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.
2.
Two configurationally stable, chiral anions (TRISPHAT, 1 ) behave as efficient hosts that control the configuration of a configurationally labile iron(II ) complex as the guest with high diastereoselectivity (>96 % de) upon ion pairing. The diastereoselectivity increases with decreasing solvent polarity.  相似文献   

3.
A light-driven system consisting of tris(2,2′-bipyridine)ruthenium(II) (Ru(bpy)32+) as the photosensitizer, semicarbazide as the electron donor and molecular oxygen as the electron acceptor has been employed for hydrogen peroxide production. The efficiency of this photosystem markedly depends on pH: while the peroxide yield is almost negligible at acid, neutral or slightly alkaline pH, it reaches significant values at high hydroxide concentrations, the initial rate of H2O2 formation drastically increasing from pH 12 to pH 14. In 1 M NaOH solutions containing Ru(bpy)32+ and semicarbazide at optimum concentrations, the number of catalytic cycles (or turnover number) undergone by the ruthenium complex over the complete course of the photochemical reaction is as high as 1.1 × 104.

Spectrofluorometric and laser flash photolysis techniques were used to study the primary photochemical reactions involving the excited state of the ruthenium complex as well as the photochemically generated species Ru(bpy)33+ and Ru(bpy)3+. It is proposed that at pH 14 a sequence of reactions leading to O2 photoreduction by electrons from semicarbazide takes place, with the concomitant formation of H2O2; the excited state of Ru(bpy)32+ appears to react via oxidative quenching by oxygen rather than via reductive quenching by semicarbazide. At neutral pH, in contrast, there is no H2O2 formation owing to the fact that semicarbazide is unable to reduce (Ru(bpy)33+ to Ru(bpy)32+, although the photoexcited ruthenium complex is quenched equally by oxygen.  相似文献   


4.
A study is presented of the quenching, by oxygen, of the luminescence of tris(2,2′-bipyridine) ruthenium(II) complexes immobilized in thin, transparent, polymer-based films. The film media consist of a water-insoluble linear polymer plasticized with a trialkylphosphate ester, in which the complex ruthenium cations are solubilized by ion pairing with organophilic anions such as tetraphenylborate.

Luminescence lifetimes were studied in relation oxygen concentration in a gas stream contiguous with the film medium, film thickness and concentration of the metal complex within the film medium. It is shown that the microheterogeneous environment of the luminescent complex, which has recently been implicated in the non-linear quenching responses of polymer-immobilized, transition metal complex oxygen sensors, may arise simply as a consequence of the limited solubility of the complex in the film medium. When solubility is limited, the partial precipitation of the complex results in a colloidal of luminescent particles which exhibit non- uniform susceptibilities to quenching by oxygen. Good solubility, and therefore linear quenching characteristics, are promoted by methyl substitution of the bipyridyl ligand and by use of a plasticizer (tributylphosphate) with marked cation solvating powers.  相似文献   


5.
Polysiloxanes containing pendant tris(2,2′-bipyridine)ruthenium(II) complex (Ru(bpy)32+) were prepared by reaction of polysiloxane-pendant 2,2′-bipyridine (PSiO-bpy) with cis-Ru(bpy)2Cl2. In methanol solution, the polymer pendant Ru(bpy)32+ showed absorption maximum at 456nm and emission maximum at around 609nm, both of which are shifted to longer wavelength than the monomeric Ru(bpy)32+. The lifetime τ0 of the excited polymer complex with low Ru(bpy)32+ content was almost the same as that of the monomeric one in methanol (830ns), but τ0 of the polymer with higher complex content was shorter because of a concentration quenching. In a solid state, τ0 was much shorter (306–503ns) than that in a methanol solution contrary to the conventional polymeric system. Higher complex content in the polymer film caused higher glass transition temperature (Tg), but shorter τ0. These results indicate concentration quenching in the polymer film. The excited polymer pendant Ru(bpy)32+ was quenched by oxygen, and the relative emission intensity followed the Stern-Volmer equation. In a methanol solution the quenching rate constant (kq) was the same order of magnitude as the monomeric complex, and independent of the complex content in the polymer. In a film, kq was higher for the polymer with higher complex content.  相似文献   

6.
The NMR method has been used to study the structure of the complexes [Cd(bipy)]SO4.4H2O, [Cd(bipy)](NO3)2.2H2O, [Cd(bipy)2](NO3)2.12H2O and [Cd(bipy)3](NO3)2.7H2O. The influence of the central ion and of diamagnetic currents of the rings in these complexes on the PMR spectrum has been investigated. In the complexes [Cd(bipy)](NO3)2.2H2O and [Cd(bipy)]SO4.4H2O two kinds of hydration isomers, with different PMR spectra, have been obtained.  相似文献   

7.
The temperature dependence of the emission lifetime of the series of complexes Ru(bpy)n(4,4′-dpb) (bpy = 2,2′bipyridine, 4,4′-dpb = 4,4′-diphenyl-2,2′-bipyridine) has been studied in propionitrile/butyronitrile (4:5 v/v) solutions in the range 90–293 K. The obtained photophysical parameters show that the energy separation between the metal-to-ligand charge tranfer (3MLCT) emitting level and the photoreactive metal-centered (3MC) level changes across the series (ΔE = 3960, 4100, 4300, and 4700 cm?1 for Ru(bpy)), Ru(bpy)2(4,4′-dpb)2+, Ru(bpy)(4,4′-dpb), and Ru(4,4′-dpb), respectively, where ΔE is the energy separation between the minimum of the 3MLCT potential curve and 3MLCT – 3MC crossing point. Comparison between spectral and electrochemical data indicated that the changes in ΔE are due to stabilization of the MLCT levels in complexes containing 4,4′-dpb with respect to Ru(bpy)2+3. The photochemical data for the same complexes (as I? salts) have been obtained in CH2Cl2 in the presence of 0.01M Cl? upon irradiation at 462 nm. The complexes containing 4,4′-dpb are more photostable than Ru(bpy). Comparison between the data for thermal population of the 3MC photoreactive state and those for photochemistry indicated that the overall photochemical process is governed by (i) a thermal redistribution between the emitting and photoreactive excited states, and (ii) mechanistic factors, likely related to the size of the detaching ligand.  相似文献   

8.
The photophysics of three complexes of the form Ru(bpy)3−(pypm)2+ (where bpy2,2′-bipyridine, pypm 2-(2′-pyridyl)pyrimidine and P=1, 2 or 3) was examined in H2O, propylene carbonate, CH3CN and 4:1 (v/v) C2H5OH---CH3OH; comparison was made with the well-known photophysical behavior of Ru(bpy)32+. The lifetimes of the luminescent metal-to-ligand charge transfer (MLCT) excited states were determined as a function of temperature (between −103 and 90 °C, depending on the solvent), from which were extracted the rate constants for radiative and non-radiative decay and ΔE, the energy gap between the MLCT and metal-centered (MC) excited states. The results indicate that *Ru(bpy)2(pypm)2+ decays via a higher lying MLCT state, whereas *Ru(pypm)32+ and *Ru(pypm)2(bpy)2+ decay predominantly via the MC state.  相似文献   

9.
The electrogenerated chemiluminescence (ECL) of the Ru(bpy)32+ (bpy, 2,2′-bipyridine)/tri-n-propylamine (TPrA) system can be produced at an oxidation-potential well before the oxidation of Ru(bpy)32+. Here, we describe the unique features of the low-oxidation-potential (LOP) ECL. The LOP ECL exhibited strong dependence on solution pH with the maximum emission at pH  7.7. Compared with the conventional ECL, the LOP ECL was much more significantly diminished at high pH (>10), probably due to the short lifetime of TPrA cation radical which is a crucial intermediate for the LOP emission. It was also found that the preceding deprotonation step played an important role in TPrA oxidation at neutral pH and would remarkably influence the emission intensity. As excess intermediate radicals were produced upon rapid TPrA oxidation, only 5 mM TPrA was needed to achieve the maximum LOP ECL intensity in detecting trace Ru(bpy)32+ (<1 μM) and the LOP ECL response to Ru(bpy)32+ concentration was linear. Compared with the conventional Ru(bpy)32+/TPrA ECL, the LOP ECL technique not only produces higher emission intensity at lower oxidation-potential, but also significantly reduces the amount of the coreactant.  相似文献   

10.
11.
The complexes Zn(bipy)Cl2 and Zn(bipy)2Cl2 as well as 2,2′-bipyridyl in aqueous solution (D2O) have been examined by the NMR method. The presence of the monocationic bipy D+ form in aqueous bipyridyl solution has been found. The changes of chemical shifts of bipyridyl protons for complexes Zn(bipy)3Cl2 and Zn(bipy)Cl2 have confirmed explicitly the essential influence of diamagnetic currents on the NMR spectrum of Zn(bipy)3Cl2. The comparison of the spectra of 2,2′-bipyridyl (in CH3OH) and of Zn(bipy)Cl2 may also suggest the presence of the nonbonding metal-proton 6 interaction.  相似文献   

12.
A post- column chemiluminescent technique for thedetection of compounds that are poor chromoshores using electorogenerated chemiluminescence following separation by capillartgy electrophoresis is described. The luminrescent signal is generated followintg the reaction of anlyres (e.g. amines) with Ru(bpy)33+, which isx electrochemically generated post-columan from Ru(bpy)32+. Tripropylamine and proline are used as two model compounds to demostrate the feasibility of the method. Detection limits for the prototype system were in the micromolar rage, suggesting that this technnique offers an alternative to indirect detection of compounds that are poor chromophores with an added selectivity advangage. The system includes the use of a conductive joint to isolate the separation field from the potential necessary to drive the elctrogenerated chemiluminescent reactiion. Addition of the chemiluminescent reagent Ru(bpy)32+ post-column did not decrease the efficiency of the separation. The design and favrication of the novel cell is discussed.  相似文献   

13.
The pyrocatechol derivatives, dopamine ( 2 ), adrenaline ( 3 ) and L -dopa (4) form in their semiquinoid oxidation states complexes with several closed-shell metal ions like Zn2+, Cd2+ and Y3+. This complex formation can be triggered by visible light via a one-electron transfer reaction from the pyrocatechol derivative to (3CT)Ru(bpy), which is thereby reduced (Schemes 1 and 2), The quenching of the triplet charge-transfer state, (3CT)Ru(bpy) cannot be measured by conventional methods (Stern-Volmer plot), because the bimolecular electron-transfer rate is too slow (kq < 10?7 M?1 s?1). ESR. spectroscopy shows, however, clearly the paramagnetic reaction products. The concentration of these paramagnetic species is strongly enhanced by complex formation. The hindered rotation of the side chain in L -dopa and in its Y3+-complex is described within a two-jump model, which gives best agreement between calculated and measured spectra assuming an energy of 37 kJmol?1 for the rotational barrier.  相似文献   

14.
《Electroanalysis》2006,18(16):1584-1589
We described here a new method for the determination of total calcium in plasma. The method is based on the precipitation of calcium with excess oxalate and the measurement of residual oxalate by flow injection analysis with Ru(bpy)32+ electrochemiluminescent detection. It has the advantages of extremely stable reagent, user‐friendly instrument, high selectivity, good analytical recovery, wide dynamic range, and nice correlation with atomic absorption spectroscopy. The calibration plot for calcium is linear over a concentration range from 0.5 mmol L?1 to 4.8 mmol L?1, which is wider than those obtained by most other methods. The analytical recoveries for plasma calcium are 98.4–101.2% with coefficients of variation (CVs) of 1.96–2.52%. The within‐day CVs range from 0.76% to 0.95%, and between‐day CVs were from 1.12% to 1.46%. The time for each injection is one minute. Because the proposed method can be readily carried out on increasingly popular instruments for Ru(bpy)32+ ECL immunoassays and DNA probe assays, Ru(bpy)32+ ECL method is suitable for routine clinical analysis of calcium.  相似文献   

15.
16.
The structures of 3,3′-dicarbometoxy-2,2′-bipyridine (dcmbpy) complexes with copper(II) and silver(I) cations have been determined using single crystal X-ray-diffraction. The crystals of Cu(dcmbpy)Cl2 are monoclinic, C2/c, a = 16.966(3), b = 18.373(3), c = 13.154(2) Å, β = 126.543(3)°. The crystals of Ag(dcmbpy)NO3 · H2O are also monoclinic, C2/c, a = 16.7547(13), b = 11.0922(9), c = 18.7789(18) Å, β = 100.228(7)°. The results have been compared with the literature data on the complexes of dcmbpy and its precursors: 2,2′-bipyridine (bpy) and 3,3′-dicarboxy-2,2′-bipyridine (dcbpy). Two types of complexes of 3,3′-carboxy derivatives of bpy are distinguished: (1) with metal atom bonded to two N atoms of the same molecule and (2) with metal atom bonded to two N atoms of two different molecules. The Cu(dcmbpy)Cl2 complex belongs to the first type, whereas Ag(dcmbpy)NO3 · H2O belongs to the second type.  相似文献   

17.
For the first time, CEC was coupled with tris(2,2‐bipyridyl) ruthenium(II) ( Ru(bpy) electrochemiluminescence detection. Efficient CEC separations of proline, putrescine, spermidine and spermine were achieved when the pH of the mobile phase is in the range of 3.5–7.0. The optimum mobile phase for CEC separation is much less acidic than that for CZE separation, which matches better with the optimum pH for Ru(bpy) electrochemiluminescence detection and dramatically shortens the analysis time because of larger EOF at higher pH. The time for CEC separation of the polyamines is less than 12.5 min, which is about half as much as the time needed for CZE. The detection limits were 1.7, 0.2, and 0.2 μM for putrescine, spermidine, and spermine, respectively. The RSD of retention time and peak height of these polyamines were less than 0.85 and 6.1%, respectively. The column showed good long‐term stability, and the RSD of retention time is below 5% for 150 runs over one‐month use. The method was successfully used for the determination of polyamines in urine samples.  相似文献   

18.
19.
20.
Electrochemiluminescence (ECL) of tris(2,2′‐bipyridine)ruthenium, Ru(bpy)32+ in the presence of various co‐reactants, such as tripropylamine (TPA), oxalate ion (C2O42?), ascorbic acid (H2A) and dehydroascorbic acid (DHA), were investigated under ultrasound irradiation. In sono‐ECL experiments, an indium‐thin‐oxide (ITO) was used as working electrode, and a titanium tipped sonic horn probe (diameter 2 mm) which operated at a frequency of 20 kHz was set in the front of the ITO electrode. Under the ultrasound irradiation, ECL signals were found to be significantly enhanced when TPA and C2O42? were used as co‐reactants, only slightly enhanced in Ru(bpy)32+/DHA system, but total quenched in Ru(bpy)32+/H2A system. The difference of Ru(bpy)32+ ECL behaviors for various co‐reactant could to be due to the different kinetics of catalytic reactions associated in ECL schemes. ECL quenching effect observed in Ru(bpy)32+/H2A system was suggested to be due to electron transfer (ET) route between the excited state *Ru(bpy)32+ and ascorbate anion HA? diffused from the bulk solution, where the diffusional HA? species served as electron donor. The effect becomes more pronounced upon sonication because the effective collision frequency between *Ru(bpy)32+ and HA? would be significantly increased by the enhanced mass transport effect of ultrasound.  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