Acessibilidade / Reportar erro

Near-Infrared Luminescence from Visible-Light-Sensitized Ruthenium(II)-Neodymium(III) Heterobimetallic Bridged Complexes Containing Alkoxy(silyl) Functional Groups

Abstract

New infrared emitting d-f (ruthenium(II)-neodymium(III)) heterobimetallic complexes with alkoxy(silyl) functional groups have been prepared. Visible excitation evidenced energy transfer processes from the ruthenium(II) donor to neodymium(III) acceptors leading to infrared emission. Energy transfer rates (kEnT) and efficiency of energy transfer (ηEnT) are, respectively, 0.61 × 107s-1 and 44% for RuL1-NdL3 complex. Larger values of kEnT (3.04 × 107 s-1) and ηEnT (84%) were detected for RuL2-NdL4 complex. RuL1-NdL3 and RuL2-NdL4 complexes were fully characterized by elementary analysis (EA), mass spectrometry (MS), Fourier transform infrared spectroscopy (FTIR) and Fourier transform Raman spectroscopy (FT-Raman). Total correlation spectroscopy (TOCSY1D), 1H{13C} heteronuclear single quantum correlation (HSQC) and 1H{13C} heteronuclear multiple bond correlation (HMBC) nuclear magnetic resonance (NMR) analyses were also carried out to characterize NdL3 and RuL1-NdL3 complexes. The presence of trialkoxysilyl-substituted ligands would allow further grafting onto any silica or silicated surface aiming at applications as new luminescent near infrared (NIR)-emitting biosensors or biomarkers.

Keywords:
ruthenium(II); neodymium(III); silylated heterobimetallic complexes; d-f energy transfer; grafting trialkoxysilyl group; visible and NIR luminescent complexes


Introduction

The design and assembly of heterobimetallic complexes based on d-block metalloligands with near-infrared (NIR) emitting lanthanides (Ln) have attracted interest in the recent years due to potential applications in non-invasive bio-analysis and bio-imaging.11 Chen, F.-F.; Chen, Z.-Q.; Bian, Z.-Q.; Huang, C.-H.; Coord. Chem. Rev. 2010, 254, 991.

2 Chen, F.-F.; Wei, H.; Bian, Z.; Liu, Z.; Ma, E.; Chen, Z.; Huang, C.; Organometallics 2014, 33, 3275.

3 Jana, A.; Baggaley, E.; Amoroso, A.; Ward, M. D.; Chem. Commun. 2015, 51, 8833.

4 Baggaley, E.; Cao, D.-K.; Sykes, D.; Botchway, S. W.; Weinstein, J. A.; Ward, M. D.; Chem. - Eur. J. 2014, 20, 8898.

5 Pope, S. J. A.; Coe, B. J.; Faulkner, S.; Laye, R. H.; Dalton Trans. 2005, 1482.
-66 Li, L.; Zhang, S.; Xu, L.; Chen, Z.-N.; Luo, J.; J. Mater. Chem. C 2014, 2, 1698. Given the inherent low molar absorption coefficients of lanthanide(III) ions (ε < 10 mol L−1 cm−1), d-block chromophores serve as effective light-harvesting groups to sensitize Ln-emissive centres.11 Chen, F.-F.; Chen, Z.-Q.; Bian, Z.-Q.; Huang, C.-H.; Coord. Chem. Rev. 2010, 254, 991. Beyond that, most of lanthanide complexes require excitation in the ultraviolet (UV) region, which is harmful for the biological systems. The aromatic residues of proteins and deoxyribonucleic acid (DNA) absorb in competition with the chromophores. In addition, the UV excitation cause damages in these biological systems.11 Chen, F.-F.; Chen, Z.-Q.; Bian, Z.-Q.; Huang, C.-H.; Coord. Chem. Rev. 2010, 254, 991. One of the strategies to overcome these limitations consists on using d-block chromophores as antenna groups to sensitize luminescence from lanthanide(III) ions with low energy f-f excited states.

Typically, ruthenium(II) complexes displaying polypyridine bridging ligands are used to sensitize NIR emitting lanthanide ions through the strong absorption of the visible light by 3MLCT (MLCT = metal-to-ligand charge transfer) excited state.11 Chen, F.-F.; Chen, Z.-Q.; Bian, Z.-Q.; Huang, C.-H.; Coord. Chem. Rev. 2010, 254, 991. These d-block chromophores, with relatively low 3MLCT states, act as good energy donors to f-f levels of praseodymium(III), neodymium(III), erbium(III) and ytterbium(III). In particular, [Ru(bpy)3]2+ (bpy = 2,2’-bipyridine) possesses a broad and strong absorption in the visible and a low triplet state (energy of triplet state = 17400 cm-1) that can allow lanthanide(III) to be efficiently excited.11 Chen, F.-F.; Chen, Z.-Q.; Bian, Z.-Q.; Huang, C.-H.; Coord. Chem. Rev. 2010, 254, 991. Authors77 Lazarides, T.; Sykes, D.; Faulkner, S.; Barbieri, A.; Ward, M. D.; Chem. - Eur. J. 2008, 14, 9389. reported the use of distinct lengthy bridging ligands (saturated or unsaturated) between energy donor (d-block) and acceptor fragments (f-block) and concluded that the role of ligands is vital in determining the ruthenium(II) → lanthanide(III) energy transfer (EnT), and the energy level of triplet state was related with the bridging processes.77 Lazarides, T.; Sykes, D.; Faulkner, S.; Barbieri, A.; Ward, M. D.; Chem. - Eur. J. 2008, 14, 9389.

Pyridine derivate ligands can also serve as bridging ligands to prepare ruthenium(II)-lanthanide(III) heterobimetallic d-f complexes. Specifically, 2,2’-bipyrimidine (bpmd) is widely used to synthesize planar heterobimetallic complexes since it can easily coordinate two metal centers through four nitrogen sites leading to a relatively short metal-metal distance. As a result, an efficient channel is available for energy and electron transfer from d-block chromophores in the visible region to NIR emitting lanthanide(III) centres.77 Lazarides, T.; Sykes, D.; Faulkner, S.; Barbieri, A.; Ward, M. D.; Chem. - Eur. J. 2008, 14, 9389.

8 Shavaleev, N. M.; Accorsi, G.; Virgili, D.; Bell, Z. R.; Lazarides, T.; Calogero, G.; Armaroli, N.; Ward, M. D.; Inorg. Chem. 2005, 44, 61.

9 Lazarides, T.; Tart, N. M.; Sykes, D.; Faulkner, S.; Barbieri, A.; Ward, M. D.; Dalton Trans. 2009, 3971.

10 Lazarides, T.; Adams, H.; Sykes, D.; Faulkner, G.; Calogero, A.; Ward, M. D.; Dalton Trans. 2008, 691.

11 Zucchi, G.; le Goff, X. F.; Polyhedron 2013, 52, 1262.

12 Zucchi, G.; Maury, O.; Thuéry, P.; Ephritikhine, M.; Inorg. Chem. 2008, 47, 10398.

13 Ward, M. D.; Coord. Chem. Rev. 2010, 254, 2634.
-1414 Fratini, A.; Richards, G.; Larder, E.; Swavey, S.; Inorg. Chem. 2008, 47, 1030.

Despite the examples of ruthenium(II)-lanthanide(III) heterobimetallic complexes being reported in the last few years, their use as luminophors for biosensors or biomarkers is yet scarce. The most likely reasons are their low water solubility and relatively weak luminescence.1515 Wei, Q.-H.; Lei, Y.-F.; Xu, W.-R.; Xie, J.-M.; Chen, G.-N.; Dalton Trans. 2012, 41, 11219. As an alternative to explore the potentialities of such heterobimetallic complexes in biological applications, these complexes need to be functionalized to be covalently anchored on silica nano and microparticles. We have demonstrated the preparation of bipyridine and diketone derivative ligands such as bpy-Si (4-methyl-4’-[methylamino-3(propyltriethoxysilyl)]-2,2’-dipyridine)1616 Cousinié, S.; Gressier, M.; Alphonse, P.; Menu, M.-J.; Chem. Mater. 2007, 19, 6492. and TTA-Si (4,4,4-trifluoro-2-(3-(trimethoxysilyl)-propyl)-1-(2-thienyl)-1,3-butanedione)1717 Duarte, A. P.; Gressier, M.; Menu, M.-J.; Dexpert-Ghys, J.; Caiut, J. M. A.; Ribeiro, S. J. L.; J. Phys. Chem. C. 2012, 116, 505. appropriately substituted by a trialkoxysilyl groups. Sequentially, silylated ruthenium(II) and lanthanides(III) complexes have been isolated and grafted at the surface1616 Cousinié, S.; Gressier, M.; Alphonse, P.; Menu, M.-J.; Chem. Mater. 2007, 19, 6492.

17 Duarte, A. P.; Gressier, M.; Menu, M.-J.; Dexpert-Ghys, J.; Caiut, J. M. A.; Ribeiro, S. J. L.; J. Phys. Chem. C. 2012, 116, 505.

18 Cousinié, S.; Mauline, L.; Gressier, M.; Kandibanda, S. R.; Datas, L.; Reber, C.; Menu, M.-J.; New J. Chem. 2012, 36, 1355.

19 Sábio, R. M.; Gressier, M.; Caiut, J. M. A.; Menu, M.-J.; Ribeiro, S. J. L.; Mater. Lett. 2016, 174, 1.

20 Cousinié, S.; Gressier, M.; Reber, C.; Dexpert-Ghys, J.; Menu, M.-J.; Langmuir 2008, 24, 6208.
-2121 Duarte, A. P.; Mauline, L.; Gressier, M.; Dexpert-Ghys, J.; Roques, C.; Caiut, J. M. A.; Deffune, E.; Maia, D. C. G.; Carlos, I. Z.; Ferreira, A. A. P.; Ribeiro, S. J. L.; Menu, M.-J.; Langmuir 2013, 29, 5878. or embedded1717 Duarte, A. P.; Gressier, M.; Menu, M.-J.; Dexpert-Ghys, J.; Caiut, J. M. A.; Ribeiro, S. J. L.; J. Phys. Chem. C. 2012, 116, 505.,1818 Cousinié, S.; Mauline, L.; Gressier, M.; Kandibanda, S. R.; Datas, L.; Reber, C.; Menu, M.-J.; New J. Chem. 2012, 36, 1355.,2121 Duarte, A. P.; Mauline, L.; Gressier, M.; Dexpert-Ghys, J.; Roques, C.; Caiut, J. M. A.; Deffune, E.; Maia, D. C. G.; Carlos, I. Z.; Ferreira, A. A. P.; Ribeiro, S. J. L.; Menu, M.-J.; Langmuir 2013, 29, 5878.,2222 Mauline, L.; Gressier, M.; Roques, C.; Hammer, P.; Ribeiro, S. J. L.; Caiut, J. M. A.; Menu, M.-J.; Biofouling 2013, 29, 775. in nanosized silica. These previous results stimulated us to design d-f heterobimetallic complexes by introducing silylated appropriated ligands, bpy-Si or TTA-Si, involving NIR emitting lanthanides as neodymium(III) and bipyrimidine as bridging d-f ligand.

In this way, two routes were employed to obtain two types of dyads based on the chemistry of bipyridine ruthenium(II) and diketonate neodymium(III) derived complexes according to the position of the silylated functional group. For the sake of brevity we will use a labelling Scheme 1 whereby the complexes are referred to as: RuL2-NdL4 dinuclear complexes, bearing the alkoxysilyl group on bipyridine ruthenium moiety (i), and RuL1-NdL3 bearing the alkoxysilyl groups on the TTA lanthanide complex (ii). In both cases bipyrimidine is the bridging ligand on ruthenium center. Successful structural characterization of the silylated d-f heterobimetallic complexes is described and discussed in this work. Furthermore, photophysical properties of these heterobimetallic complexes are examined with respect to the efficiency of energy transfer processes from ruthenium(II) moieties to the luminescent NIR-emitting neodymium(III) ions.

Scheme 1
Synthesis of the heterobimetallic complexes, RuL1-NdL3 (i) and RuL2-NdL4 (ii). Numbering of the hydrogen and carbon atoms are identified as red and blue labels.

Experimental

Materials

The silylating ligands, 4-methyl-4’-[methylamino-3(propyltriethoxysilyl)]-2,2’-dipyridine (bpy-Si), and 4,4,4-trifluoro-2-(3-(trimethoxysilyl)-propyl)-1-(2-thienyl)-1,3-butanedione (TTA-Si), were synthesized as previously described by Menu and co-workers.1616 Cousinié, S.; Gressier, M.; Alphonse, P.; Menu, M.-J.; Chem. Mater. 2007, 19, 6492.,1717 Duarte, A. P.; Gressier, M.; Menu, M.-J.; Dexpert-Ghys, J.; Caiut, J. M. A.; Ribeiro, S. J. L.; J. Phys. Chem. C. 2012, 116, 505. [cis-RuCl2(bpy)2], (bpy = 2,2’-bipyridine), [RuCl2(bpy)(bpy-Si)] and [Nd(TTA)3].2H2O were synthesized as previously described.1616 Cousinié, S.; Gressier, M.; Alphonse, P.; Menu, M.-J.; Chem. Mater. 2007, 19, 6492.,1717 Duarte, A. P.; Gressier, M.; Menu, M.-J.; Dexpert-Ghys, J.; Caiut, J. M. A.; Ribeiro, S. J. L.; J. Phys. Chem. C. 2012, 116, 505. Thenoyltrifluoro-acetone (TTA), neodymium chloride anhydrous, ytterbium chloride anhydrous were purchased from Aldrich (Toulouse, France). 2,2’-Bipyrimidine were purchased from Acros (Toulouse, France). These reagents were used as received. Acetone, dichloromethane, diethylether, ethanol and pentane were purified by distillation in an inert atmosphere, degassed and dried using cryogenic and drying procedures, respectively. All manipulations concerning the preparation of ligands and complexes were performed in an inert atmosphere using the Schlenk tube technique.

Characterization methods

Fourier transform infrared (FTIR) spectra were obtained in the spectral range from 4000 to 650 cm-1 (4 cm-1 of resolution) with a Bruker Vector 22 (KBr dispersion). The Raman spectra from all complexes were collected on a RFS 100 FT-Raman Bruker spectrometer equipped with a Ge detector using liquid nitrogen as the coolant and a Nd:YAG laser emitting at 1064 nm. The laser light, with a power varying from 30 to 150 mW, was introduced and focused on the sample, and the scattered radiation was collected at 180°. For each spectrum, an average of 1024 scans was performed at a resolution of 4 cm-1 over a range from 3500 to 50 cm-1. The OPUS 6.0 (Bruker Optik, Ettlingen, Germany) software was used for Raman data acquisition. For the lanthanide and heterobimetallic complexes, nuclear magnetic resonance (NMR) spectra were obtained using a Bruker Advance 600 MHz spectrometer at room temperature in deuterated solvents. Total correlation spectroscopy (TOCSY1D) spectra were recorded by selective irradiation of different aromatic hydrogen atoms of the complex with a selective pulse. The spectra were recorded after magnetization transfer at 120 ms of mixing time. 1H{13C} heteronuclear single quantum correlation (HSQC) NMR experiments were recorded using the spectral widths of 12 000 and 24 000 Hz for the 1H and 13C dimensions, respectively. The number of collected complex points was 1024 for the 1H dimension with a recycle delay of 1.5 s. The number of transients was from 24 to 48 depending of the sample and 128 time increments were recorded in the 13C dimension. The 1JH-C used was 145 Hz. Prior to Fourier transformation, the data matrices were zero filled to 1024 points in the 13C dimension. 1H{13C} heteronuclear multiple bond correlation (HMBC) NMR experiments were carried out using the spectral widths of 12000 and 24000 Hz for the 1H and 13C dimensions, respectively. The number of collected complex points was 1024 for the 1H dimension with a recycle delay of 2.0 s. The number of transients was from 16 to 32 depending of the sample and 128 time increments were recorded in the 13C dimension. The 3JH-C used was 8 Hz. Prior to Fourier transformation, the data matrices were zero filled to 1024 points in the 13C dimension. Data processing was performed using standard Bruker Topspin-NMR software. The central solvent line was used as an internal chemical shift reference point. Mass spectra (MS) were recorded using Q TOF 1er (Waters) spectrometer: source electrospray (ESI) or DSQ Thermo Fisher Scientific spectrometer equipped with chemical ionization source (NH3). UV-Vis spectra were recorded using a Varian spectrophotometer model Cary 5000 in the region of 800 to 200 nm (resolution of 1 nm) using Cary WinUV software. Elemental analyses of C, H, N and S were performed on a Carlo Erba instrument (EA 1110). Luminescence spectra were measured at room temperature using a Jobin-Yvon Model Fluorolog FL3-22 spectrometer equipped with a H10330-75 Hamamatsu detector, TE: cooled NIR-photomultiplier module and a 450 W Xe excitation lamp. Excitation and emission spectra were recorded under CW excitation and were corrected with respect to the Xe lamp intensity and spectrometer response. Fluorescence intensity decays were obtained using the time-correlated single-photon counting technique. The excitation source was a mode-locked Ti:saphire laser (Tsunami 3950 pumped by Millennia X Spectra Physics) producing 5 ps FWHM pulses ranging from 0.5 to 8.0 MHz repetition rate, regulated by the 3980 Spectra Physics pulse picker. The laser was tuned to give output at 892 nm and a second harmonic generator BBO crystal (GWN-23PL Spectra Physics) gave the 448 nm excitation pulses that were directed to an Edinburgh FL900 spectrometer adjusted in L-format configuration. The emission wavelength of 620 and 670 nm were selected by a monochromator and emitted photons were detected by a cooled Hamamatsu R3809U microchannel plate photo-multiplier. The whole instrument response function was typically 100 ps. Energy transfer rate constant (kEnT) and efficiency of energy transfer (ηEnT) were obtained using ruthenium 3MLCT decay values from ruthenium precursors and the respective heterobimetallic complexes.

Synthesis

[Nd(TTA-Si)3] (NdL3)

This complex was obtained by the procedure adapted from Duarte et al.1717 Duarte, A. P.; Gressier, M.; Menu, M.-J.; Dexpert-Ghys, J.; Caiut, J. M. A.; Ribeiro, S. J. L.; J. Phys. Chem. C. 2012, 116, 505. 6 mmol (2.30 g) of the TTA-Si and 2 mmol (0.50 g) of neodymium chloride anhydrous were added to 20 mL of anhydrous ethanol. The reaction mixture was kept under nitrogen atmosphere and stirred for 18 h. The solvent was evaporated and the powder was then washed with pentane and diethylether, and dried under vacuum. The Nd(TTA-Si)3 complex (NdC42H54F9O15S3Si3, MW 1294.54) was obtained as a greenish powder (1.9 g, yield: 77%). Elemental analysis found: C 38.45, H 3.5; calcd. for NdC41H52O15F9S3Si3 (one methoxy group hydrolyzed): C 38.45, H 4.1%; UV (EtOH) λ / nm 267 and 340; IR (KBr) ν / cm-1 3108w (vCH), 2960m and 2923m (vasCH2,CH3), 2855w (vsCH2,CH3), 1609vs-br (vCO), 1539s and 1510m (vCCO), 1410s and 1356s (vCCSthenoyl), 1304vs-br (vCF3), 1060s-br (vSiOCH3), 788s (δCHthenoyl), 721m (δCF3); 1H NMR (600 MHz, CDCl3) δ 0.83 (2H, CH2, 11-H), 1.90 (2H, CH2, 10-H), 3.57 (2H, CH2, 9-H and 9H, CH3, 12-H), 7.19 (1H, dd, JH-H 4.5 and 3.4 Hz, CH, 2-H), 7.75 (1H, d, JH-H 4.9 Hz, CH, 1-H), 7.79 (1H, d, JH-H 3.4 Hz, CH, 3-H); MS (ESI) m/z, found for [(M - OCH3+ OH) - TTA-Si]: 895.9.

[Nd(TTA)3(H2O)2] (NdL4)

This complex was obtained by the procedure described by Duarte et al.1717 Duarte, A. P.; Gressier, M.; Menu, M.-J.; Dexpert-Ghys, J.; Caiut, J. M. A.; Ribeiro, S. J. L.; J. Phys. Chem. C. 2012, 116, 505. 3 mmol (0.666 g) of TTA were added to 30 mL of ethanol. Then 450 mL of NH4OH were added. 1 mmol (0.279 g) of neodymium chloride were dissolved in 30 mL of water and added to the previous solution. The reaction mixture was kept under stirring for 3 h. The obtained powder was recovered and dissolved in acetone; then it was re-crystalized by precipitation on cold water and dried under vacuum. The [Nd(TTA)3H2O2] complex (NdC24H16O8F9S3, MW 843.80) was obtained as a yellow powder (0.59 g, yield: 70%). Elemental analysis found: C 34.1, H 1.5; calcd. for NdC24H16O8F9S3: C 34.1, H 1.9%; UV (EtOH) λ / nm 268 and 340; IR (KBr) ν / cm-1 3330m-br (vOH), 3115 (vCH), 1606vs (vCO), 1542s and 1514s (vCCO), 1411s and 1358s (vCCSthenoyl), 1300ns-br (vCF3), 789w (δCHthenoyl), 716w (δCF3); MS (desorption chemical ionization, DCI, NH3) m/z, found for [M + H+ - 2H2O]: 807.8; [M + H+ - H2O]: 825.8.

[Ru(bpy)2(bpmd)]Cl2 (RuL1)

1.88 mmol (0.307 g) of 2,2’-bipyrimidine and 1.71 mmol (0.830 g) of [cis-RuCl2(bpy)2], were added to 40 mL of anhydrous and degassed ethanol. The reaction mixture was kept under nitrogen atmosphere and refluxing at 78 °C for 72 h. The solvent was evaporated, and the powder was then washed with pentane and diethyl ether, and dried under vacuum. The [Ru(bpy)2(bpmd)]Cl2 complex (RuC28H22N8Cl2, MW 642.50) was obtained as a brown powder (1.1 g, yield: 85%). Elemental analysis found: C 49.7, H 3.4, N 16.8; calcd. for RuC28H22N8Cl2.H2O: C 52.3, H 3.45, N 17.4%; UV (EtOH) λ / nm 245, 286 and 424; IR (KBr) ν / cm-1 3064w (vCH), 1630m-br and 1601s (vCN), 1572s, 1463s, 1443s 1420m and 1400vs (vCC), 772vs (δCHar); MS (ESI) m/z, found for [M2+/2]: 286.1.

[Ru(bpy)(bpy-Si)(bpmd)]Cl2 (RuL2)

1.38 mmol (0.225 g) of 2,2’-bipyrimidine and 1.38 mmol (1.01 g) of [RuCl2(bpy)(bpy-Si)] were added to 40 mL of anhydrous and degassed ethanol. The reaction mixture was kept under nitrogen atmosphere and refluxing at 78 °C for 72 h. The solvent was evaporated, and the powder was washed with diethyl ether, and dried under vacuum. The [Ru(bpy)(bpy-Si)(bpmd)]Cl2 complex (RuC39H47N9O3SiCl2, MW 889.91) was obtained as a black powder (0.98 g, yield: 80%). Elemental analysis found: C 50.0, H 4.9, N 13.65; calcd. for RuC39H47N9O3SiCl2: C 52.6, H 5.3, N 14.1%; UV (EtOH) λ / nm 250, 287 and 443; IR (KBr) ν / cm-1 3064w (vCH), 2967w and 2923w (vasCH2,CH3), 2867w (vsCH2,CH3), 1637sh and 1618s (vCN), 1553s, 1464m, 1444m,1421m and 1400vs (vCC), 1072s-br (vSiOC), 771s (δCHar); MS (ESI) m/z, found for [M2+/2]: 409.4.

[Ru(bpy)2(bpmd)Nd(TTA-Si)3]Cl2 (RuL1-NdL3)

0.33 mmol (0.208 g) of the [Ru(bpy)2(bpmd)]Cl2 (RuL1) was added to 10 mL of anhydrous and degassed ethanol and then 0.33 mmol (0.420 g) of Nd(TTA-Si)3 (NdL3) diluted in 10 mL of anhydrous and degassed dichlorometane were added to the first solution. The reaction mixture was kept under nitrogen atmosphere for 12 h at room temperature. The solvent was half evaporated and filtered. The obtained powder was washed with pentane and diethyl ether, and dried under vacuum. The [Ru(bpy)2(bpmd)Nd(TTA-Si)3]Cl2 complex (C70H76N8O15F9S3Si3RuNdCl2, MW: 1937.05) was obtained as a dark brown powder (0.35 g, yield: 55%). Elemental analysis found: C 41.4, H 2.95, N 5.8, S 5.1; calcd. for C70H76N8O15F9S3Si3RuNdCl2: C 43.4, H 3.95, N 5.8, S 5.0%; UV (EtOH) λ / nm 257, 270, 285, 339 and 423; IR (KBr) ν / cm-1 3070w (vCH), 2972w (vasCH2,CH3), 1631sh and 1605vs (vCN), 1535s and 1508m (vCCO), 1576w, 1463m and 1446m (vCC), 1411s-br and 1354m (vCCSthenoyl), 1302vs (vCF3), 1060w (vSiOCH3), 785w (δCHthenoyl), 767w (δCHar), 728w (δCF3); 1H NMR (600 MHz, CDCl3) δ 0.85 (br s, CH2, 11-H), 1.91 (br s, CH2, 10-H), 3.59 (br s, CH2, 9-H and br s, CH3, 12-H ), 7.15 (br s, CH, IV’-H), 7.18 (dd, JH-H 6.2 Hz, V-H), 7.29 (dd, JH-H6.4 Hz, CH, 5-H), 7.38 (d, JH-H 4.9 Hz, VI-H), 7.53 (br s, CH, 2-H), 7.55 (br s, CH, 6-H), 7.65 (br s, CH, V’-H), 7.70 (dd, JH-H6.2 Hz, CH, IV-H), 7.80 (dd, JH-H7.7 Hz, CH, 4-H), 7.97 (d, JH-H 3.8 Hz, CH, 1-H), 8.15 (d, JH-H5.9 Hz, CH, III-H), 8.19 (d, JH-H 7.4 Hz, CH, 3-H), 8.72 (br s, CH, III’-H), 9.08 (br s, CH, 3-H); MS (ESI) m/z, found for [Ru(bpy)2(bpmd)]2+/2]: 286.1; [M - TTA-Si]: 1551.0.

[Nd(TTA)3(bpmd)Ru(bpy)(bpy-Si)]Cl2 (RuL2-NdL4)

0.28 mmol (0.250 g) of the [Ru(bpy)(bpy-Si)(bpmd)]Cl2 (RuL2) was added to 10 mL of anhydrous and degassed ethanol and then 0.28 mmol (0.240 g) of [Nd(TTA)3(H2O)2], (NdL4) diluted in 10 mL of anhydrous and degassed dichloromethane were added to the ruthenium solution. The reaction mixture was kept under nitrogen atmosphere for 12 h at room temperature. The solvent was half evaporated and filtered. The obtained powder was washed with pentane and diethyl ether, and dried under vacuum. The [Ru(bpy)(bpy-Si)(bpmd)Nd(TTA)3]Cl2 complex (C63H59N9O9F9S3SiRuNdCl2, MW 1697.68) was obtained as a black powder (0.26 g, yield: 55%). Elemental analysis found: C 44.15, H 3.5, N 7.6; calcd. for C63H59N9O9F9S3SiRuNdCl2: C 44.6, H 3.7, N 7.4%; UV (EtOH) λ / nm 258, 287, 339 and 437; IR (KBr) ν / cm-1 3069w (vCH), 2978w and 2923w (vasCH2,CH3), 1634sh and 1609vs (vCN), 1555s and 1506m (vCCO), 1575w, 1465m and 1446m (vCC), 1413s, and 1355m (vCCSthenoyl), 1302vs-br (vCF3), 1076w (vSiOC), 783m (CHthenoyl), 767w (δCHar), 717w (δCF3); MS (ESI) m/z, found for [M+ - Cl]: 1662.2.

Results and Discussion

Synthesis and structural characterization of silylated ruthenium(II)-neodymium(III) heterobinetallic complexes

The synthesis routes of three new mononuclear complexes denoted by (i) RuL1, (ii) RuL2 and (iii) NdL3 has been achieved in Supplementary Information (SI) section (Figure S1). Equimolar reaction of ruthenium complex bearing the bridging ligand, RuL1, with silylated tris(diketonate) neodymium complex, NdL3, gives rise to RuL1-NdL3 heterobimetallic complex (i,Scheme 1). On the other hand, the combination of RuL2 complex, containing both silyl and bridging groups, with tris(diketonate) neodymium complex, Nd(TTA)3 denoted by NdL4 (synthesis route and structure of NdL4 were done in Figure S2, SI section), leads to heterobimetallic RuL2-NdL4 complex with the trialkoxysilyl functional group on the ruthenium moiety (ii, Scheme 1).

FTIR spectra (Figure 1) exhibit indicative bands for lanthanide units in NdL3, NdL4 and both heterobimetallic complexes with the stronger vibration bands in the range 1603-1610 (vC=O), 1539-1543 (vC-C-O), 1411-1412 (vCS thienyl), 1300-1310 cm-1(vCF3) related to diketonate ligands. Concerning the NdL3 and RuL1-NdL3 complexes (Figure 1a red and blue lines, respectively), the presence of bands at 1060 (vSi-O-CH3) suggests the successful synthesis of both methoxysilyl complexes. These results are in agreement with EA and MS (molecular peaks for RuL1 and RuL1-NdL3 at m/z = 286.1 [M2+/2] and 286.1 [Ru(bpy)2(bpmd)]2+/2]; 1551.0 [M - TTA-Si], respectively).

Figure 1
FTIR spectra for the (a) NdL3, RuL1-NdL3 and RuL1; (b) NdL4, RuL2-NdL4 and RuL2 complexes. All analyses were carried out as KBr pellets.

For RuL2 and RuL2-NdL4 complexes (Figure 1b red and blue lines, respectively), three bands 1072 and 1076 cm-1 (vSi-O-C), respectively confirming that the reaction was performed avoiding the hydrolysis and self-polymerization of the ethoxysilyl groups. These results are in agreement with EA and MS (molecular peaks for RuL2 and RuL2-NdL4 at m/z = 409.4 [M2+/2] and 1662.2 [M+- Cl], respectively). The NdL4 complex (Figure 1b, red line) was successfully prepared and characterized as already described in the Experimental section.

Raman spectroscopy allows accurate assignment of CCH stretching vibration bands at 1028, 1040, 1202, 1490 and 1580 cm-1 of heterocycles of the bpy and bpmd ligands of ruthenium units in RuL1-NdL3 and RuL2-NdL4 heterobimetallic complexes. The ruthenium(II) coordination was confirmed by the band observed at 1030 cm-1 (vRu-N).2323 Silverstein, D. W.; Milojevich, C. B.; Camden, J. P.; Jensen, L.; J. Phys. Chem. C 2013, 117, 20855.,2424 Kokošková, M.; Procházka, M.; Šloufová, I.; Vlčková, B.; J. Phys. Chem. C 2013, 117, 1044. Both spectroscopies appeared well complement one another. Raman scattering spectra are shown in Figure S3 (SI section).

The steric arrangement of the ligands in the metal coordination sphere in NdL3 and RuL1-NdL3 complexes have been elucidated by assignment of chemical shifts, JH-C and JH-H coupling constant and using complementary NMR studies such as TOCSY1D, 1H{13C}-HSQC and 1H{13C}-HMBC experiments (Figures S4, S5 for NdL3 and S6-S9 for RuL1-NdL3, SI section). Numbering of the hydrogen and carbon atoms used in the further characterization is detailed in Figure S1 (SI section) and Scheme 1 for each complexes. Most probable because of the presence of the silyl group, compounds behave as hygroscopic powders and all attempts to grow single crystals have been unsuccessful.

Photophysical properties

Figure 2 shows the UV-Vis absorption spectra for the two types of heterobimetallic complexes and respective precursors. Figure 2a black line, shows the spectrum obtained for the RuL1 complex. Strong and broad bands are observed at 245 and 424 nm attributed to d→π* 1MLCT transitions (RuL1 to bpy and bpmd ligands, respectively) and a strong and sharp band at 286 nm attributed to p→π* transitions of bpy and bpmd ligands.2525 Zhang, L.-Y.; Hou, Y.-J.; Pan, M.; Chen, L.; Zhu, Y.-X.; Yin, S.-Y.; Shao, G.; Su, C.-Y.; Dalton Trans. 2015, 44, 15212. For NdL3 complex (Figure 2a, red line), the absorption bands at 269 and 340 nm were assigned to singlet-to-singlet transitions in the TTA-Si ligands.1717 Duarte, A. P.; Gressier, M.; Menu, M.-J.; Dexpert-Ghys, J.; Caiut, J. M. A.; Ribeiro, S. J. L.; J. Phys. Chem. C. 2012, 116, 505. The spectrum obtained for RuL1-NdL3 complex (Figure 2a, blue line) shows the strong and broad absorption bands at 257, 285 and 423 nm, attributed to the π→π* transitions of the ligands coordinated to the ruthenium center and d→π* 1MLCT transitions from the ruthenium center to bpy and bpmd ligands. The absorption bands at 270 and 339 nm are ascribed to singlet-to-singlet ligands transitions present in the neodymium(III) center. Concerning the RuL2 complex, the bpy-Si was confirmed by bands at 250 and 443 nm, attributed to d→π* 1MLCT transitions from the ruthenium(II) center to bpy-Si and bpmd ligands (Figure 2b, black line) and a strong and sharp band at 287 nm attributed to π→π* transitions of bpy-Si and bpmd ligands. Comparing spectra obtained for RuL1 and RuL2 one may note a red shift of around 20 nm for 1MLCT state with the addition of the silylated ligand. The spectrum obtained for NdL4 complex (Figure 2b, red line) displays two bands at 268 and 340 nm assigned to singlet-to-singlet transitions in the TTA ligands.1717 Duarte, A. P.; Gressier, M.; Menu, M.-J.; Dexpert-Ghys, J.; Caiut, J. M. A.; Ribeiro, S. J. L.; J. Phys. Chem. C. 2012, 116, 505. The spectrum of RuL2-NdL4 complex (Figure 2b, blue line) displays strong and broad absorption bands at 258, 287 and 437 nm, characteristic of π→π* transitions of the ligands coordinated to the ruthenium center and d→π*1MLCT transitions from the ruthenium to bpy-Si and bpmd ligands. The absorption bands at 258 to 339 nm are assigned to the singlet-to-singlet ligands transitions present in the neodymium(III) complex.

Figure 2
Absorption spectra (2.5 × 10-5 mol L-1, ethanol) of (a) NdL3 (red line), RuL1-NdL3 (blue line), RuL1 (black line) and (b) NdL4 (red line), RuL2-NdL4 (blue line), RuL2 (black line).

The characteristic infrared emission from the lanthanide counterpart is observed when excitation is set in the visible at the MLCT ruthenium(II) band (at 450 nm). For RuL1-NdL3, neodymium(III) 4F3/24I11/2 and 4F3/24I13/2 transitions at 1065 (sharp band) and 1334 nm (low intensity band) were detected as depicted in Figure 3a blue line.99 Lazarides, T.; Tart, N. M.; Sykes, D.; Faulkner, S.; Barbieri, A.; Ward, M. D.; Dalton Trans. 2009, 3971.,2525 Zhang, L.-Y.; Hou, Y.-J.; Pan, M.; Chen, L.; Zhu, Y.-X.; Yin, S.-Y.; Shao, G.; Su, C.-Y.; Dalton Trans. 2015, 44, 15212.,2626 Singaravadivel, S.; Babu, E.; Velayudham, M.; Lu, K.-L.; Rajagopal, S.; J. Organomet. Chem. 2013, 738, 49. The RuL2-NdL4 (Figure 3b, blue line) displayed similar transitions bands at 1066 (4F3/24I11/2) and 1334 nm (4F3/24I13/2), respectively.99 Lazarides, T.; Tart, N. M.; Sykes, D.; Faulkner, S.; Barbieri, A.; Ward, M. D.; Dalton Trans. 2009, 3971.,2525 Zhang, L.-Y.; Hou, Y.-J.; Pan, M.; Chen, L.; Zhu, Y.-X.; Yin, S.-Y.; Shao, G.; Su, C.-Y.; Dalton Trans. 2015, 44, 15212.,2626 Singaravadivel, S.; Babu, E.; Velayudham, M.; Lu, K.-L.; Rajagopal, S.; J. Organomet. Chem. 2013, 738, 49. Excitation spectra obtained by monitoring the NIR emission (λem: 1065 and 1066 nm for RuL1-NdL3 and RuL2-NdL4, respectively) are also shown in Figures 3a-3b, red lines. For both heterobimetallic complexes, three sharp bands were detected in visible range and ascribed to neodymium(III) intraconfigurational f-f transitions at 581, 746 and 800 nm.99 Lazarides, T.; Tart, N. M.; Sykes, D.; Faulkner, S.; Barbieri, A.; Ward, M. D.; Dalton Trans. 2009, 3971.,2626 Singaravadivel, S.; Babu, E.; Velayudham, M.; Lu, K.-L.; Rajagopal, S.; J. Organomet. Chem. 2013, 738, 49. The broad band from 270 to 400 nm were assigned to transitions centered at TTA-Si, bpy and bpmd ligands.1414 Fratini, A.; Richards, G.; Larder, E.; Swavey, S.; Inorg. Chem. 2008, 47, 1030.,1717 Duarte, A. P.; Gressier, M.; Menu, M.-J.; Dexpert-Ghys, J.; Caiut, J. M. A.; Ribeiro, S. J. L.; J. Phys. Chem. C. 2012, 116, 505.,1818 Cousinié, S.; Mauline, L.; Gressier, M.; Kandibanda, S. R.; Datas, L.; Reber, C.; Menu, M.-J.; New J. Chem. 2012, 36, 1355.,2626 Singaravadivel, S.; Babu, E.; Velayudham, M.; Lu, K.-L.; Rajagopal, S.; J. Organomet. Chem. 2013, 738, 49. Bands ranging from 400 to 600 nm were attributed to the ruthenium(II) 3MLCT transition.2525 Zhang, L.-Y.; Hou, Y.-J.; Pan, M.; Chen, L.; Zhu, Y.-X.; Yin, S.-Y.; Shao, G.; Su, C.-Y.; Dalton Trans. 2015, 44, 15212. Further information on the energy transfer from the ruthenium counterpart to the lanthanide one is these heterobimetallic compounds can be obtained by analyzing the changes observed for the Ru visible emission properties (Figures S10 and S11, SI section).

Figure 3
Room temperature excitation (red lines) and emission (blue lines) spectra from (a) RuL1-NdL3em: 1065 nm and λex: 450 nm) and (b) RuL2-NdL4em: 1066 nm and λex: 450 nm) complexes in solid state.

Table 1 shows results for excited state lifetimes. Considering bpmd as a planar bridging ligand that provides short metal-metal distance and a pathway that favors efficient energy transfer in heterobimetallic complexes, especially for neodymium(III) and ytterbium(III) ions,1010 Lazarides, T.; Adams, H.; Sykes, D.; Faulkner, G.; Calogero, A.; Ward, M. D.; Dalton Trans. 2008, 691. the lifetime values obtained for RuL1 and RuL2 substantially reduced in RuL1-NdL3 and RuL2-NdL4 complexes suggesting energy transfer processes for both heterobimetallic complexes. The kEnT and ηEnT may be evaluated from lifetimes as shown in Table 1.

Table 1
Photophysical properties for MLCT-based visible emission and neodymium(III) based NIR. Lifetimes data (τ) were obtained for ruthenium(II) and heterobimetallic complexes in solid state

A pathway of energy transfer (EnT) is presented in Scheme 2. The proposed energy level diagram suggests that the RuL2 and RuL1 3MLCT energy are suitable for the sensitization of the neodymium(III) luminescence. The RuL2 complex 3MLCT energy level is centered at 16129 cm-1 and the energy transfer to neodymium(III) could be evaluated. Considering the RuL2-NdL4 complex, the kEnT and ηEnT were 3.04 × 107 s-1and 84%, respectively. For RuL1 complex, the 3MLCT energy level is located at 14925 cm-1. Values of kEnT and ηEnT for the RuL1-NdL3 were 0.61 × 107 s-1 and 44%, respectively. It is worth emphasizing that higher EnT rate was observed for RuL2-NdL4 complex. This observation could be explained considering that neodymium(III) presents excited levels better located for energy transfer from RuL2 3MLCT states.

Scheme 2
Schematic energy transfer processes from RuL1 and RuL2 complexes (ruthenium(II)) to neodymium(III) (A = absorption; ISC = inter-system cross; P = phosphorescence; EnT = energy transfer; F = fluorescence) in the RuL1-NdL3 and RuL2-NdL4 complexes.

Conclusions

New ruthenium(II) complexes, named RuL1 and RuL2, both containing the bridging ligand bpym and the last one containing also a silylated bpy ligand (L2) have been prepared. From these complexes and new NdL3 diketonate complexes (where L3 is methoxysilyl modified diketonate), ruthenium(II)-neodymium(III) heteronuclear compounds have been prepared. A structural study was performed by EA, MS, FTIR and FT-Raman and 1D and 2D NMR techniques confirming the presence of the alkoxysilyl function and the fabrication of two new silylated d-f heterobimetallic complexes. UV-Vis spectra depicted characteristic transition bands in ruthenium(II) and neodymium(III) mononuclear precursors and both transition bands contributions in RuL1-NdL3 and RuL2-NdL4 complexes corroborating all structural analysis. The luminescent properties of the silylated d-f heterobimetallic complexes were evaluated with energy transfer processes being established from the ruthenium(II) donor to neodymium(III) acceptor units in the IR region. The decrease of lifetime and ruthenium quantum yield resulting from ruthenium(II) units compared to ruthenium(II)-neodymium(III) confirm effective energy transfer in these systems. For RuL1-NdL3 complex, kEnT and ηEnT values of 0.61 × 107 s-1 and 44% were obtained whereas RuL2-NdL4 complex showed kEnT and ηEnT values of 3.04 × 107 s-1and 84%, respectively. Our results emphasize that new silylated heterobimetallic ruthenium(II)-neodymium(III) complexes can be excited in the visible range via ruthenium(II) MLCT transitions promoting energy transfer to neodymium(III) units emitting in the IR region. The successful preparation of the silylated ruthenium(II)-neodymium(III) complexes allow grafting onto any silicated matrix suggesting their application as new luminescent NIR-emitting biosensors or biomarkers.

Supplementary Information

Supplementary data (1D/2D NMR data (Table S1); synthesis routes of the mononuclear precursors, RuL1, RuL2 and NdL3 (Figure S1); NdL4 synthesis route and structure (Figure S2); Raman scattering spectra (Figure S3); 1D/2D NMR spectra (Figures S4-S9); FTIR spectra (Figure S10); luminescence spectra monitoring Ru units of the heterobimetallic complexes (Figure S11) and of the ruthenium(II) precursors complexes (Figure S12)) are available free of charge at http://jbcs.sbq.org.br as PDF file.

Acknowledgments

This work was supported by the Brazilian agencies FAPESP, CNPq, CAPES and CAPES-COFECUB Brazil-France cooperation program for grant to R. M. Sábio.

References

  • 1
    Chen, F.-F.; Chen, Z.-Q.; Bian, Z.-Q.; Huang, C.-H.; Coord. Chem. Rev. 2010, 254, 991.
  • 2
    Chen, F.-F.; Wei, H.; Bian, Z.; Liu, Z.; Ma, E.; Chen, Z.; Huang, C.; Organometallics 2014, 33, 3275.
  • 3
    Jana, A.; Baggaley, E.; Amoroso, A.; Ward, M. D.; Chem. Commun. 2015, 51, 8833.
  • 4
    Baggaley, E.; Cao, D.-K.; Sykes, D.; Botchway, S. W.; Weinstein, J. A.; Ward, M. D.; Chem. - Eur. J. 2014, 20, 8898.
  • 5
    Pope, S. J. A.; Coe, B. J.; Faulkner, S.; Laye, R. H.; Dalton Trans 2005, 1482.
  • 6
    Li, L.; Zhang, S.; Xu, L.; Chen, Z.-N.; Luo, J.; J. Mater. Chem. C 2014, 2, 1698.
  • 7
    Lazarides, T.; Sykes, D.; Faulkner, S.; Barbieri, A.; Ward, M. D.; Chem. - Eur. J. 2008, 14, 9389.
  • 8
    Shavaleev, N. M.; Accorsi, G.; Virgili, D.; Bell, Z. R.; Lazarides, T.; Calogero, G.; Armaroli, N.; Ward, M. D.; Inorg. Chem. 2005, 44, 61.
  • 9
    Lazarides, T.; Tart, N. M.; Sykes, D.; Faulkner, S.; Barbieri, A.; Ward, M. D.; Dalton Trans 2009, 3971.
  • 10
    Lazarides, T.; Adams, H.; Sykes, D.; Faulkner, G.; Calogero, A.; Ward, M. D.; Dalton Trans 2008, 691.
  • 11
    Zucchi, G.; le Goff, X. F.; Polyhedron 2013, 52, 1262.
  • 12
    Zucchi, G.; Maury, O.; Thuéry, P.; Ephritikhine, M.; Inorg. Chem. 2008, 47, 10398.
  • 13
    Ward, M. D.; Coord. Chem. Rev. 2010, 254, 2634.
  • 14
    Fratini, A.; Richards, G.; Larder, E.; Swavey, S.; Inorg. Chem. 2008, 47, 1030.
  • 15
    Wei, Q.-H.; Lei, Y.-F.; Xu, W.-R.; Xie, J.-M.; Chen, G.-N.; Dalton Trans 2012, 41, 11219.
  • 16
    Cousinié, S.; Gressier, M.; Alphonse, P.; Menu, M.-J.; Chem. Mater. 2007, 19, 6492.
  • 17
    Duarte, A. P.; Gressier, M.; Menu, M.-J.; Dexpert-Ghys, J.; Caiut, J. M. A.; Ribeiro, S. J. L.; J. Phys. Chem. C. 2012, 116, 505.
  • 18
    Cousinié, S.; Mauline, L.; Gressier, M.; Kandibanda, S. R.; Datas, L.; Reber, C.; Menu, M.-J.; New J. Chem. 2012, 36, 1355.
  • 19
    Sábio, R. M.; Gressier, M.; Caiut, J. M. A.; Menu, M.-J.; Ribeiro, S. J. L.; Mater. Lett. 2016, 174, 1.
  • 20
    Cousinié, S.; Gressier, M.; Reber, C.; Dexpert-Ghys, J.; Menu, M.-J.; Langmuir 2008, 24, 6208.
  • 21
    Duarte, A. P.; Mauline, L.; Gressier, M.; Dexpert-Ghys, J.; Roques, C.; Caiut, J. M. A.; Deffune, E.; Maia, D. C. G.; Carlos, I. Z.; Ferreira, A. A. P.; Ribeiro, S. J. L.; Menu, M.-J.; Langmuir 2013, 29, 5878.
  • 22
    Mauline, L.; Gressier, M.; Roques, C.; Hammer, P.; Ribeiro, S. J. L.; Caiut, J. M. A.; Menu, M.-J.; Biofouling 2013, 29, 775.
  • 23
    Silverstein, D. W.; Milojevich, C. B.; Camden, J. P.; Jensen, L.; J. Phys. Chem. C 2013, 117, 20855.
  • 24
    Kokošková, M.; Procházka, M.; Šloufová, I.; Vlčková, B.; J. Phys. Chem. C 2013, 117, 1044.
  • 25
    Zhang, L.-Y.; Hou, Y.-J.; Pan, M.; Chen, L.; Zhu, Y.-X.; Yin, S.-Y.; Shao, G.; Su, C.-Y.; Dalton Trans 2015, 44, 15212.
  • 26
    Singaravadivel, S.; Babu, E.; Velayudham, M.; Lu, K.-L.; Rajagopal, S.; J. Organomet. Chem. 2013, 738, 49.

Publication Dates

  • Publication in this collection
    23 Mar 2020
  • Date of issue
    Apr 2020

History

  • Received
    23 Apr 2019
  • Accepted
    01 Oct 2019
Sociedade Brasileira de Química Instituto de Química - UNICAMP, Caixa Postal 6154, 13083-970 Campinas SP - Brazil, Tel./FAX.: +55 19 3521-3151 - São Paulo - SP - Brazil
E-mail: office@jbcs.sbq.org.br