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.
ABSTRACT

The ruthenium species [Ru(NH3)6]3+ was adsorbed by the surface of (α-Sn(HPO4)2.2H2O [SnP]) α-tin(IV) hydrogen phosphate. The ruthenium(II)-containing cation [Ru(NH3)6]2+, however, has been directly intercalated into (SnP). Since the corresponding ruthenium(III) complex cation was not so intercalated, a new self-catalysed intercalation mechanism involving labile ammonia ligands from the ruthenium(II) has been proposed. At high loadings, guest ruthenium(II) species were oxidised to ruthenium(III).  相似文献   

2.
A hybrid material (AB-1) was obtained by immobilizing photolabile ruthenium nitrosyl complex [Ru(LSBH)(PPh3)2(NO)Cl](ClO4) (1) into alginate polymer. A controlled release of nitric oxide was observed when a suspension of AB-1 was exposed to visible light. The amount of photoreleased NO from polymeric hybrid material was estimated using Griess reagent assay.  相似文献   

3.
Homoleptic Ru(II)–diphosphine and Ru(II)–diarsine complexes [Ru(L-L)3](OTf)2 (L-L=Me2P(CH2)nPMe2; n=1, dmpm; n=2, dmpe) and 1,2-C6H4(AsMe2)2 (diars) have been isolated, in which the Ru(II) state is stabilised to an unprecedented degree, and the crystal structure of [Ru(diars)3]Cl2 has been determined.  相似文献   

4.
Electroactive 2,2′: 6,2″-terpyridinyl ligands ( 3, 5, 6 ) and their iron(II) ( 7a–9a ) and ruthenium(II) complexes ( 7b–9b ) were synthesized. Bis[3-(aminophenyl)-2,2′ :6,2″-terpyridinyl]metal(II) complexes ( 7a, 7b ) and bis[2-(hydroxyphenyl)-2,2′ :6,2″-terpyridinyl]metal(II) complexes ( 8a, 8b ) were electropolymerized on to the surface of Pt or In-SnO2 (ITO) electrodes in acetonitrile containing Bu4NCIO4 by scanning the potential between O and + 1.6V (for 7a and 7b ), and ?0.8 and +1.6V (for 8a and 8b ) versus saturated calomel electrode. The electrodes obtained by electropolymerization exhibited reversible electrochromism based on Fe(II)/Fe(III) or Ru(II)/Ru(III) redox couple. Photoresponses to visible light were found in the modified electrode obtained by electropolymerization of ruthenium complex 7b in an aqueous LiClO4 solution containing methylviologen (cation MV2+) under an O2 atmosphere. The mechanism for the photoresponded cathodic current was explained in terms of an excitation of bis(terpyridinyl)ruthenium(II) complex [Ru(terpy)22+] by visible light, an electron transfer from the excited state [Ru(terpy)2+*2] to MV2+, reduction of Ru(terpy)3+2 at an electrode, and oxidation of MV+* with O2.  相似文献   

5.
The modification of a gold electrode surface by electropolymerization of trans-[Ru(NH3)4(Ist)SO4]+ to produce an electrochemical sensor for nitric oxide was investigated. The influence of dopamine, serotonin and nitrite as interferents for NO detection was also examined using square-wave voltammetry (SWV). The characterization of the modified electrode was carried out by cyclic voltammetry, electrochemical quartz crystal microbalance (EQCM) and SERS techniques. The gold electrode was successfully modified by the trans-[Ru(NH3)4(Ist)SO4]+ complex ion using cyclic voltammetry. The experiments show that a monolayer of the film is achieved after ten voltammetric cycles, that NO in solution can coordinate to the metal present in the layer, that dopamine, serotonin and nitrite are interferents for the detection of NO, and that the response for the nitrite is much less significant than the responses for dopamine and serotonin. The proposed modified electrode has the potential to be applied as a sensor for NO.  相似文献   

6.
Lignin gasification in supercritical water over charcoal supported ruthenium trivalent salts was studied using a batch reactor at 673 K. Ruthenium (III) nitrosyl nitrate on charcoal (Ru(NO)(NO3)3/C) was more active than ruthenium (III) chloride on charcoal (RuCl3/C) for the gasification reaction. EXAFS analysis revealed that ruthenium metal particles were formed in both RuCl3/C and Ru(NO)(NO3)3/C catalysts during the lignin gasification and that the size of ruthenium metal in Ru(NO)(NO3)3/C was smaller than that in RuCl3/C. It was concluded that well-dispersed ruthenium metal particles were active for the lignin gasification in supercritical water.  相似文献   

7.
The allylation of 1,3‐dicarbonyl compounds and malononitrile with aliphatic allylic substrates is achieved under mild conditions in the presence of new ruthenium catalysts. The ruthenium complex [Ru(C5Me5)(2‐quinolinecarboxylato)(CH2CHCH‐n‐Pr)] [BF4] as a precatalyst, allows the synthesis of mono‐allylated branched derivatives. On the other hand, the parent complex [Ru(C5Me5)(MeCN)3] [PF6] as a precatalyst, straightforwardly favours the bis‐allylation of the procarbonucleophiles leading to bis‐allylated bis‐linear products. The involvement of the two precatalysts provides a sequential synthesis of unsymmetrical mixed linear‐branched bis‐allylated derivatives.  相似文献   

8.
The diruthenium(2.5) complex [(Me3TACN)Ru(μ-Cl)3Ru(Me3TACN)]-(PF6)2, Me3TACN = 1,4,7-trimethyl-1,4,7-triazacyclononane, has been crystallized for structural characterization. The results are reproduced by density functional theory (DFT) calculations and confirm the sensitivity of the central Ru(μ-Cl)3Ru core to contacts between the Cl bridging atoms and the co-ligands. The singly occupied MO is characterized as a σ* MO involving the metal dz2 orbitals and a small halide contribution by DFT calculations and EPR.  相似文献   

9.
Tris(N-phenyldithiocarbamato) ruthenium(III) complexes, [Ru(L1)3] (1); tris(N-(4-methylphenyl)dithiocarbamato)) ruthenium(III), [Ru(L2)3] (2); and tris(N-(4-methoxyphenyl)dithiocarbamato)) ruthenium(III), [Ru(L3)3] (3) were synthesized and characterized by elemental analysis, thermogravimetric analysis, FTIR, UV–VIS and NMR spectroscopy. TGA analyses show major degradation of all complexes in the range 120–350°C, leading to the formation of residual weight corresponding to ruthenium (III) sulfides. The 1H-NMR spectra of the ligands and complexes are in agreement with the proposed structures. FTIR studies confirmed that the ligands coordinate the Ru3+ ion in a bidentate chelating mode. The complexes were thermolysed at 180°C to prepare hexadecylamine-capped Ru2S3 nanoparticles. Powder X-ray diffraction patterns revealed the formation of hexagonal-phase Ru2S3 nanoparticles with average crystallite sizes ranging from 8.3 to 9.5?nm. TEM images showed the crystalline clusters with shapes ranging from square to hexagonal, while SEM images elucidated that the particles were agglomerated. Energy-dispersive X-ray spectra confirmed the presents of Ru2S3 nanoparticles.  相似文献   

10.
A new type of binuclear Ru(II) complex [Ru(TAP)2(TAP-TAP)Ru(TAP)2]4+ (1) (TAP-TAP=2,2′-bis(1,4,5,8-tetraazaphenanthrene)) produced by photo-reaction of Ru(TAP)32+ in the presence of 5′-guanosine-monophosphate (5′-GMP) has been isolated and characterised. The formation of 1 is proposed to proceed via mono-reduced Ru(TAP)32+, i.e. [Ru2+(TAP)2(TAP⋅−)]+. The emission lifetime of 1 is much greater than that of Ru(TAP)32+, consistent with an increased energy gap between the 3MLCT and 3MC states in 1. This arises because of stabilisation of its 3MLCT state due to the localisation of the excited state electron on the TAP-TAP ligand.  相似文献   

11.
The syntheses and properties of trans-[Ru(NH3)4(L)(NO)](BF4)3 (L = isonicotinic acid (inaH) (I) or ina-Tat48–60 (II)) are described. Tat48–60, a cell penetrating peptide fragment of the Tat regulatory protein of the HIV virus, was linked to the ruthenium nitrosyl through inaH. I and II release NO after reduction forming trans-[Ru(NH3)4(L)(H2O)]3 +. The IC50 values against B16-F10 melanoma cells of I and II (21 μmol L 1 and 23 μmol L 1, respectively) are close to that of the commercially available cisplatin (33 μmol L 1) and smaller than similar complexes. The cytotoxicity is assigned to the NO released from I and II.  相似文献   

12.
A new and promising nitrosyl ruthenium complex, [Ru(NO)(bdqi-COOH)(terpy)](PF6)3, bdqi-COOH is 3,4-diiminebenzoic acid and terpy is 2,2′-terpyridine, has been synthesized as a NO donor agent. The procedure used for [Ru(NO)(bdqi-COOH)(terpy)](PF6)3 synthesis has, apparently, yielded the formation of two isomers in which the ligand bdqi-COOH appears to be coordinated in its reduced form (bdcat-COOH), which could have differences in their pharmacological properties. Therefore, it was intended to separate the two possible isomers by high-performance liquid chromatography (HPLC) and to characterize them by high resolution mass spectrometry (QTOF MS) and by magnetic nuclear resonance spectroscopy (NMR). The results obtained by MS showed that the ESI-MS mass spectra of both HPLC column fractions, e.g. peak 1 and peak 2, are essentially equal, showing that both isomers display nearly identical gas-phase behavior with clusters of isotopologue ions centered at m/z 573, m/z 543 and m/z 513. Regarding the NMR analysis, the results showed that the positional isomerism is located in the bdqi-COOH ligand. From the observed results it can be concluded that the synthesis procedure that has been used results in the formation of two [Ru(terpy)(bdqi-COOH)NO](PF6)3 isomers.  相似文献   

13.
New ion-conducting polymer composite films have been prepared, and their ionic conducting properties have been investigated. The polymer composite films are fabricated from partially phosphorylated poly(vinyl alcohol) with tetramethylammonium salt (P-PVA·Me4N+) and poly(acrylic acid) (PAA) or poly(ethylene glycol) (PEG). For P-PVA·Me4N+/PEG composite films, the ionic conductivity and carrier density sharply increased, and carrier mobility sharply decreased around [PEG]/[PO3]P-PVA of 2. The ionic conductivity is dominated by both carrier density and carrier mobility at [PEG]/[PO3]P-PVA<2 and only by carrier density at [PEG]/[PO3]P-PVA>2. This is attributed to the fact that the ionic conduction in P-PVA·Me4N+/PEG composite films occurred through the PEG-Me4N+ complex which was independent of the carrier mobility. On the other hand, the ionic conductivity in P-PVA·Me4N+/PAA composite films showed a bell-shaped dependence on the PAA contents with a maximum value at [CO2H]PAA/[OH]P-PVA=1. FTIR spectrum measurements demonstrated that part of the carboxylic acid residues was dissociated in the composite films. This fact implied that the ionic conduction was mediated by PAA at the low PAA content. At high PAA content, however, an excess of the carboxylic acid residues formed trapping sites for the Me4N+ ion, leading to a decrease in the ionic conductivity. Furthermore, we proposed a unique mechanism of the ionic conduction.  相似文献   

14.
Two new aldehyde-decorated tpy and bpy-containing ruthenium(II) complexes, [Ru(1)(bpy)2][PF6]2 and [Ru(1)(tpy)Cl][PF6] in which 1 is 5,5′-bis(4-formylphenyl)-2,2′-bipyridine, have been prepared and fully characterized. The packing in both solid state structures involves extensive Oaldehyde···HCpyridine contacts, but π-stacking interactions are important only between [Ru(1)(tpy)Cl]+ cations.  相似文献   

15.
Organometallic ruthenium(II) complexes of general formula [(η6‐arene)Ru(curcuminato)Cl], with arene being piPrC6H4Me ( 1 ), C6H6 ( 2 ), and C6Me6 ( 3 ), were synthesized, characterized, and evaluated for their antitumor effects. Specifically, we explored their ability to regulate the proteasome, a validated pharmacological target in cancer treatment. Ruthenium complexes inhibited isolated proteasomes to various extents, with the biological activity of these complexes depending on the nature of the bound arene; in particular, [(η6‐arene)Ru(curcuminato)Cl] 2 suppressed proteasomal activities more potently than 1 , 3 , or free curcumin. Each complex also inhibited proteasomes in cultured colon cancer cells and consequently triggered apoptosis, with the [(η6‐benzene)Ru(curcuminato)Cl] complex 2 being the most active. The influence on the oxidative status of HCT116 cells and the DNA binding ability of the [(η6‐arene)Ru(curcuminato)Cl] complexes were studied. Complex 2 showed the highest antioxidant capacity; moreover, complexes 1 and 2 were shown to bind isolated DNA with higher affinity (up to threefold) than free curcumin. Collectively, our results demonstrate that the complexation of curcumin with ruthenium(II) is a promising starting point for the development of curcumin‐based anticancer drugs.  相似文献   

16.
By reacting a stoichiometric amount of BPM or Me2BPM (BPM = bis(pyrazol-1-yl)methane, Me2BPM = bis(3,5-dimethylpyrazol-1-yl)methane) with Pd(OAc)2 (OAc = acetate) in CH2Cl2, the new neutral complexes [Pd(OAc)2(BPM)] (1) and [Pd(OAc)2(Me2BPM)] (1Me) were isolated and characterized by NMR and IR spectroscopy. Acidolysis of one Pd-OAc bond of compound 1 by using aqueous perchloric acid in methanol led to the formation of the stable cationic aqua-complex [Pd(OAc)(H2O)(BPM)](ClO4) (2-ClO4). The [Pd(OAc)(H2O)(BPM)]+ cation was obtained also by reacting 1 with trifluoromethanesulfonic acid. NMR studies on inter- and intra-molecular hydrogen bonding of the aqua-ligand have been carried out and the data obtained have been interpreted on the basis of the results of DFT calculations.  相似文献   

17.
A novel dinuclear ruthenium complex, [(tpy)Ru(tmbimbpyH4)Ru(tpy)](ClO4)4, has been synthesized to introduce a new proton-induced molecular switching system, where tpy=2,2′:6′,2″-terpyridine and tmbimbpyH4=2,6,2′,6′-tetra(4,5-dimethylbenzimidazol-2-yl)-4,4′-bipyridine. Both fully protonated and deprotonated forms of the dinuclear Ru complex were isolated and the forms were interchangeable by the addition of acid or base. A longer wavelength shift and a negative oxidation potential shift were observed when the protonated Ru complex was fully deprotonated.  相似文献   

18.
Reaction of 2,6-bis(pyrrolidin-2-yl)pyridine (LH4) with RuCl3·3H2O in refluxing methanol/water mixtures gives rise to the formation of the octahedral complexes [Ru(LH4)(L)]2+, in which one of the two trihapto ligands has been dehydrogenated as 2,6-bis(3,4-dihydro-2H-pyrrol-5-yl)pyridine (L), even if LH4 was present in excess. With the three stereoisomers of LH4, the complexes [Ru(R,S-LH4)(L)]2+ (meso), [Ru(R,R-LH4)(L)]2+ and [Ru(S,S-LH4)(L)]2+ have been isolated as the perchlorate salts and characterised by X-ray structure analysis and by CD spectra.  相似文献   

19.
The phase relationships in binary, ternary, and more complex Me 2 +O–Me 2+O–Me 2 3+O3Me 4+O2–TiO2 systems (Me + = Li+, K+, Rb+, Cs+; Me 2+ = Mg2+, Sr2+, Ba2+, Zn2+; Me 3+ = Al3+, Fe3+, Ga3+; Me 4+ = Sn4+, Zr4+) are investigated in the concentration regions corresponding to the compositions of titanates with a tunnel structure: Li2(Me 2+,Me 3+) y (Me 4+,Ti)4O8 ramsdellites, (Me +,Me 2+) x (Me 2+,Me 3+) y (Me 4+,Ti)8O16 hollandites, and (Ba,Me 2+)2(Me 4+,Ti)9O20 phases. The homogeneity regions of the solid solutions with the above structures are determined, and their crystal chemical characteristics, phase transformations, and thermal and electrical properties are studied. The results obtained and the data available in the literature are analyzed and generalized. The general approaches to the prediction of changes in the structure and properties of the studied titanates with a variation in the chemical composition due to isomorphous substitutions in different structural positions of the crystal lattice are discussed.  相似文献   

20.
A novel dinuclear ruthenium(II) complex [(bpy)2Ru(bdptb)Ru(bpy)2]4+ (bpy=2,2-bipyridyl; bdptb=2,2-bis(5,6-diphenyl-1,2,4-triazin-3-yl)-4,4-bipyridine) has been synthesized and characterized. The DNA-binding behaviors of this complex have been studied by viscosity, absorption and circular dichroism (CD) spectra. It has firstly been found that the dinuclear ruthenium(II) complex display the enantiopreferential DNA-binding after equilibrium dialysis.  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