, Soto J (2001) Colourimetric detection of Hg2+ by a chromogenic reagent based on methyl orange and open-chain polyazaoxaalkanes. Tetrahedron Lett 42: 4321–4323.
  23. 23. Kim HN, Ren WX, Kim JS, Yoon J (2012) Fluorescent and colorimetric sensors for detection of lead, cadmium, and mercury ions. Chem Soc Rev 41: 3210–3244.
  24. 24. Yin Z, Tam AYY, Wong KMC, Tao CH, Li B, et al. (2012) Functionalized BODIPY with various sensory units - a versatile colorimetric and luminescent probe for pH and ions. Dalton Trans 41: 11340–11350.
  25. 25. Brulatti P, Gildea RJ, Howard JAK, Fattori V, Cocchi M, et al. (2012) Luminescent Iridium(III) Complexes with N∧C∧N-Coordinated Terdentate Ligands: Dual Tuning of the Emission Energy and Application to Organic Light-Emitting Devices. Inorg Chem 51: 3813–3826.
  26. 26. Yang C-H, Mauro M, Polo F, Watanabe S, Muenster I, et al. (2012) Deep-Blue-Emitting Heteroleptic Iridium(III) Complexes Suited for Highly Efficient Phosphorescent OLEDs. Chem Mater 24: 3684–3695.
  27. 27. Chan DSH, Lee HM, Che CM, Leung CH, Ma DL (2009) A selective oligonucleotide-based luminescent switch-on probe for the detection of nanomolar mercury(ii) ion in aqueous solution. Chem Commun 7479– 7481.
  28. 28. Ma DL, Ma VPY, Chan DSH, Leung KH, He HZ, et al. (2012) Recent advances in luminescent heavy metal complexes for sensing. Coord Chem Rev. doi:https://doi.org/10.1016/j.ccr.2012.07.005.
  29. 29. Leung KH, Ma VPY, He HZ, Chan DSH, Yang H, et al. (2012) A highly selective G-quadruplex-based luminescent switch-on probe for the detection of nanomolar strontium(ii) ions in sea water. RSC Adv 2: 8273–8276.
  30. 30. Wadas TJ, Wang Q-M, Kim Y-j, Flaschenreim C, Blanton TN, et al. (2004) Vapochromism and Its Structural Basis in a Luminescent Pt(II) Terpyridine−Nicotinamide Complex. J Am Chem Soc 126: 16841–16849.
  31. 31. Ma DL, Che CM, Yan SC (2008) Platinum(II) Complexes with Dipyridophenazine Ligands as Human Telomerase Inhibitors and Luminescent Probes for G-Quadruplex DNA. J Am Chem Soc 131: 1835–1846.
  32. 32. Zhao Q, Li F, Huang C (2010) Phosphorescent chemosensors based on heavy-metal complexes. Chem Soc Rev 39: 3007–3030.
  33. 33. Zhou J, Liu Z, Li F (2012) Upconversion nanophosphors for small-animal imaging. Chem Soc Rev 41: 1323–1349.
  34. 34. Tian X, Gill MR, Cantón I, Thomas JA, Battaglia G (2011) Live Cell Luminescence Imaging As a Function of Delivery Mechanism. ChemBioChem 12: 548–551.
  35. 35. Zhao Q, Huang C, Li F (2011) Phosphorescent heavy-metal complexes for bioimaging. Chem Soc Rev 40: 2508–2524.
  36. 36. Baggaley E, Weinstein JA, Williams JAG (2012) Lighting the way to see inside the live cell with luminescent transition metal complexes. Coord Chem Rev 256: 1762–1785.
  37. 37. Murphy L, Congreve A, Palsson L-O, Williams JAG (2010) The time domain in co-stained cell imaging: time-resolved emission imaging microscopy using a protonatable luminescent iridium complex. Chem Commun 46: 8743–8745.
  38. 38. Yang Y, Zhao Q, Feng W, Li F (2013) Luminescent Chemodosimeters for Bioimaging. Chem Rev 113: 192–270.
  39. 39. Liu J, Liu Y, Liu Q, Li C, Sun L, et al. (2011) Iridium(III) Complex-Coated Nanosystem for Ratiometric Upconversion Luminescence Bioimaging of Cyanide Anions. J Am Chem Soc 133: 15276–15279.
  40. 40. Li C, Liu Y, Wu Y, Sun Y, Li F (2013) The cellular uptake and localization of non-emissive iridium(III) complexes as cellular reaction-based luminescence probes. Biomaterials 34: 1223–1234.
  41. 41. Li C, Yu M, Sun Y, Wu Y, Huang C, et al. (2011) A Nonemissive Iridium(III) Complex That Specifically Lights-Up the Nuclei of Living Cells. J Am Chem Soc 133: 11231–11239.
  42. 42. Ge G, He J, Guo H, Wang F, Zou D (2009) Highly efficient phosphorescent iridium (III) diazine complexes for OLEDs: Different photophysical property between iridium (III) pyrazine complex and iridium (III) pyrimidine complex. J Organomet Chem 694: 3050–3057.
  43. 43. Jia QX, Qian XB, Wu HH, Wang QL, Gao EQ (2009) Manganese(II) coordination polymers with bis(5-tetrazolyl)methane: Synthesis, structure and magnetic properties. Inorg Chim Acta 362: 2213–2216.
  44. 44. Prokhorov AM, Santoro A, Williams JAG, Bruce DW (2012) Phosphorescent Mesomorphic Dyads Based on Tetraacetylethane Complexes of Iridium(III). Angew Chem Int Ed 51: 95–98.
  45. 45. Mauro M, De Paoli G, Otter M, Donghi D, D'Alfonso G, et al. (2011) Aggregation induced colour change for phosphorescent iridium(iii) complex-based anionic surfactants. Dalton Trans 40: 12106–12116.
  46. 46. Fernández-Hernández JM, Yang C-H, Beltrán JI, Lemaur V, Polo F, et al. (2011) Control of the Mutual Arrangement of Cyclometalated Ligands in Cationic Iridium(III) Complexes. Synthesis, Spectroscopy, and Electroluminescence of the Different Isomers. J Am Chem Soc 133: 10543–10558.
  47. 47. Bronner C, Veiga M, Guenet A, De Cola L, Hosseini MW, et al. (2012) Excited State Properties and Energy Transfer within Dipyrrin-Based Binuclear Iridium/Platinum Dyads: The Effect of ortho-Methylation on the Spacer. Chem Eur J 18: 4041–4050.
  48. 48. Santoro A, Prokhorov AM, Kozhevnikov VN, Whitwood AC, Donnio B, et al. (2011) Emissive Metallomesogens Based on 2-Phenylpyridine Complexes of Iridium(III). J Am Chem Soc 133: 5248–5251.
  49. 49. Mauro M, Schuermann KC, Prétôt R, Hafner A, Mercandelli P, et al. (2010) Complex Iridium(III) Salts: Luminescent Porous Crystalline Materials. Angew Chem Int Ed 49: 1222–1226.
  50. 50. Zhao Q, Liu S, Li F, Yi T, Huang C (2008) Multisignaling detection of Hg2+ based on a phosphorescent iridium(III) complex. Dalton Trans 3836–3840.
  51. 51. Liu Y, Li M, Zhao Q, Wu H, Huang K, et al. (2011) Phosphorescent Iridium(III) Complex with an N∧O Ligand as a Hg2+-Selective Chemodosimeter and Logic Gate. Inorg Chem 50: 5969–5977.
  52. 52. Wu Y, Jing H, Dong Z, Zhao Q, Wu H, et al. (2011) Ratiometric Phosphorescence Imaging of Hg(II) in Living Cells Based on a Neutral Iridium(III) Complex. Inorg Chem 50: 7412–7420.
  53. 53. Tong B, Mei Q, Lu M (2012) A highly selective chemosensor for mercury(II) cations based on cyclometalated iridium(III) complex. Inorg Chim Acta 391: 15–19.
  54. 54. Guerrero-Martinez A, Vida Y, Dominguez-Gutierrez D, Albuquerque RQ, De Cola L (2008) Tuning Emission Properties of Iridium and Ruthenium Metallosurfactants in Micellar Systems. Inorg Chem 47: 9131–9133.
  55. 55. Kim CY, Ha DG, Kang HH, Yun HJ, Kwon SK, et al. (2012) Synthesis and characterization of new blue light emitting iridium complexes containing a trimethylsilyl group. J Mater Chem
  56. 56. Nelson A, Auffret N, Borlakoglu J (1990) Interaction of hydrophobic organic compounds with mercury adsorbed dioleoylphosphatidylcholine monolayers. BBA-Biomembranes 1021: 205–216.
  57. 57. Jia SM, Liu XF, Li P, Kong DM, Shen HX (2011) G-quadruplex DNAzyme-based Hg2+ and cysteine sensors utilizing Hg2+-mediated oligonucleotide switching. Biosens Bioelectron 27: 148–152.
  58. 58. Zhang M, Le HN, Wang P, Ye BC (2012) A versatile molecular beacon-like probe for multiplexed detection based on fluorescence polarization and its application for a resettable logic gate. Chem Commun 48: 10004–10006.
  59. 59. Lowry MS, Hudson WR, Pascal RA, Bernhard S (2004) Accelerated Luminophore Discovery through Combinatorial Synthesis. J Am Chem Soc 126: 14129–14135.