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.
Two new aminophosphines – furfuryl‐(N‐dicyclohexylphosphino)amine, [Cy2PNHCH2–C4H3O] ( 1 ) and thiophene‐(N‐dicyclohexylphosphino)amine, [Cy2PNHCH2–C4H3S] ( 2 ) – were prepared by the reaction of chlorodicyclohexylphosphine with furfurylamine and thiophene‐2‐methylamine. Reaction of the aminophosphines with [Ru(η6p‐cymene)(μ‐Cl)Cl]2 or [Ru(η6‐benzene)(μ‐Cl)Cl]2 gave corresponding complexes [Ru(Cy2PNHCH2–C4H3O)(η6p‐cymene)Cl2] ( 1a ), [Ru(Cy2PNHCH2–C4H3O)(η6‐benzene)Cl2] ( 1b ), [Ru(Cy2PNHCH2–C4H3S)(η6p‐cymene)Cl2] ( 2a ) and [Ru(Cy2PNHCH2–C4H3S)(η6‐benzene)Cl2] ( 2b ), respectively, which are suitable catalyst precursors for the transfer hydrogenation of ketones. In particular, [Ru(Cy2PNHCH2–C4H3S)(η6‐benzene)Cl2] acts as a good catalyst, giving the corresponding alcohols in 98–99% yield in 30 min at 82 °C (up to time of flight ≤ 588 h?1). Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

2.
BaCeO3‐a and BaCeO3‐b, with strong basic sites, were synthesized by using a co‐precipitation method at different calcination temperatures, and used as supports to evaluate their performance in ammonia synthesis. The ammonia synthesis rate with the 1.25 % Ru/BaCeO3‐a catalyst is 24 mmol g?1 h?1, which is higher than that of 1.25 % Ru/BaCeO3‐b catalyst (18 mmol g?1 h?1) at 3 MPa and 450 °C. Moreover, the performance of the 4 % Cs‐1.25 % Ru/BaCeO3‐a catalyst was further improved to 28 mmol g?1 h?1, and no sign of deactivation was observed after a reaction time of 120 h. The XPS and H2 temperature‐programmed reduction analyses indicated that the Ru/BaCeO3‐a catalyst has more oxygen vacancies than the Ru/BaCeO3‐b catalyst. In addition, the average Ru particle size of the Ru/BaCeO3‐a catalyst is closer to 2 nm than the Ru/BaCeO3‐b catalyst, which promotes the generation of B5‐type sites (the active site for N2 dissociation). The CO2 temperature‐programmed desorption analysis indicates that BaCeO3‐a has a high basic density, which is beneficial for electron transfer to Ru and further facilitates the dissociation of N≡N bonds.  相似文献   

3.
A novel Ru‐Fe‐B/ZrO2 catalyst for the selective hydrogenation of benzene to cyclohexene was prepared by the chemical reduction method. A yield of cyclohexene of 57.3% was achieved at benzene conversion of 80.6% on this catalyst. The activity and yield of cyclohexene were higher than those studied previously. The structural characterizations of the catalyst were performed by TEM‐SAED, XRD, and N2‐physisorption. Moreover, cyclohexene selectivities on this catalyst increased and the activities decreased with the increase of the ZnO dosages, however, the activities increased and cyclohexene selectivities decreased with the increase of the H2SO4 dosages. Different feeding manners of H2SO4 or ZnO exerted definitely influence on the performances of this catalyst, but the degrees of influence were different due to the character of chemisorptions. Furthermore, the activity and cyclohexene selectivity on the catalysts could be reversibly modified by adding H2SO4 or ZnO into reaction slurry, which provides an easy method to recover the activity and selectivity of Ru‐Fe‐B/ZrO2 catalysts during the process of producing cyclohexene. And the modifiable mechanisms involved were speculated.  相似文献   

4.
用沉淀法制备了单金属纳米Ru(0)催化剂,考察了ZnSO4和La2O3作共修饰剂对该催化剂催化苯选择加氢制环己烯性能的影响,并用X射线衍射(XRD)、X射线荧光(XRF)光谱、X射线光电子能谱(XPS)、俄歇电子能谱(AES)、透射电镜(TEM)和N2物理吸附等手段对加氢前后催化剂进行了表征. 结果表明,在ZnSO4存在下,随着添加碱性La2O3量的增加,ZnSO4水解生成的(Zn(OH)23(ZnSO4)(H2O)x(x=1,3)盐量增加,催化剂活性单调降低,环己烯选择性单调升高. 当La2O3/Ru 物质的量比为0.075 时,Ru催化剂上苯转化率为77.6%,环己烯选择性和收率分别为75.2%和58.4%. 且该催化体系具有良好的重复使用性能. 传质计算结果表明,苯、环己烯和氢气的液-固扩散限制和孔内扩散限制都可忽略. 因此,高环己烯选择性和收率的获得不能简单归结为物理效应,而与催化剂的结构和催化体系密切相关. 根据实验结果,我们推测在化学吸附有(Zn(OH)23(ZnSO4)(H2O)x(x=1,3)盐的Ru(0)催化剂有两种活化苯的活性位:Ru0和Zn2+. 因为Zn2+将部分电子转移给了Ru,Zn2+活化苯的能力比Ru0弱. 同时由于Ru和Zn2+的原子半径接近,Zn2+可以覆盖一部分Ru0活性位,导致解离H2的Ru0活性位减少. 这导致了Zn2+上活化的苯只能加氢生成环己烯和Ru(0)催化剂活性的降低. 本文利用双活性位模型来解释Ru基催化剂上的苯加氢反应,并用Hückel分子轨道理论说明了该模型的合理性.  相似文献   

5.
In order to increase the catalyst activity for Fischer–Tropsch synthesis (FTS), the preparation methods of two new catalysts were studied. The chemically identical bimetallic Co–Mn/Al2O3 catalysts were synthesized by different synthetic methods: (a) via thermal decomposition of the complex [Co1.33Mn0.667(C7H3NO4)2(H2O)5].2H2O ( 1 ) and (b) by the impregnation technique. The complex was characterized by the single‐crystal analysis, elemental analysis, and Fourier‐transform infrared (FT‐IR) spectroscopy. Both catalysts were characterized by powder X‐ray diffraction (XRD), scanning electron microscopy (SEM), energy dispersive X‐ray spectrometry (EDS), Brunauer–Emmett–Teller (BET) specific surface area, hydrogen temperature‐programmed reduction (H2‐TPR), and H2‐chemisorption. The catalysts' activity was investigated for the Fischer–Tropsch synthesis in a fixed bed microreactor. Higher activity was obtained for the catalyst prepared by thermal decomposition of the inorganic precursor due to its small particle size, superior dispersion, and higher surface area. The results show that the catalyst prepared thermal decomposition has 21% ethylene, 10% propylene, and 50% C5+ selectivity, while methane selectivity of this catalyst is 11% at 250°C. On the other hand, the catalyst obtained by the impregnation method displays 15% ethylene, 8% propylene, 29% C5+, and 29% methane selectivity at the same temperature.  相似文献   

6.
Addition of the amine–boranes H3B ? NH2tBu, H3B ? NHMe2 and H3B ? NH3 to the cationic ruthenium fragment [Ru(xantphos)(PPh3)(OH2)H][BArF4] ( 2 ; xantphos=4,5‐bis(diphenylphosphino)‐9,9‐dimethylxanthene; BArF4=[B{3,5‐(CF3)2C6H3}4]?) affords the η1‐B? H bound amine–borane complexes [Ru(xantphos)(PPh3)(H3B ? NH2tBu)H][BArF4] ( 5 ), [Ru(xantphos)(PPh3)(H3B ? NHMe2)H][BArF4] ( 6 ) and [Ru(xantphos)(PPh3)(H3B ? NH3)H][BArF4] ( 7 ). The X‐ray crystal structures of 5 and 7 have been determined with [BArF4] and [BPh4] anions, respectively. Treatment of 2 with H3B ? PHPh2 resulted in quite different behaviour, with cleavage of the B? P interaction taking place to generate the structurally characterised bis‐secondary phosphine complex [Ru(xantphos)(PHPh2)2H][BPh4] ( 9 ). The xantphos complexes 2 , 5 and 9 proved to be poor precursors for the catalytic dehydrogenation of H3B ? NHMe2. While the dppf species (dppf=1,1′‐bis(diphenylphosphino)ferrocene) [Ru(dppf)(PPh3)HCl] ( 3 ) and [Ru(dppf)(η6‐C6H5PPh2)H][BArF4] ( 4 ) showed better, but still moderate activity, the agostic‐stabilised N‐heterocyclic carbene derivative [Ru(dppf)(ICy)HCl] ( 12 ; ICy=1,3‐dicyclohexylimidazol‐2‐ylidene) proved to be the most efficient catalyst with a turnover number of 76 h?1 at room temperature.  相似文献   

7.
Selective hydrogenation of aromatic amines,especially chemicals such as aniline and bis(4-aminocyclohexyl)methane for non-yellowing polyurethane,is of particular interests due to the extensive applications.To conquer the existing difficulties,in selective hydrogenation,,the Ru~0-Ru~(δ+)/CeO_2 catalyst with solid frustrated Lewis pairs was developed for aromatic amines hydrogenation with excellent activity and selectivity under relative milder conditions.The morphology,electronic and chemical properties,especially the Ru~O-Ru~(δ+) clusters and reducible ceria were characterized by X-ray diffraction(XRD),transmission electron microscopy(TEM),sca nning electronic microscopy(SEM),X-ray photoelectron sp ectroscopy(XPS),CO_2 tempe rature programmed deso rption(CO_2-TPD),H_2 tempe rature programmed reduction(H_2-TPR),H_2 diffuse reflectance Fourier transform infrared spectroscopy(H_2-DRIFT),Raman,etc.The 2% Ru/CeO_2 catalyst exhibited good conversion of 95% and selectivity greater than 99% toward cyclohexylamine.The volcano curve describing the activity and Ru state was found.Owning to the "acidic site isolation" by surrounding alkaline sites,condensation between the neighboring amine molecules could be effectively suppressed.The catalyst also showed good stability and applicability for other aromatic amines and heteroarenes containing different functional groups.  相似文献   

8.
Three Ru(bpy)32+ derivatives tethered to multiple viologen acceptors, [Ru(bpy)2(4,4′‐MV2)]6+, [Ru(bpy)2(4,4′‐MV4)]10+, and [Ru(bpy)(4,4′‐MV4)2]18+ [bpy=2,2′‐bipyridine, 4,4′‐MV2=4‐ethoxycarbonyl‐4′‐(N‐G1‐carbamoyl)‐2,2′‐bipyridine, and 4,4′‐MV4=4,4′‐bis(N‐G1‐carbamoyl)‐2,2′‐bipyridine, where G1=Asp(NHG2)‐NHG2 and G2=‐(CH2)2‐N+C5H4‐C5H4N+‐CH3] were prepared as “photo‐charge separators (PCSs)”. Photoirradiation of these complexes in the presence of a sacrificial electron donor (EDTA) results in storage of electrons per PCS values of 1.3, 2.7, and 4.6, respectively. Their applications in the photochemical H2 evolution from water in the presence of a colloidal Pt H2‐evolving catalyst were investigated, and are discussed along with those reported for [Ru(bpy)2(5,5′‐MV4)]10+, [Ru(4,4′‐MV4)3]26+, and [Ru(5,5′‐MV4)3]26+ (Inorg. Chem. Front. 2016 , 3, 671–680). The PCSs with high dimerization constants (Kd=105–106 m ?1) are superior in driving H2 evolution at pH 5.0, whereas those with lower Kd values (103–104 m ?1) are superior at pH 7.0, where Kd=[(MV+)2]/[MV+ . ]2. The (MV+)2 site can drive H2 evolution only at pH 5.0 as a result of its 0.15 eV lower driving force for H2 evolution relative to MV+ . , whereas the PCSs with lower Kd values exhibit higher performance at pH 7.0 owing to the higher population of free MV+ . . Importantly, the rate of electron charging over the PCSs is linear to the apparent H2 evolution rate, and shows an intriguing quadratic dependence on the number of MV2+ units per PCS.  相似文献   

9.
Monodisperse Ru–B amorphous alloy catalysts were synthesized by ultrasound-assisted chemical reduction of (NH4)2RuCl6 with BH 4 ? . With the characterization of X-ray diffraction (XRD), selective area electronic diffraction (SAED), X-ray photoelectron spectroscopy (XPS), differential scanning calorimetry (DSC), and transmission electron microscopy (TEM), the resulting Ru–B nanoparticles were identified to be amorphous alloys ranging in size from 2.4 to 4.9 nm. During liquid-phase maltose hydrogenation, the as-synthesized Ru–B catalyst was extremely active compared to the regular Ru–B obtained via the reduction of RuCl3 with BH 4 ? . The Ru–B sample prepared under ultrasonication with 60 W was proven to be the most active catalyst. Its catalytic activity was nearly 11 times that of industrial Raney Ni, and could be used repetitively for more than six times without significant deactivation.  相似文献   

10.
Hydrogen transfer reduction processes are attracting increasing interest from synthetic chemists in view of their operational simplicity. Reaction of [Ph2PNHCH2‐C4H3S] with [Ru(η6‐benzene)(µ‐Cl)Cl]2, [Rh(µ‐Cl)(cod)]2 and [Ir(η5‐C5Me5)(µ‐Cl)Cl]2 gave a range of new monodendate complexes [Ru(Ph2PNHCH2‐C4H3S)(η6‐benzene)Cl2], 1, [Rh(Ph2PNHCH2‐C4H3S)(cod)Cl], 2, and [Ir(Ph2PNHCH2‐C4H3S)(η5‐C5Me5)Cl2], 3, respectively. All new complexes were fully characterized by analytical and spectroscopic methods. 1H? 31P NMR, 1H? 13C HETCOR or 1H? 1H COSY correlation experiments were used to confirm the spectral assignments. 1–3 are suitable catalyst precursors for the transfer hydrogenation of acetophenone derivatives. Notably [Ru(Ph2PNHCH2‐C4H3S)(η6‐benzene)Cl2], 1, acts as an excellent catalyst, giving the corresponding alcohols in 98–99% yields in 30 min at 82 °C (TOF ≤200 h?1) for the transfer hydrogenation reaction in comparison to analogous rhodium or iridium complexes. This transfer hydrogenation is characterized by low reversibility under these conditions. Copyright © 2011 John Wiley & Sons, Ltd.  相似文献   

11.
Ruthenium catalysts supported on zinc-promoted amorphous-niobium mixed oxides were prepared, characterized, and studied in the additive-free partial hydrogenation of benzene reaction. The amorphous matrix of Nb2O5 was responsible for a highly active Ru/Nb2O5 catalyst, although less selective than those containing zinc. The ZnO-containing supports were prepared by wet impregnation technique, followed by incipient wetness of ruthenium chloride salt. The catalysts were characterized by textural analysis, X-ray fluorescence, X-ray diffraction, H2 chemisorption, temperature-programmed reduction (TPR), Scanning electron microscopy, H2 temperature-programmed desorption, and X-ray photoelectron spectroscopy (XPS) of the calcined-reduced samples. Chlorine retention was observed on zinc-containing samples. An unexpected ZnNb2O6 oxide phase, ascribed to a selectivity increase with less activity loss, was obtained for the supports at lower temperatures than those related on the literature. A very complex electronic environment of Ru- and Zn-containing species interactions was observed by TPR. The presence of surface-reduced (Ru0) and partially reduced (Ruδ+) Ru species observed by XPS justified well, respectively, the activity and selectivity achieved with every catalyst. The addition of water as a solvent resulted in very constant yield to cyclohexene, as expected, despite activity diminution due to low solubility of the reactants.  相似文献   

12.
The electrochemical reduction of oxygen on binary Pt–Ru alloy deposited onto microporous–mesoporous carbon support was studied in 0.5 M H2SO4 solution using cyclic voltammetry, rotating disk electrode (RDE), and impedance method. The microporous–mesoporous carbon support C(Mo2C) with specific surface area of 1,990 m2?g?1 was prepared from Mo2C at 600 °C using the chlorination method. Analysis of X-ray diffraction, photoelectron spectroscopy, and high-resolution transmission electron microscopy data confirms that the Pt–Ru alloy has been formed and the atomic fraction of Ru in the alloy was ~0.5. High cathodic oxygen reduction current densities (?160 A?m?2 at 3,000 rev?min?1) have been measured by the RDE method. The O2 diffusion constant (1.9?±?0.3?×?10?5?cm2?s?1) and the number of electrons transferred per electroreduction of one O2 molecule (~4), calculated from the Levich and Koutecky–Levich plots, are in agreement with literature data. Similarly to the Ru/RuO2 system in H2SO4 aqueous solution, nearly capacitive behavior was observed from impedance data at very low ac frequencies, explained by slow electrical double-layer formation limited by the adsorption of reaction intermediates and products into microporous–mesoporous Pt–Ru–C(Mo2C) catalyst. All results obtained for C(Mo2C) and Pt–Ru–C(Mo2C) electrodes have been compared with corresponding data for commercial carbon VULCAN® XC72 (C(Vulcan)) and Pt–Ru–C(Vulcan) electrodes processed and measured in the same experimental conditions. Higher activity for C(Mo2C) and Pt–Ru–C(Mo2C) has been demonstrated.  相似文献   

13.
The effects of Ru on the self-reducibility of Ru-doped Ni/MgAl2O4 catalysts, which do not need pre-reduction treatment with H2, were investigated in the steam reforming of methane (SRM). The Ru-promoted Ni/MgAl2O4 catalysts with various amounts of Ru (0–0.5 wt%) were prepared by stepwise impregnation and co-impregnation methods using hydrotalcite-like MgAl2O4 support. For comparison, Ru/MgAl2O4 catalysts with the same amount of Ru were also prepared by the impregnation method. The catalysts were characterized by the N2-sorption, XRD, H2-TPR, H2-chemisorption, and XPS methods. Ni/MgAl2O4 catalyst in the presence of even the trace amount of Ru (Ru content ≥0.05 wt%) showed higher conversion without pre-reduction as compared to Ru/MgAl2O4 catalysts in SRM under the same conditions. The self-activation of Ru–Ni/MgAl2O4 catalysts is mainly attributed to the spillover of hydrogen, which is produced on Ru at first and then reduces NiO species under reaction conditions. Besides, Ru doping makes the reduction of NiO easier. The stepwise impregnated Ru/Ni/MgAl2O4 catalyst produced superior performance as compared to co-impregnated Ru–Ni/MgAl2O4 catalyst for SRM.  相似文献   

14.
The mononuclear amidinate complexes [(η6‐cymene)‐RuCl( 1a )] ( 2 ) and [(η6‐C6H6)RuCl( 1b )] ( 3 ), with the trimethylsilyl‐ethinylamidinate ligands [Me3SiC≡CC(N‐c‐C6H11)2] ( 1a ) and[Me3SiC≡CC(N‐i‐C3H7)2] ( 1b ) were synthesized in high yields by salt metathesis. In addition, the related phosphane complexes[(η5‐C5H5)Ru(PPh3)( 1b )] ( 4a ) [(η5‐C5Me5)Ru(PPh3)( 1b )] ( 4b ), and [(η6‐C6H6)Ru(PPh3)( 1b )](BF4) ( 5 ‐BF4) were prepared by ligand exchange reactions. Investigations on the removal of the trimethyl‐silyl group using [Bu4N]F resulted in the isolation of [(η6‐C6H6)Ru(PPh3){(N‐i‐C3H7)2CC≡CH}](BF4) ( 6 ‐BF4) bearing a terminal alkynyl hydrogen atom, while 2 and 3 revealed to yield intricate reaction mixtures. Compounds 1a / b to 6 ‐BF4 were characterized by multinuclear NMR (1H, 13C, 31P) and IR spectroscopy and elemental analyses, including X‐ray diffraction analysis of 1b , 2 , and 3 .  相似文献   

15.
In contrast to ruthenocene [Ru(η5‐C5H5)2] and dimethylruthenocene [Ru(η5‐C5H4Me)2] ( 7 ), chemical oxidation of highly strained, ring‐tilted [2]ruthenocenophane [Ru(η5‐C5H4)2(CH2)2] ( 5 ) and slightly strained [3]ruthenocenophane [Ru(η5‐C5H4)2(CH2)3] ( 6 ) with cationic oxidants containing the non‐coordinating [B(C6F5)4]? anion was found to afford stable and isolable metal?metal bonded dicationic dimer salts [Ru(η5‐C5H4)2(CH2)2]2[B(C6F5)4]2 ( 8 ) and [Ru(η5‐C5H4)2(CH2)3]2[B(C6F5)4]2 ( 17 ), respectively. Cyclic voltammetry and DFT studies indicated that the oxidation potential, propensity for dimerization, and strength of the resulting Ru?Ru bond is strongly dependent on the degree of tilt present in 5 and 6 and thereby degree of exposure of the Ru center. Cleavage of the Ru?Ru bond in 8 was achieved through reaction with the radical source [(CH3)2NC(S)S?SC(S)N(CH3)2] (thiram), affording unusual dimer [(CH3)2NCS2Ru(η5‐C5H4)(η3‐C5H4)C2H4]2[B(C6F5)4]2 ( 9 ) through a haptotropic η5–η3 ring‐slippage followed by an apparent [2+2] cyclodimerization of the cyclopentadienyl ligand. Analogs of possible intermediates in the reaction pathway [C6H5ERu(η5‐C5H4)2C2H4][B(C6F5)4] [E=S ( 15 ) or Se ( 16 )] were synthesized through reaction of 8 with C6H5E?EC6H5 (E=S or Se).  相似文献   

16.
Ru 前驱体对 Ru/MgO-CeO2 氨合成催化剂性能的影响   总被引:1,自引:0,他引:1  
王秀云  王榕  倪军  林建新  魏可镁 《催化学报》2010,31(12):1452-1456
 分别以 K2RuO4, Ru(Ac)3 和 RuCl3 为 Ru 前驱体, 制备了 Ru/MgO-CeO2 催化剂, 并运用 X 射线衍射、X 射线荧光光谱, CO 吸附、N2 物理吸附和 H2程序升温还原等技术对催化剂进行了表征, 考察了 Ru 前驱体对 Ru/MgO-CeO2 催化剂氨合成性能的影响. 结果表明, Ru 前驱体对载体 MgO-CeO2 和 Ru 的还原性能、氯残留量和催化剂比表面积的影响都很大, 从而导致催化剂的氨合成性能的不同. 其中以 K2RuO4 为 Ru 前驱体制备的催化剂的载体和 Ru 容易还原, 无氯离子, 且比表面积较高, 因而催化剂活性和氨合成转换频率较高. 在 10 MPa, 425 °C, 10 000 h-1 条件下, K2RuO4, Ru(Ac)3 和 RuCl3 作前驱体制备的催化剂上氨合成转换频率比为 1.33:1.05:1.  相似文献   

17.
Piano‐stool‐shaped platinum group metal compounds, stable in the solid state and in solution, which are based on 2‐(5‐phenyl‐1H‐pyrazol‐3‐yl)pyridine ( L ) with the formulas [(η6‐arene)Ru( L )Cl]PF6 {arene = C6H6 ( 1 ), p‐cymene ( 2 ), and C6Me6, ( 3 )}, [(η6‐C5Me5)M( L )Cl]PF6 {M = Rh ( 4 ), Ir ( 5 )}, and [(η5‐C5H5)Ru(PPh3)( L )]PF6 ( 6 ), [(η5‐C5H5)Os(PPh3)( L )]PF6 ( 7 ), [(η5‐C5Me5)Ru(PPh3)( L )]PF6 ( 8 ), and [(η5‐C9H7)Ru(PPh3)( L )]PF6 ( 9 ) were prepared by a general method and characterized by NMR and IR spectroscopy and mass spectrometry. The molecular structures of compounds 4 and 5 were established by single‐crystal X‐ray diffraction. In each compound the metal is connected to N1 and N11 in a k2 manner.  相似文献   

18.
Organometallic Ru(arene)–peptide bioconjugates with potent in vitro anticancer activity are rare. We have prepared a conjugate of a Ru(arene) complex with the neuropeptide [Leu5]‐enkephalin. [Chlorido(η6p‐cymene)(5‐oxo‐κO‐2‐{(4‐[(N‐tyrosinyl‐glycinyl‐glycinyl‐phenylalanyl‐leucinyl‐NH2)propanamido]‐1H‐1,2,3‐triazol‐1‐yl)methyl}‐4H‐pyronato‐κO)ruthenium(II)] ( 8 ) shows antiproliferative activity in human ovarian carcinoma cells with an IC50 value as low as 13 μM , whereas the peptide or the Ru moiety alone are hardly cytotoxic. The conjugation strategy for linking the Ru(cym) (cym=η6p‐cymene) moiety to the peptide involved N‐terminal modification of an alkyne‐[Leu5]‐enkephalin with a 2‐(azidomethyl)‐5‐hydroxy‐4H‐pyran‐4‐one linker, using CuI‐catalyzed alkyne–azide cycloaddition (CuAAC), and subsequent metallation with the Ru(cym) moiety. The ruthenium‐bioconjugate was characterized by high resolution top‐down electrospray ionization mass spectrometry (ESI‐MS) with regard to peptide sequence, linker modification and metallation site. Notably, complete sequence coverage was obtained and the Ru(cym) moiety was confirmed to be coordinated to the pyronato linker. The ruthenium‐bioconjugate was analyzed with respect to cytotoxicity‐determining constituents, and through the bioconjugate models [{2‐(azidomethyl)‐5‐oxo‐κO‐4H‐pyronato‐κO}chloride (η6p‐cymene)ruthenium(II)] ( 5 ) and [chlorido(η6p‐cymene){5‐oxo‐κO‐2‐([(4‐(phenoxymethyl)‐1H‐1,2,3‐triazol‐1‐yl]methyl)‐4H‐pyronato‐κO}ruthenium(II)] ( 6 ) the Ru(cym) fragment with a triazole‐carrying pyronato ligand was identified as the minimal unit required to achieve in vitro anticancer activity.  相似文献   

19.
Development of novel bioanalytical methods for monitoring of H2S is key toward understanding the physiological and pathological functions of this gasotransmitter in live organisms. A ruthenium(II)‐complex‐based luminescence probe, Ru‐MDB (MDB: 4’‐methyl‐[2,2’‐bipyridine]‐4‐yl)methyl 2‐((2,4‐dinitrophenyl)thio)benzoate), was developed by introducing a new H2S responsive masking moiety to a red‐emitting RuII luminophore. Cleavage of this masking group by a H2S‐triggered reaction leads to a luminescence “off–on” response. The long‐lived emissions of Ru‐MDB and its reaction product with H2S allowed quantitative detection of H2S in autofluorescence‐rich human sera and adult zebrafish organs using the time‐gated luminescence mode. Ru‐MDB exhibits red emission, a large Stokes shift, high specificity and sensitivity for H2S detection, and low cytotoxicity, which enables imaging and flow cytometry analysis of lysosomal H2S generation in live inflamed cells under drug stimulation. Monitoring of H2S in live Daphnia magna, zebrafish embryos, adult zebrafish, and mice, was conducted by in vivo imaging using Ru‐MDB as a probe.  相似文献   

20.
Several Ru(II) complexes (η5-C5H4CO2H)Ru(η2-L)I have been prepared by the hydrolysis of the ester linkage in (η5-C5H4CO2t-Bu)Ru(η2-L)Cl with trimethylsilyl iodide. The hydrides (η5-C5H4CO2H)Ru(η2-L)H may be prepared by reduction of the iodide complexes in KOH/MeOH solutions followed by acidification. Complexes with several chelating bisphosphine ligands have been prepared in this way. The carboxylate anions [(η5-C5H4CO2)Ru(η2-L)H] are readily protonated by weak acids to give the carboxyCp complexes. The pKa of the carboxy proton of (η5-C5H4CO2H)Ru(dppe)H (dppe = 1,2-bis(diphenylphosphino)ethane) is 11.3 in DMSO. Protonation of the neutral hydride complex (η5-C5H4CO2H)Ru(dppf)H gives the cationic dihydride (η5-C5H4CO2H)Ru(dppf)H+2; the dihydride structure has been confirmed by measuring the T1 of its 1H NMR hydride resonance over a range of temperatures. The oxidations of the halide complexes (η5-C5H4CO2H)Ru(dppf)I and (η5-C5H4CO2t-Bu)Ru(dppf)Cl (dppf = 1,1′-bis(diphenylphosphino)ferrocene) have been studied by cyclic voltammetry.  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