首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Kim HJ  Jeong YC  Rhee JI 《Talanta》2008,76(5):1070-1076
To circumvent the leaching problem of optical sensing membranes used for dissolved oxygen (DO) measurements, the encapsulation of Ru(II) complexes linked with bulky dendron(s) in a sol-gel matrix was investigated. A dendron, readily formed via chemical transformations such as amidation and catalytic reduction, was covalently incorporated into tris(4,7-diphenyl-1,10-phenanthroline)ruthenium(II) complex, leading to an increase in the size and lipophilicity of dye molecules. Sol-gel-based sensing membranes encapsulating these Ru(II) complexes displayed a strong luminescence emission at 590 nm induced by radiation at 480 nm, and showed excellent DO sensing properties and stability for repeated measurements in aqueous solution. The encapsulation of the dendron-incorporated Ru(II) complexes in sol-gels prevented the dyes from leaching out of the membranes.  相似文献   

2.
Poe DP  Diehl H 《Talanta》1974,21(10):1065-1071
Tris(4,7-dihydroxy-1,10-phenanthroline)iron(II) reacts rapidly and quantitatively with dissolved oxygen in alkaline aqueous solution. In ammoniacal solution, the reaction is accompanied by the disappearance of the intense red colour of the iron(II) compound, which gives way to the pale gray, slightly-dissociated ion tris(4,7-dihydroxy-1,10-phenanthrolinefiron)(III). By measurement of the absorbance of a solution containing the ferrous compound before and after the injection of an oxygen-containing solution, the concentration of dissolved oxygen in the sample can be accurately determined in the range 1-20 ppm.  相似文献   

3.
Absorption at the excitation wavelength recovers in a sub-nanosecond, two stage process following bleaching of tris(1,10-phenanthroline) iron(II) by a single picosecond pulse at 530 nm. Absorption coefficients and decay times suggest that a CT and a dd excited state are consecutively occupied before ground state repopulation.  相似文献   

4.
《Polyhedron》1986,5(3):815-819
Stereoselective interactions between tris(acetylacetonato)cobalt(III) [Co(acac)3] and optically active Λ-(+)546-tris(1,10-phenanthroline)ruthenium(II) chloride [Λ-[Ru(phen)3]Cl2] were investigated by measuring the distribution coefficients of racemic Co(acac)3 between carbon tetrachloride and water containing Λ-[Ru(phen)3]2+ or pure water, and the optical rotations of the aqueous phase was measured over the temperature range 10–40° C. The Λ-isomer of Co(acac)3 is “salted-in” more strongly by Λ-[Ru(phen)3]2+ than the Δ-isomer. The association constants between Λ-[Ru(phen)]2+ and Λ- or Δ-Co(acac)3 were calculated by using the optical rotation to give K(Λ-Λ) = 3.86 and K(Λ-Δ) = 3.80 at 25°C, respectively. The temperature dependence of the association constants showed that the enthalpy and entropy changes for the Λ-Λ association is slightly more positive than those for the Λ-Δ association. This was discussed from the viewpoint of hydrophobic interactions.  相似文献   

5.
6.
The complex tris(4,7-diphenyl-1,10-phenanthroline)ruthenium(II), prepared via a simple microwave-assisted synthesis, was used to prepare a single-layer light-emitting electrochemical cell. This device reaches a high power efficiency of 1.9 Lum/W at a brightness of 390 cd/m2. Moreover, its lifetime is an order of magnitude longer than that of a similar cell making use of tris(bipyridine)ruthenium(II) as the emitting complex.  相似文献   

7.
Electrochemiluminescence (ECL) of ruthenium complexes has broad applications and the immobilization of Ru(bpy)32+ has received extensive attention. In comparison with Ru(bpy)32+, Ru(phen)32+ can be immobilized more easily because of its better adsorbability. In this study, immobilization of Ru(phen)32+ for ECL analysis has been demonstrated for the first time by using graphene oxide (GO) as an immobilization matrix. The immobilization of Ru(phen)32+ is achieved easily by mixing Ru(phen)32+ with GO without using any ion exchange polymer or covalent method. The strong binding of Ru(phen)32+ with GO is attributed to both the π–π stacking interaction and the electrostatic interaction. The Ru(phen)32+/GO modified electrode was characterized by using tripropylamine (TPA) as the coreactant. The linear range of TPA is from 3 × 10−7 to 3 × 10−2 mol L−1 with the detection limit of 3 × 10−7 mol L−1. The ECL sensor demonstrates outstanding long-term stability. After the storage in the ambient environment for 90 days, the ECL response remains comparable with its original signal.  相似文献   

8.
9.
10.
A new copper(II) complex, [Cu(epoxy)(phen)(NO3)](NO3)·H2O (epoxy = 5,6-dihydro-5,6-epoxy-1,10-phenanthroline, phen = 1,10-phenanthroline), has been synthesized and characterized by elemental analysis, IR and UV–vis spectroscopy, and single crystal X-ray diffraction. The complex crystallized in a monoclinic system with space group P21/n, a = 13.3199(8) Å, b = 11.2917(7) Å, c = 16.0832(9) Å, α = 90°, β = 107.827(4)°, and γ = 90°. The complex cation possesses a distorted octahedral geometry, with Jahn–Teller distortion occurring in the CuN4O2 core. The binding interaction with calf thymus DNA (CT–DNA) was investigated by UV–vis and fluorescence spectroscopy, cyclic voltammetry, and viscometry. The results support a partially intercalative binding mode. Efficient cleavage of plasmid DNA by the complex was observed by gel electrophoresis.  相似文献   

11.
Outer-sphere oxidation of phenols is under intense scrutiny because of questions related to the dynamics of proton-coupled electron transfer (PCET). Oxidation by cationic transition-metal complexes in aqueous solution presents special challenges because of the potential participation of the solvent as a proton acceptor and of the buffers as general base catalysts. Here we report that oxidation of phenol by a deficiency of [Os(phen)(3)](3+), as determined by stopped-flow spectrophotometry, yields a unique rate law that is second order in [osmium(III)] and [phenol] and inverse second order in [osmium(II)] and [H(+)]. A mechanism is inferred in which the phenoxyl radical is produced through a rapid PCET preequilibrium, followed by rate-limiting phenoxyl radical coupling. Marcus theory predicts that the rate of electron transfer from phenoxide to osmium(III) is fast enough to account for the rapid PCET preequilibrium, but it does not rule out the intervention of other pathways such as concerted proton-electron transfer or general base catalysis.  相似文献   

12.
A novel dinuclear ruthenium(II) complex [(phen)2Ru(mbpibH2)Ru(phen)2]4+ [phen = 1,10-phenanthroline; mbpibH2 = 1,3-bis(1,10-phenanthroline[5,6-d]imidazol-2-yl)-benzene] has been synthesized and characterized. The DNA-binding behavior of this complex has been studied by spectroscopic and viscosity measurements. Results suggest that the dinuclear ruthenium(II) complex intercalates into DNA base pairs via its bridging moiety. It has also been found that the dinuclear ruthenium(II) complex displays the enantiopreferential DNA-binding after equilibrium dialysis.  相似文献   

13.
A tripodal tris(urea) ligand with 2,2'-bipyridyl (bpy) substituents (L) has been designed and synthesized, which coordinates with three equivalents of Ru(bpy)(2)Cl(2)·2H(2)O, followed by treatment with NH(4)PF(6), to afford the anion receptor [(bpy)(6)Ru(3)L](PF(6))(6) (1). The anion-binding behavior of the ligand L and the Ru(II)-bpy functionalized receptor 1 toward different anions was investigated by (1)H NMR (for L and 1), fluorescence, and UV-vis spectroscopy (for 1). Both compounds showed selective recognition of SO(4)(2-) or H(2)PO(4)(-) ions in the 1:1 binding mode in the NMR studies. The Ru(II) complex 1 displayed the metal-to-ligand charge transfer emission at 600 nm, which was quenched on addition of the sulfate and dihydrogen phosphate ions. Quantitative fluorescence titration experiments were carried out and the stability constants (log K) of the complex 1 with SO(4)(2-) and H(2)PO(4)(-) ions were obtained to be 4.73 and 4.69 M(-1) (1:1 binding mode), respectively.  相似文献   

14.
The electrochemical behaviour of Ru(bipy)2(CN)2 and Ru(phen)2(CN)2 (bipy=2,2′-bipyridine; phen=1,10-phenanthroline) has been investigated in dimethylformamide. Both complexes exhibit one oxidation wave and three reduction waves. In the case of Ru(bipy)2-(CN)2 the anodic process and the first two cathodic processes involve one electron and are reversible in the time scale of polarographic and cyclic voltammetric experiments. The third reduction step is irreversible and has been attributed to the addition of three electrons to Ru(bipy)2(CN)2 followed by liberation of one or more ligands and reduction of liberated bipyridine. The features of the redox processes for the Ru(phen)2(CN)2 are similar to those found for the bipy complex except for the first reduction wave, which is complicated by adsorption phenomena. A qualitative MO discussion of the redox processes is also reported.  相似文献   

15.
The electrochemical behavior of 5-amino-1,10-phenanthroline and tris[5-amino-1,10-phenanthroline]-iron(II) at carbon paste, glassy carbon, and platinum electrodes is reported. The iron complex undergoes electrochemically induced oxidative polymerization from acetonitrile solutions and the resulting polymers are very stable. Charge transport through the polymer films occurs with a charge transfer diffusion coefficient, Dct, equal to 3.1 × 10−8 cm2 s−1 corresponding to an electron self-exchange rate of 5.2×107M−1 s−1. The activation energy and the entropy change for the charge transfer diffusion process are (approximate values) 32.0 ± 0.12 kJ mol−1 and −24.7 ± 0.4 J K−1 mol−1, respectively.  相似文献   

16.
17.
D. C. polarography and cyclic voltammetry were used for investigating the reduction processes of the tris(1,10-phenanthroline)cobalt(II) and bis(2,9-dimethyl-1, 10-phenanthroline)-cobalt(II) perchlorates in 0.1 M solutions of tetraethylammonium perchlorate in acetonitrile. The first complex gave a four-step reduction wave; the first two steps were found to be diffusion controlled and reversible reductions from Co(phen)+ to Co(phen)3+ to Co(phen) to Co(phen;) occured. The second complex gave a six-step reduction wave; the first three steps were found to be diffusion controlled and were to be considered as successive reversible reductions from Co(2, 9dm-phen)+ to Co(2, 9dmphen), from Co(2, 9dmphen) to Co(2, 9dmphen)2 and from Co(2, 9dmphen)2 to Co(2, 9dmphen).  相似文献   

18.
Previous studies have suggested that tris(4,7-diphenyl-1,10-phenanthrolinedisulfonate)ruthenium(II) (Ru(BPS)34−) has great potential as a chemiluminescence reagent in acidic aqueous solution. We have evaluated four different samples of this reagent (two commercially available and two synthesised in our laboratory) in comparison with tris(2,2′-bipyridine)ruthenium(II) (Ru(bipy)32+) and tris(1,10-phenanthroline)ruthenium(II) (Ru(phen)32+), using a range of structurally diverse analytes. In general, Ru(BPS)34− produced more intense chemiluminescence, but the oxidised Ru(BPS)33− species is less stable in aqueous solution than Ru(bipy)33+ and produced a greater blank signal than Ru(bipy)33+ or Ru(phen)33+, which had a detrimental effect on sensitivity. Although the complex is often depicted with the sulfonate groups of the BPS ligand in the para position on the phenyl rings, NMR characterisation revealed that the commercially available BPS material used in this study was predominantly the meta isomer.  相似文献   

19.
20.
The redox properties of a series of [Ru(phen)2(py)X]n+ cations (X = pyridine, NH3, Cl, Br, I, CN, SCN, N3 and NO2) have been investigated in acctonitrile. Two reversible reduction steps are seen at ? 1.35 and ? 1.6 V vs Ag/AgCl; the invariance of these processes with X-group is indicative of electron addition to molecular orbitals mainly of phenanthroline ligand π* origin. Irreversible multi-electron reductions follow below ? 2.20 V. The Ru(II)/Ru(III) couple is seen as a reversible wave near + 0.8 V vs the normal hydrogen electrode, from calibration with ferrocene, except in the cases of the NO2 and SCN complexes, where rapid reactions involving these ligands occur.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号