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 31 毫秒
1.
The spectroscopic and photophysical properties of a synthetically versatile ruthenium complex [Ru(bpy)2(LH2)]2+ where LH2 is 2-(4-carboxyphenyl)imidazo[4,5-f][1,10]phenanthroline and bpy is 2,2-bipyridyl and its analogue, [Ru(bpy)2(LOMe)]2+ where the carboxyphenyl functionality is methylated are reported. Both complexes exhibit long-lived luminescence which for [Ru(bpy)2(LH2)]2+ is remarkably enhanced in aqueous compared to organic media. The pH dependence of the electronic absorption and emission spectra in water and acetonitrile are described and the influence of the protonation state of the 2-(4-carboxyphenyl)imidazo[4,5-f][1,10]phenanthroline ligand on the electronic structure of [Ru(bpy)2(LH2)]2+ is discussed. Oxidative quenching of the excited state of the complex by anthraquinone-2-carboxylic acid is investigated for both complexes. In polar media, this is a dynamic process suggesting that the quenching rate is controlled by bimolecular collision with a quenching rate constant, kq, of approximately 6.7 × 109 M−1 s−1 for [Ru(bpy)2(LH2)]2+. In contrast in aprotic solvent, dichloromethane, quenching occurs through a purely static mechanism indicating association between the luminophore and quencher, most likely through hydrogen bonding, between the carboxylic acid moieties of the ruthenium complex and the anthraquinone carboxylic derivative. The association constant for formation of the dyad was determined to be 565 L mol−1 in dichloromethane and the rate of electron transfer was estimated to be 4.7 × 107 s−1. By contrast, for the analogous complex in which the carboxylate is methyl protected mixed static and dynamic quenching behaviour in aprotic solvent.  相似文献   

2.
New ruthenium(II) complexes carrying methionine and phenylalanine in the bipyridine ligand, [Ru(bpy)2(4-Me-4′-(CONH-l-methionine methyl ester)-2,2′-bipyridine)](PF6)2 (IV) and [Ru(bpy)2(4-Me-4′-(CONH-l-phenylalanine ethyl ester)-2,2′-bpy)](PF6)2(V) have been synthesized and characterized and their photophysical properties studied. Flash photolysis measurements of complex IV, in the presence of an electron acceptor, methyl viologen (MV2+) show that an intermolecular electron transfer from the excited state of Ru(II) in complex IV, to MV2+ takes place, forming Ru(III) and the methyl viologen cation radical, MV+. The formation of MV+ in this system is confirmed using time-resolved transient absorption spectroscopy. This intermolecular electron transfer is followed by intramolecular electron transfer from the thioether moiety (methionine) to the photogenerated Ru(III), regenerating Ru(II).  相似文献   

3.
The reaction of [Ru(salen)(PPh3)Cl] and the 5-imidazol-substituted nitronyl nitroxide radical (NIT-(5)ImH) yields the [Ru(salen)(PPh3)(NIT-(5)ImH)](ClO4) (1) complex which has been characterized by single crystal X-ray diffraction. This analysis reveals that the Ru(III) ion is coordinated to a tetradentate salen2? ligand in equatorial positions while one PPh3 ligand and one NIT-(5)ImH radical are coordinated in axial positions. This led to RuIII ions in tetragonally elongated octahedral geometry. From the magnetic point of view ferromagnetic intramolecular interaction (J1 = +2.47 cm?1) have been found between the Ru(III) ion and the coordinated NIT-(5)ImH while no significant intermolecular antiferromagnetic interactions are observed at low temperature leading to a ground spin state S = 1. The absence of intermolecular magnetic interaction is explained by considering the crystal packing of (1) where the [Ru(salen)(PPh3)(NIT-(5)ImH)]+ moieties are relatively well isolated. This has to be compared with the situation observed in the previously reported [Ru(salen)(PPh3)-(NIT)]+ compound (2) where ferromagnetic RuIII–NIT interaction were identified and the crystal packing generate intermolecular antiferromagnetic interactions that complicated the study. The analysis of this compound confirms the rather isotropic g values that were found of (2) and of [Ru(salen)(PPh3)(N3)], (3) a radical-free analogue. Moreover it is also a step towards extended structures based on RuIII–NIT moieties since this compound possesses a free bischelating site likely to coordinate additional metallic ions.  相似文献   

4.
A new silver(II) complex, {[Ag(L1)](NO3)2·4H2O}n (1) (L = 3,14-dimethyl-2,6,13,17-tetraazatricyclo[14,4,01.18,07.12] docosane) has been synthesized and structurally characterized by a combination of analytical, spectroscopic, electrochemical and X-ray diffraction methods. The complex 1 exhibits a 1D supramolecular polymer with the silver(II) macrocycle L1 and nitrate ions, where 1D chain is formed by hydrogen bonds between the two sets of pre-organized N-H groups of the macrocycle and nitrate ions. The lattice water molecules mediate to interconnect each 1D chain to form the 2D supramolecular sheet. In 1 the unusual high oxidation state of Ag(II) is stabilized by the tetraazamacrocyclic ligand L1. The cyclic voltammogram for 1 indicates that the electrochemical oxidation of [Ag(L1)]2+ is an irreversible process.  相似文献   

5.
《Inorganica chimica acta》1986,115(2):193-196
The binuclear complexes [Cl(OC)3ReI(bipym)ReI(CO)3Cl] (bipym=2,2′-bipyrimidine), [(bipy)2RuII(bipym)ReI(CO)3Cl](PF6)2 (bipy=2,2′-bipyridine) and their mononuclear component [Re(bipym)(CO)3Cl] were prepared. The electronic absorption spectra of these complexes display low-energy Re(I) →π*(bipym) and Ru(II)→π*(bipym) charge transfer (CT) bands. While [Re(bipym)(CO)3Cl] shows a strong emission from its lowest CT state, the dimer [Cl(OC)3Re(bipym)Re(CO)3Cl] is not luminescent. The cation [(bipy)2Ru(bipym)Re(CO)3Cl]2+ emits from the lowest-energy Ru→bipym CT state. The emission behavior of the binuclear complexes is described in terms of intramolecular excited state electron or energy transfer.  相似文献   

6.
Two new ruthenium complexes [Ru(bpy)2(mitatp)](ClO4)21 and [Ru(bpy)2(nitatp)](ClO4)22 (bpy = 2,2′-bipyridine, mitatp = 5-methoxy-isatino[1,2-b]-1,4,8,9-tetraazatriphenylene, nitatp = 5-nitro-isatino[1,2-b]-1,4,8,9-tetraazatriphenylene) have been synthesized and characterized by elemental analysis, 1H NMR, mass spectrometry and cyclic voltammetry. Spectroscopic and viscosity measurements proved that the two Ru(II) complexes intercalate DNA with larger binding constants than that of [Ru(bpy)2(dppz)]2+ (dppz = dipyrido[3,2-a:2′,3′-c]phenazine) and possess the excited lifetime of microsecond scale upon binding to DNA. Both complexes can efficiently photocleave pBR322 DNA in vitro under irradiation. Singlet oxygen (1O2) was proved to contribute to the DNA photocleavage process, the 1O2 quantum yields was determined to be 0.43 and 0.36 for 1 and 2, respectively. Moreover, a photoinduced electron transfer mechanism was also found to be involved in the DNA cleavage process.  相似文献   

7.
A novel ruthenium(II) complex of dipyridophenazine (DPPZ) with the ancillary ligand imidazole[4,5-f] [1,10]phenanthroline (IP), [Ru(IP)2(DPPZ)] (PF6)2, has been synthesized and characterized by elemental analysis, 1D and 2D 1H NMR, fast-atom bombardment mass spectra (FABMS), electronic spectroscopy and cyclic voltammetry. The DNA-binding properties of the complex were studied by spectroscopic methods. The intrinsic binding constant, K =2.1 × 107M−1, of the complex to calf thymus DNA has been determined by absorption titration in 5 mmol dm−3 Tris-HCl, 50 mmol dm−3 NaCl buffer (pH 7.0). The excited state lifetimes and luminescence quenching with [Fe(CN)6]4− as the quencher in the presence of DNA were also tested and mono-exponentiality was observed for the emission decay curves. Viscosity measurements together with the optical titrations unambiguously proved that the complex bound with DNA intercalatively and that the binding affinity to DNA was several times larger than that of the parent complex [Ru(bpy)2(DPPZ)]2+.  相似文献   

8.
The trinuclear [{RuII(bpy)2(bpy-terpy)}2CoII]6+ complex (16+) in which a Co(II)-bis-terpyridine-like centre is covalently linked to two Ru(II)-tris-bipyridine-like moieties by a bridging bipyridine-terpyridine ligand has been synthesised and characterised. Its electrochemical, photophysical and photochemical properties have been investigated in CH3CN. The cyclic voltammetry exhibits two successive reversible oxidation processes, corresponding to the CoIII/CoII and RuIII/RuII redox couples at E1/2 = −0.06 and 0.91 V vs Ag/Ag+ 10 mM, respectively. The one-electron oxidized form of the complex, [{RuII(bpy)2(bpy-terpy)}2CoIII]7+ (17+) obtained after exhaustive electrolysis carried out at 0.2 V is fully stable. 16+ and 17+ are only poorly luminescent, indicating that the covalent linkage of the Ru(II)-tris-bipyridine centre to the cobalt subunit leads to a strong quenching of the RuII excited state by an intramolecular process. Luminescence lifetime experiments carried out at different temperatures indicate that the transfer is more efficient for 17+ compare to 16+ due to lower activation energy. Continuous irradiation of 17+ performed at 405 nm in the presence of P(Ph)3 acting as sacrificial electron donor leads to its quantitative reduction into 16+, whereas similar experiment starting from 16+ with a sulfonium salt as sacrificial electron acceptor converts 16+ into 17+ with a slower rate and a maximum yield of 80%. These photoinduced electron transfers were followed by UV-Visible spectroscopy and compared with those obtained with a simple mixture of both mononuclear parent complexes i.e. [RuII(bpy)3]2+ and [CoII(tolyl-terpy)2]2+ or [CoIII(tolyl-terpy)2]3+ (tolyl-terpy = 4′-(4-methylphenyl)-2,2′:6′,2′′-terpyridine).  相似文献   

9.
In order to explore the electronic effects of Ru(II) complexes binding to DNA, a series of Ru(II) complexes [Ru(phen)2 (p-MOPIP)]2+ (1), [Ru(phen)2 (p-HPIP)]2+ (2), and [Ru(phen)2(p-NPIP)]2+ (3) were synthesized and characterized by elementary, 1H NMR, and ES-MS analysis. The binding properties of these complexes to CT-DNA were investigated with spectroscopic methods and viscosity experiments. Furthermore, the computations for these complexes applying the density functional theory (DFT) method have also been performed. The results show that all of these complexes can well bind to DNA in intercalation mode and DNA-binding affinity of these complexes is greatly influenced by electronic effects of intercalating ligands. The intrinsic binding constants for 1, 2, and 3 are 0.20, 0.69, and 1.56 × 105 M−1, respectively. This order is in accordance with that of the electron-withdrawing ability of substituent [-OR < -OH < -NO2]. Such a trend in electronic effects of Ru(II) complexes binding to DNA can be reasonably explained by the DFT calculations.  相似文献   

10.
The new complex, [RuII(bpy)2(4-HCOO-4′-pyCH2 NHCO-bpy)](PF6)2 · 3H2O (1), where 4-HCOO-4′-pyCH2NHCO-bpy is 4-(carboxylic acid)-4′-pyrid-2-ylmethylamido-2,2′-bipyridine, has been synthesised from [Ru(bpy)2(H2dcbpy)](PF6)2 (H2dcbpy is 4,4′-(dicarboxylic acid)-2,2′-bipyridine) and characterised by elemental analysis and spectroscopic methods. An X-ray crystal structure determination of the trihydrate of the [Ru(bpy)2(H2dcbpy)](PF6)2 precursor is reported, since it represented a different solvate to an existing structure. The structure shows a distorted octahedral arrangement of the ligands around the ruthenium(II) centre and is consistent with the carboxyl groups being protonated. A comparative study of the electrochemical and photophysical properties of [RuII(bpy)2(4-HCOO-4′-pyCH2NHCO-bpy)]2+ (1), [Ru(bpy)2(H2dcbpy)]2+ (2), [Ru(bpy)3]2+ (3), [Ru(bpy)2Cl2] (4) and [Ru(bpy)2Cl2]+ (5) was then undertaken to determine their variation upon changing the ligands occupying two of the six ruthenium(II) coordination sites. The ruthenium(II) complexes exhibit intense ligand centred (LC) transition bands in the UV region, and broad MLCT bands in the visible region. The ruthenium(III) complex, 5, displayed overlapping LC bands in the UV region and a LMCT band in the visible. 1, 2 and 3 were found, via cyclic voltammetry at a glassy carbon electrode, to exhibit very positive reversible formal potentials of 996, 992 and 893 mV (versus Fc/Fc+) respectively for the Ru(III)/Ru(II) half-cell reaction. As expected the reversible potential derived from oxidation of 4 (−77 mV (versus Fc/Fc+)) was in excellent agreement with that found via reduction of 5 (−84 mV (versus Fc/Fc+)). Spectroelectrochemical experiments in an optically transparent thin-layer electrochemical cell configuration allowed UV-Vis spectra of the Ru(III) redox state to be obtained for 1, 2, 3 and 4 and also confirmed that 5 was the product of oxidative bulk electrolysis of 4. These spectrochemical measurements also confirmed that the oxidation of all Ru(II) complexes and reduction of the corresponding Ru(III) complex are fully reversible in both the chemical and electrochemical senses.  相似文献   

11.
A new ligand L1 has been prepared in which two 1,10-phenanthroline fragments are separated by an 18-crown-6 macrocyclic spacer. This was used to prepare the heterodinuclear complex [(bipy)2Ru(μ-L1)Re(CO)3Cl][PF6]2 [Ru(L1)Re] in which the {Ru(bipy)2(phen)}2+ and {Re(CO)Cl(phen)} chromophores are separated by a saturated and fairly flexible crown-ether fragment. On the basis of photophysical studies on Ru(L1)Re and associated mononuclear Ru(II) and Re(I) complexes, Re → Ru photoinduced energy-transfer occurs with a rate constant of 1.9 × 108 s−1 in solution room temperature leading to near-complete quenching of the Re(I)-based luminescence. At 77 K the Re(I)-based luminescence component is completely quenched. Calculations on the efficiency of both Förster and Dexter energy-transfer as a function of Re?Ru distance in this system suggest that a folded conformation of the complex, in which the Re?Ru separation is much shorter than that implied by the extended conformation detected crystallographically, is responsible for the energy-transfer, since neither Förster nor Dexter Re → Ru energy-transfer should be possible with the complex in an extended conformation. Addition of K+ or Ba2+ salts to solutions of Ru(L1)Re had no effect on the photophysical properties, probably because the association constants are too low to give significant metal-ion binding in the macrocycle at the low concentrations employed.  相似文献   

12.
《Inorganica chimica acta》1988,149(2):177-185
CpRuCl(PPh3)2 reacted with excess R-DAB in refluxing toluene to give CpRuCl(R-DAB(4e)) (1a: R = i-Pr; 1b: R = t-Bu; 1c: R = neo-Pent; 1d: R =p-Tol). 1H NMR and 13C NMR spectroscopic data indicated that in these complexes the R-DAB ligand is bonded in a chelating 4e coordination mode.Reaction of 1a and 1b with one equivalent of [Co(CO)4] afforded CpRuCo(CO)3(R-DAB(6e)) (2a: R = i-Pr; 2b: R = t-Bu). The structure of 2b was determined by a single crystal X-ray structure determination. Crystals of 2b are monoclinic, space group P21/n, with four molecules in a unit cell of dimensions: a = 16.812(4), b = 12.233(3), c = 9.938(3) Å and β = 105.47(3)°. The structure was solved via the heavy atom method and refined to R = 0.060 and Rw = 0.065 for the 3706 observed reflections. The molecule contains a RuCo bond of 2.660(3) Å and a cyclopentadienyl group that is η5-coordinated to ruthenium [RuC(cyclopentadienyl) = 2.208(3) Å (mean)]. Two carbonyls are terminally coordinated to cobalt (CoC(1) = 1.746(7) and CoC(2) = 1.715(6) Å) while the third is slightly asymmetrically bridging the RuCo bond (RuC(3) = 2.025(6) and CoC(3) = 1.912(6) Å). The RuC(3)O(3) and CoC(3)O(3) angles are 138.4(5)° and 136.5(5)°, respectively. The t-Bu-DAB ligand is in the bridging 6e coordination mode: σ-N coordinated to Ru (RuN(2) = 2.125(4) Å), μ2-N′ bridging the RuCo bond and η2-CN coordinated to Co (RuN(1) = 2.113(5), CoN(1) = 1.941(4) and CoC(4) = 2.084(5) Å). The η2-CN′ bonded imine group has a bond length of 1.394(7) Å indicating substantial π-backbonding from Co into the anti-bonding orbital of this CN bond.1H NMR spectroscopy indicated that 2a and 2b are fluxional on the NMR time scale. The fluxionality of 6e bonded R-DAB ligands is rarely observed and may be explained by the reversible interchange of the σ-N and η2-CN′ coordinated imine parts of the R-DAB ligand.  相似文献   

13.
[Pd(sac)(terpy)](sac)·4H2O (1), [Pt(sac)(terpy)](sac)·5H2O (2), [PdCl(terpy)](sac)·2H2O (3) and [PtCl(terpy)](sac)·2H2O (4) (sac = saccharinate, and terpy = 2,2′:6′,2″-terpyridine) have been synthesized and characterized by elemental analysis, FT-IR, 1H NMR and 13C NMR. In 1 and 2, a tridentate terpy ligand together with an N-coordinated sac ligand form the square-planar geometry around the palladium(II) or platinum(II) ions, while one sac anion remains outside the coordination sphere as a counter-ion. X-ray single crystal studies show that the [M(sac)(terpy)]+ ions in 1 and 2 reside in the centers of a hydrogen bonded honeycomb network formed by the uncoordinated sac ions and the lattice water molecules. Complexes 3 and 4 are isostructural and consist of a [M(Cl)(terpy)]+ cation, a sac anion and two lattice water molecules. The [M(Cl)(terpy)]+ ions interact with each other via M-M and π-π stacking interactions and these π interacted units are assembled to a 2D network by water bridges involving the sac ions and lattice water molecules. Convenient synthetic paths for 1-4 are also presented, and spectral, luminescence and thermal properties were discussed.  相似文献   

14.
Cation effects are studied on the excitation energy transfer reaction between anionic complexes, i.e., [Tb(dpa)3]3− (dpa=2,6-pyridinedicarboxylate) quenched by [Cr(ox)3]3− (ox=oxalate ion), [Cr(mal)3]3− (mal=malonate ion) and [Nd(dpa)3]3− in aqueous solutions in the presence of alkali metal ions added for adjustments of ionic strengths. In the quenching reaction of [Cr(ox)3]3−, magnitudes of quenching rate constants (kq) and energy transfer rate constant in encounter complex (k1) are changed by the cations in the order of Li+ < Na+ < K+ ≈ Rb+ ≈ Cs+, that is quite contrary of the cation effect on energy transfer reaction between [Ru(N-N)3]2+ and [Cr(ox)3]3−, reported in the previous paper. On the other hand, the rate constants in quenching reactions by [Cr(mal)3]3− and [Nd(dpa)3]3− remain almost constant. This result indicates that more separated donor-acceptor pair is not sensitive to coexisting cations.  相似文献   

15.
This study describes the quenching effects of p‐aminobenzenesulfonic acid (p‐ABSA) based on electrochemiluminescence (ECL) of the tris (2,2‐bipyridyl)‐ruthenium(II)(Ru(bpy)32+)/tri‐n‐propylamine (TPrA) system in aqueous solution. Quenching behaviours were observed with a 200‐fold excess of p‐ABSA over Ru(bpy)32+. In the presence of 0.1 M TPrA, the Stern‐Volmer constant (KSV) of ECL quenching was as high as 1.39 × 104 M‐1 for p‐ABSA. The logarithmic plot of inhibited ECL versus concentration of p‐ABSA was linear over the range of 6.0 × 10‐6 ‐3.0 × 10‐4 mol/L. The corresponding limit of detection was 1.2 × 10‐6 mol/L for p‐ABSA (S/N = 3). The mechanism of quenching is believed to involve an energy transfer from the excited‐state luminophore to a dimer of p‐ABSA and the adsorption of free radicals of p‐ABSA at the electrode surface that impeded the oxidation of the Ru(bpy)32+/TPrA system. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

16.
The reaction of [Ru(CO)2Cl2]n with bis(2-pyridylmethyl)amine (bpma) in refluxing ethanol followed by anion exchange yields two products: cis,fac-[Ru(bpma)(CO)2Cl]PF6 (1a, 71%) and trans,fac-[Ru(bpma)(CO)2Cl]PF6 (1b, 29%). Reaction of 1a with AgBF4 in acetone, followed by acetonitrile and then anion exchange gave cis,fac-[Ru(bpma)(CO)2(CH3CN)](PF6)2 (2a). In the same way, 1b afforded trans,fac-[Ru(bpma)(CO)2(CH3CN)](PF6)2 (2b). Reaction of depolymerized [Ru(CO)2Cl2]n with bpma in ethanol at room temperature afforded cis,cis-[Ru(η2-bpma)(CO)2Cl2] (3). In refluxing ethanol, 3 was converted to cis,fac-[Ru(bpma)(CO)2Cl]Cl (1a-Cl). Heating 3 in chlorobenzene afforded 1b-Cl, exclusively; heating 3 in ethylene glycol gave mainly 1a-Cl. Heating 1a-Cl in ethanol resulted in no isomerization, but heating in chlorobenzene gave a mixture of 3 and 1b-Cl. Anion exchange for PF6 with 1a-Cl and 1b-Cl afforded 1a and 1b, respectively, whereas anion exchange for BPh4 afforded 1a-BPh4. Compounds 1a, 1b, 2a and 3 have been structurally characterized.  相似文献   

17.
Specific salt effects were studied on the quenching reaction of excited [Ru(N-N)3]2+ (N-N=2,2-bipyridine (bpy), 1,10-phenanthrorine (phen)) and [Cr(bpy)3]3+ by [Cr(ox)3]3− (ox=oxalate ion) and [Cr(mal)3]3− (mal=malonate ion) in aqueous solutions as a function of alkali metal ions which were added for adjustment of ionic strength. The value of kq, quenching rate constants, and k1, energy transfer rate constant in encounter complex, is changed by the cations as the order of Li+ > Na+ > K+ ≈ Rb+ ? Cs+, although diffusion rate constants are not changed by the co-existing cations. Among the quenching reactions investigated in this work, a ratio of k1 values in the aqueous solutions whose ionic strength was adjusted with LiCl and KCl, k1LiCl/k1KCl, is larger for quenching systems of closely approached donor-acceptor pair than loosely bounded pair. These results indicate that co-existing alkali cation tunes the distance between donor and acceptor in encounter complex where energy transfer occurs.  相似文献   

18.
The previously reported complex [Ru(ttpy)(CN)3] [ttpy = 4′(p-tolyl)-2,2′:6′,2″-terpyridine] is conveniently synthesised by reaction of ttpy with Ru(dmso)4Cl2 to give [Ru(ttpy)(dmso)Cl2], which reacts in turn with KCN in aqueous ethanol to afford [Ru(ttpy)(CN)3] which was isolated and crystallographically characterised as both its (PPN)+ and K+ salts. The K+ salt contains clusters containing three complex anions and three K+ cations connected by end-on and side-on cyanide ligation to the K+ ions. The solution speciation behaviour of [Ru(ttpy)(CN)3] was investigated with both Zn2+ and K+ salts in MeCN, a solvent sufficiently non-competitive to allow the added metal cations to associate with the complex anion via the externally-directed cyanide lone pairs. UV-Vis spectroscopic titration of (PPN)[Ru(ttpy)(CN)3] with Zn(ClO4)2 showed a blue shift of 2900 cm−1 in the 1MLCT absorption manifold due to the ‘metallochromism’ effect; a series of distinct binding events could be discerned corresponding to formation of 4:1, 1:1 and then 1:3 anion:cation adducts, all with high formation constants, as the titration proceeded. In contrast titration of (PPN)[Ru(ttpy)(CN)3] with the more weakly Lewis-acidic KPF6 resulted in a much smaller blue-shift of the 1MLCT absorptions, and the titration data corresponded to formation of 1:1 and then 2:1 cation:anion adducts with weaker stepwise association constants of the order of 104 and then 103 M−1. Although association of [Ru(ttpy)(CN)3] resulted in a blue-shift of the 1MLCT absorptions, the luminescence was steadily quenched, as raising the 3MLCT level makes radiationless decay via a low-lying 3MC state possible.  相似文献   

19.
Three new silver(I) complexes of 5,5-diethlybarbiturate (barb), [Ag(barb)(apy)]·H2O (1), {[Ag(μ-ampy)][Ag(μ-barb)2]}n (2) and [Ag(barb)(dmamhpy)] (3) [apy = 2-aminopyridine, ampy = 2-aminomethylpyridine and dmamhpy = 2-(dimethylaminomethyl)-3-hydroxypyridine] have been synthesized and characterized by elemental analysis and FT-IR. Single crystal X-ray diffraction analyses showed that complexes 1 and 3 are mononuclear. In 1, the silver(I) ion is linearly coordinated by a barb anion and a ampy ligand, while a bidentate dmamhpy ligand together with an N-coordinated barb anion forms a trigonal coordination geometry around silver(I) in 3. Complex 2 is a one-dimensional coordination polymer in which silver(I) ions are bridged by ampy ligands, leading to a cationic chain . The [Ag(barb)2] units contains two N-bonded barb ligands, bridging the silver centers in the cationic and anionic units via the carbonyl O atoms. Thus, complex 2 contains two-coordinated and four-coordinated silver ions. All complexes display hydrogen-bonded network structures and exhibit appreciable fluorescence at room temperature. Thermal analysis (TG-DTA) data are in agreement with the structures of the complexes.  相似文献   

20.
The interactions of π-arene-Ru(II)-chloroquine complexes with human serum albumin (HSA), apotransferrin and holotransferrin have been studied by circular dichroism (CD) and UV-Visible spectroscopies, together with isothermal titration calorimetry (ITC). The data for [Ru(η6-p-cymene)(CQ)(H2O)Cl]PF6 (1), [Ru(η6-benzene)(CQ)(H2O)Cl]PF6 (2), [Ru(η6-p-cymene)(CQ)(H2O)2][PF6]2 (3), [Ru(η6-p-cymene)(CQ)(en)][PF6]2 (4), [Ru(η6-p-cymene)(η6-CQDP)][BF4]2 (5) (CQ: chloroquine; DP: diphosphate; en: ethylenediamine), in comparison with CQDP and [Ru(η6-p-cymene)(en)Cl][PF6] (6) as controls demonstrate that 1, 2, 3, and 5, which contain exchangeable ligands, bind to HSA and to apotransferrin in a covalent manner. The interaction did not affect the α-helical content in apotransferrin but resulted in a loss of this type of structure in HSA. The binding was reversed in both cases by a decrease in pH and in the case of the Ru-HSA adducts, also by addition of chelating agents. A weaker interaction between complexes 4 and 6 and HSA was measured by ITC but was not detectable spectroscopically. No interactions were observed for complexes 4 and 6 with apotransferrin or for CQDP with either protein. The combined results suggest that the arene-Ru(II)-chloroquine complexes, known to be active against resistant malaria and several lines of cancer cells, also display a good transport behavior that makes them good candidates for drug development.  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