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.
以半三明治钌(Ⅱ)(Ru1~Ru3)与1,3-双咪唑基苯配体(1~3)为前体, 合成了一系列钌(Ⅱ)矩形大环组装体 (4~12); 采用核磁共振波谱、 高分辨质谱、 元素分析和单晶X射线衍射表征了其组成和结构. 噻唑蓝(MTT)实验结果表明, 草酸型和苯醌型钌(Ⅱ)组装体4~9对测试的4种肿瘤细胞株均无毒性(IC50>100 μmol/L), 而萘醌型组装体10~12表现出较好的抗癌活性(IC50<2.22 μmol/L). 斑马鱼卵的急性毒性实验表明, 带有长烷氧基链的组装体12的毒性较低; 斑马鱼体内肿瘤实验表明, 组装体12在生物体内对A549(人肺癌细胞)有较好的抗癌作用. 研究表明, 组装体与DNA作用的主要方式是插入作用.  相似文献   

2.
The reactions of ruthenium(III), rhodium(III) and iridium(III) chlorides in molten lithium nitrite—sodium nitrite, lithium nitrite—potassium nitrit and sodium nitrite—potassium nitrite eutectics were studied and compared with those of their first row congeners. Ruthenium(III) reacted to form hexanitroruthenate(II) with the evolution of nitrogen dioxide, whereas rhodium(III) and iridium(III) formed hexanitrorhodate(III) and hexanitroiridate(III), respectively. These complexes decomposed at higher temperatures to form ruthenium(IV), rhodium(III) and iridium(IV) oxides, respectively, with the evolution of nitrogen oxides. The stoichiometries of these reactions were established by thermogravimetry and the products were characterized by their IR, visible and UV spectra, and X-ray diffraction patterns.  相似文献   

3.
建立了激光剥蚀-电感耦合等离子体质谱(LA-ICP-MS)法测定纯钌中Mg、Al、Fe、Ni、Cu、Zn、Rb、Rh、Pd、Mo、Ag、Cd、Sn、Ba、Ir、Pt、Au、Pb和Si等19种杂质元素的分析方法.优化了仪器参数,给出了激光能量为60%,剥蚀孔径为110μm,扫描速率为50μm/s,脉冲频率为10 Hz,载...  相似文献   

4.
Spirocyclic oxindoles undergo an enantioselective oxygenation reaction (nine examples; e.r. up to 97:3) upon catalysis by a chiral ruthenium porphyrin complex (1 mol %). The catalyst exhibits a lactam ring, which is responsible for substrate association through hydrogen bonds, and an active ruthenium center, which is in a defined spatial relationship to the oxygenation substrate. DFT calculations illustrate the perfect alignment of the active site with the reactive C? H bond and suggest—in line with the kinetic isotope effect—an oxygen rebound mechanism for the reaction.  相似文献   

5.
C?H activations with challenging arylacetamides were accomplished by versatile ruthenium(II) biscarboxylate catalysis. The distal C?H functionalization offers ample scope—including twofold oxidative C?H functionalizations and alkyne hydroarylations—through facile base‐assisted internal electrophilic‐type substitution (BIES) C?H ruthenation by weak O‐coordination.  相似文献   

6.
The structure of dilanthanum ruthenium pentoxide was solved by powder neutron diffraction at room temperature and 1.5 K. High‐temperature La2RuO5 crystallizes in the monoclinic space group P21/c. Upon cooling, the sample undergoes a phase transition to the triclinic low‐temperature form (space group P). This transition leads to pronounced changes in the Ru—O—Ru bond distances, resulting in a dimerization of the ruthenium ions.  相似文献   

7.
In the two ruthenium(II)–porphyrin–carbene complexes ­(di­benzoyl­carbenyl‐κC)(pyridine‐κN)(5,10,15,20‐tetra‐p‐tolyl­porphyrinato‐κ4N)­ruthenium(II), [Ru(C15H10O2)(C5H5N)(C48H36N4)], (I), and (pyridine‐κN)(5,10,15,20‐tetra‐p‐tolyl­porphyrinato‐κ4N)[bis(3‐tri­fluoro­methyl­phenyl)­carbenyl‐κC]­ruthenium(II), [Ru(C15H8F6)(C5H5N)(C48H36N4)], (II), the pyridine ligand coordinates to the octahedral Ru atom trans with respect to the carbene ligand. The C(carbene)—Ru—N(pyridine) bonds in (I) coincide with a crystallographic twofold axis. The Ru—C bond lengths of 1.877 (8) and 1.868 (3) Å in (I) and (II), respectively, are slightly longer than those of other ruthenium(II)–porphyrin–carbene complexes, owing to the trans influence of the pyridine ligands.  相似文献   

8.
The reaction of 2,2′:4,4′′:4′,4′′′‐quaterpyridyl (qtpy), with d6 ruthenium(II) (RuII), and rhenium(I) (ReI) metal centers has been investigated. The pendant pyridyl groups on the products have also been methylated to produce a second series of complexes containing coordinated Meqtpy2+. The absorption spectra of the complexes are dominated by intraligand and charge‐transfer bands. The ruthenium(II) complexes display broad unstructured luminescence consistent with emission from a Ru(d)→diimine(π*) manifold in acetonitrile solutions. In aqueous solutions, their emissions are weaker and the lifetimes are shorter. This effect is particularly acute for complexes incorporating coordinated dipyridylpyrazine, dppz, ligands. Although the emission of the ruthenium(II) complexes containing Meqtpy2+ is generally shorter than their qtpy analogs, it is notable that solvent‐dependent effects are much less intense. The rhenium(I) complexes also display broad unstructured luminescence but, compared with the ruthenium(II) systems, they have a relatively short lifetime in acetonitrile. Electrochemical studies reveal that all of the RuII complexes display chemically reversible metal‐based oxidations. ReI complexes only display irreversible metal‐based oxidations. In most cases, the reduction processes were not fully chemically reversible. The electrochemical and optical studies reveal that the nature of the lowest excited state of these complexes—particularly, the systems incorporating dppz—is highly dependent on the nature of the coordinated ligands. Calculations indicate that, although the excited state of most of the complexes is centered on the qtpy or Meqtpy2+ ligands, the excited state of the complexes containing dppz ligands is switched away from the dppz by qtpy methylation. A crystallographic study on one of the dicationic ruthenium(II) structures reveals that it forms an inclusion complex with benzene.  相似文献   

9.
With the right auxiliary phosphane ligands —for example, pentafluorophenyldiphenylphosphane or the sodium salt of 3,3′,3″-phosphinidynetris(benzenesulfonic acid)—the ruthenium(II )-catalyzed hydration of terminal alkynes to aldehydes proceeds by the previously unknown anti-Markovnikov addition of water [Eq. (a)]. Complexes of the type [RuCl2(PR2R2R″)x] (R=alkyl, Ph) are discussed as catalytically active species.  相似文献   

10.
The title compound, cis‐di‐μ‐perfluoroheptanoato‐κ4O:O′‐bis[dicarbonyl(dimethyl sulfoxide‐κS)ruthenium(I)](RuRu), [Ru2(C7F13O2)2(C2H6OS)2(CO)4], is a sawhorse‐type dinuclear ruthenium complex with two bridging perfluoroheptanoate ligands, and with two dimethyl sulfoxide (DMSO) ligands in the axial positions coordinating via the S atoms. It is a new example of a compound with an aliphatic fluorinated carboxylate ligand. The Ru—Ru bond distance of 2.6908 (3) Å indicates a direct Ru—Ru interaction. The compound is an active catalyst in transvinylation of propionic acid with vinyl acetate.  相似文献   

11.
Hexachlororuthenat (III) and the ruthenylion (RuO2+) react with dithizone on heating in weak acid water—ethanol solution. The complexes formed can be extracted into chloroform. Based on this reaction, a reproducible, selective determination of ruthenium (20 ppb) is achieved by means of substoichiometric isotope dilution analysis, using106Ru.  相似文献   

12.
Ruthenium complexes are well known to catalyze the oxidation of alcohols,1 where a ruthenium alkoxide has been assumed as a key intermediate. The formation of symmetrical ethers from certain carbinols is also catalyzed by ruthenium in a similar manner.2 We now wish to report that the mixed etherification of allyl alcohols is catalyzed by RuCl3, which acts as a Lewis acid to cleave the allylic carbon—oxygen bond.  相似文献   

13.
The T—silyl functionalized diamine‐bis(ether‐phosphine)ruthenium(II) complexes 1a(T ° ) — 1g(T ° ) (Scheme 1) were sol‐gel processed in the presence of different amounts of the co‐condensation agents CH3Si(OMe)3 (Me—T ° ) and (MeO)2SiMe—(CH2)6—MeSi(OMe)2 (D ° —C6—D ° ) to produce a library of the interphase catalysts X1a — X1c , X2a — X2g , and X3a — X3g . Due to the remarkable electronic and steric effects of the co‐ligands on the catalytic activity of such complexes, a series of aliphatic and aromatic diamines was selected. The new polymers were investigated by multinuclear CP/MAS solid‐state NMR spectroscopy as well as by EXAFS, EDX, SEM, and BET methods. Selected interphase catalysts show high activities and selectivities in the hydrogenation of trans‐4‐phenyl‐3‐butene‐2‐one.  相似文献   

14.
The title compound, mer‐[RuCl3N(C18H15As)2], is the first structurally characterized example of a nitride complex in which ruthenium is six‐coordinated to monodentate ligands only. The Ru[triple‐bond]N bond length [1.6161 (15) Å] is relatively long, and the trans influence of the nitride ligand is reflected by the difference between the Ru—Cltrans and Ru—Clcis bond lengths [0.1234 (4) Å]. The N—Ru—Cltrans axis is sited on a twofold axis.  相似文献   

15.
Dibarium μ‐oxido‐bis[pentachloridoruthenate(IV)] decahydrate, Ba2Ru2Cl10O·10H2O, has been prepared from ruthenium(III) chloride and barium chloride in hydrochloric acid. It crystallizes in the monoclinic system (space group C2/c). The structure consists of alternating layers of [Ru2Cl10O]4− and [Ba(H2O)7]2+ complex ions along the a direction. The O atom bonded to ruthenium occupies the 4e site, with symmetry, while the other atoms occupy general 8f sites. The overall structure is held together by O—H...O hydrogen bonds and O—H...Cl dipole–dipole interactions.  相似文献   

16.
The understanding of structure–function relationships within synthetic biomimetic systems is a fundamental challenge in chemistry. Herein we report the direct correlation between the structure of short peptoid ligands—N-substituted glycine oligomers incorporating 2,2′-bipyridine groups—varied in their monomer sequence, and the photoluminescence of RuII centers coordinated by these ligands. Based on circular dichroism and fluorescence spectroscopy we demonstrate that while helical peptoids do not affect the fluorescence of the embedded RuII chromophore, unstructured peptoids lead to its significant decay. Transmittance electron microscopy (TEM) revealed significant differences in the arrangements of metal-bound helical versus unstructured peptoids, suggesting that only the latter can have through-space interactions with the ruthenium dye leading to its quenching. High-resolution TEM enabled the remarkable direct imaging of singular ruthenium-bound peptoids and bundles, supporting our explanation for structure-depended quenching. Moreover, this correlation allowed us to fine-tune the luminescence properties of the complexes simply by modifying the sequence of their peptoid ligands. Finally, we also describe the chiral properties of these Ru–peptoids and demonstrate that remote chiral induction from the peptoids backbone to the ruthenium center is only possible when the peptoids are both chiral and helical.  相似文献   

17.
Mixtures of iridium(IV) and ruthenium(III) as IrCl2?6 and RuClt—6 are separated by ion flotation. Iridium (IV) is selectively floated from mixtures of the metal complexes in aqueous 1.0 M hydrochloric acid with hexadecylpyridinium bromide (HPB) and nitrogen. Ruthenium(III) does not float under the same conditions. In order to assess the usefulness of this procedure, the separation was also investigated with hexadecyltrimethylammonium bromide and hexadecyltripropylammonium bromide, from solutions of varying concentrations of sodium chloride, sodium nitrite and hydrochloric acid. Under optimal conditions at the 5 × 10?5 M level, 78% of the iridium is recovered free of ruthenium, provided that excess of HPB and > 1 M chloride are present.  相似文献   

18.
A new ruthenium(II) complex [Ru(NO2)4(CO)(H2O)]Cl2 · 2H2O has been prepared and characterized structurally. The compound crystallizes in monoclinic space group C2 /m, with the unit cell parameters a = 12.913(2), b = 14.605(2), c = 7.4494(1)Å, ß =121.49(2)°; V = 1198.0(3)Å3, Z = 4, Dc = 2.429 Mgm—3; μ = 1.83 mm—1; R = 0.0455, wR = 0.1552. The complex contains four neutral NO2 ligands. The ruthenium atom is six‐coordinated to four nitrogen atoms of nitrogen dioxide, one carbon atom from carbon monoxide and one oxygen atom from water molecule, forming slightly distorted octahedral coordination. The preparation procedure has been discussed.  相似文献   

19.
Kinetics of uncatalyzed and ruthenium(III) catalyzed oxidation of monoethanolamine by N-bromosuccinimide (NBS) has been studied in an aqueous acetic acid medium in the presence of sodium acetate and perchloric acid, respectively. In the uncatalyzed oxidation the kinetic orders are: the first order in NBS, a fractional order in the substrate. The rate of the reaction increased with an increase in the sodium acetate concentration and decreased with an increase in the perchloric acid concentration. This indicates that free amine molecules are the reactive species. Addition of halide ions results in a decrease in the kinetic rate, which is noteworthy. Both in absence and presence of a catalyst, a decrease in the dielectric constant of the medium decreases the kinetic rate pointing out that these are dipole—dipole reactions. A relatively higher oxidation state of ruthenium i.e., Ru(V) was found to be the active species in Ru(III) catalyzed reactions. A suitable mechanism consistent with the observations has been proposed and a rate law has been derived to explain the kinetic orders.  相似文献   

20.
Lithium hydride (LiH) has a strong effect on iron leading to an approximately 3 orders of magnitude increase in catalytic ammonia synthesis. The existence of lithium–iron ternary hydride species at the surface/interface of the catalyst were identified and characterized for the first time by gas‐phase optical spectroscopy coupled with mass spectrometry and quantum chemical calculations. The ternary hydride species may serve as centers that readily activate and hydrogenate dinitrogen, forming Fe‐(NH2)‐Li and LiNH2 moieties—possibly through a redox reaction of dinitrogen and hydridic hydrogen (LiH) that is mediated by iron—showing distinct differences from ammonia formation mediated by conventional iron or ruthenium‐based catalysts. Hydrogen‐associated activation and conversion of dinitrogen are discussed.  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