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 78 毫秒
1.
Decomposition of Mn3Mo2TeO12 during oxidation of toluene to benzaldehyde was observed. Depending on the surface composition of the initial catalyst, the decomposition leads to less active but highly selective MnMoTeO6 or to MnMoO4 which is not selective in toluene oxidation.
Mn3Mo2TeO12 . , MnMoTeO6, MnMoO6.
  相似文献   

2.
Abstract

Reaction of one half and one equivalents of H2O2 with K[RuIII(pdta-H)Cl].2H2O gives rise to the μ-peroxo complexes [RuIII (pdta-H)]2H2O2 and [RuIV(pdta-H)]2O2, respectively. Equilibrium constants for the formation of the various peroxo species were determined between pH 3-11, in the temperature range 283-313 K and with μ = 0.10 M in KC1. The existence of the various peroxo species was substantiated by potentiometry, spectrophotometry and electrochemical studies. Thermodynamic quantities associated with the formation of the (pdta)RuIII and (pdta)RuIV-μ-peroxo species and their hydrolysis products are reported.  相似文献   

3.
王红森 《中国化学》2003,21(10):1330-1334
Methanol oxidation on smooth Pt electrode modified with differ-ent coverage of Ru was studied using cyclic voltammetry and potential step combined with differential electrochemical mass spectroscopy. The current efficiency of formed CO2 was calcu-lated from faraday current and ion current of m/z = 44. The results show that Ru modified Pt electrode with the coverage of ca. 0.3 has the highest catalytic activity for methanol electroox-idation, i. e. faraday current and the current efficiency of CO2 at the low potentials reach to the maximum. In addition, Ru loses its co-catalytic properties at the high potentials.  相似文献   

4.
RuHCl(PPh3)3 reacts quantitatively with cycloheptatriene in CH2Cl2 at 35°C in 15 min to give Ru(η5-C7H9)Cl(PPh3)2 and PPh3. The major isomer adopts a conformation with inequivalent phosphorus ligands and no plane of symmetry through the C7H9 ligand, but rapid intramolecular scrambling with δG3 = 10.6 kcal mol−1 results in an averaged 1H, 13C, and 31P NMR spectrum at room temperature. RuHCl(PPh3)3 reacts with cyclohepta-1,3-diene to give initially Ru(η3-C7H11)Cl(PPh3)2, but in a subsequent reaction this is dehydrogenated to give Ru(η5-C7H9)Cl(PPh3)2.  相似文献   

5.
Electrochemical and photochemical properties of the tetrahedral cluster [Ru3Ir( 3-H)(CO)13] were studied in order to prove whether the previously established thermal conversion of this cluster into the hydrogenated derivative [Ru3Ir(-H)3(CO)12] also occurs by means of redox or photochemical activation. Two-electron reduction of [Ru3Ir( 3-H)(CO)13] results in the loss of CO and concomitant formation of the dianion [Ru3Ir( 3-H)(CO)12]2–. The latter reduction product is stable in CH2Cl2 at low temperatures but becomes partly protonated above 283K into the anion [Ru3Ir(-H)2(CO)12] by traces of water. The dianion [Ru3Ir( 3-H)(CO)12]2– is also the product of the electrochemical reduction of [Ru3Ir(-H)3(CO)12] accompanied by the loss of H2. Stepwise deprotonation of [Ru3Ir(-H)3(CO)12] with Et4NOH yields [Ru3Ir(-H)2(CO)12] and [Ru3Ir( 3-H)(CO)12]2–. Reverse protonation of the anionic clusters can be achieved, e.g., with trifluoromethylsulfonic acid. Thus, the electrochemical conversion of [Ru3Ir( 3-H)(CO)13] into [Ru3Ir(-H)3(CO)12] is feasible, demanding separate two-electron reduction and protonation steps. Irradiation into the visible absorption band of [Ru3Ir( 3-H)(CO)13] in hexane does not induce any significant photochemical conversion. Irradiation of this cluster in the presence of CO with irr>340nm, however, triggers its efficient photofragmentation into reactive unsaturated ruthenium and iridium carbonyl fragments. These fragments are either stabilised by dissolved CO or undergo reclusterification to give homonuclear clusters. Most importantly, in H2-saturated hexane, [Ru3Ir( 3-H)(CO)13] converts selectively into the [Ru3Ir(-H)3(CO)12] photoproduct. This conversion is particularly efficient at irr >340nm.  相似文献   

6.
The reaction of Ru3(CO)12 (1) with LiEt3BH at −78°C affords the transient cluster formyl complex Ru3(CO)11(CHO) (2) which is observed to decompose by CO loss to give the known hydride cluster Ru3(CO)11(H) (3) rapidly at temperatures above −50°C. The formyl cluster has been characterized by low temperature FT-IR, 1H and 13C NMR measurements. Formyl trapping experiments and the effect of Bu3SnH on the rate of formyl decomposition are briefly described.  相似文献   

7.
Platinum–ruthenium catalysts are widely used as anode materials in polymer electrolyte fuel cells (PEMFCs) operating with reformate gas and in direct methanol fuel cells (DMFCs). Ruthenium dissolution from the Pt–Ru anode catalyst at potentials higher than 0.5?V vs. DHE, followed by migration and deposition to the Pt cathode can give rise to a decrease of the activity of both anode and cathode catalysts and to a worsening of cell performance. A major challenge for a suitable application of Pt–Ru catalysts in PEMFC and DMFC is to improve their stability against Ru dissolution. The purpose of this paper is to provide a better knowledge of the problem of Ru dissolution from Pt–Ru catalysts and its effect on fuel cell performance. The different ways to resolve this problem are discussed.  相似文献   

8.
The reaction of dichlorodifluoromethane and hydrogen has been studied in the gas phase at temperatures 438–538 K and atmospheric pressure over Pd and Ru supported AlF3 catalysts prepared by sol–gel method. For the hydrogenation of CF2Cl2, CH2F2 and CH4 represented more than 97% of the products. The catalytic properties of the catalysts are unchanged with time and they showed no significant difference in their activities. At the steady state, the kinetics of the reaction described by a mechanism of a halogenation/dehalogenation of the Pd and Ru surfaces by CF2Cl2 and H2, respectively. The values of the respective rate constants were then determined. It was concluded that at 448 K, the interaction between the Pd and Ru surfaces with CF2Cl2 or H2 is of the same order of magnitude. The conversion ratio on Ru/Pd supported catalysts within the temperature range used was increased from 1.5 to 4.1, while the selectivity of CH2F2/CH4 ratio was decreased from about 17.4 to 1.8 on the surfaces of both catalysts. This leads to the proposition that the high dispersion of Pd and Ru over the support are responsible for the high activity and high selectivity in CH2F2.  相似文献   

9.
It is important to understand the chemisorption of oxygen and CO on Ru(0001) surface. CO oxidation at oxygen precovered Ru(0001) surface at low oxygen coverages gave an extremely low CO oxidation rate, and it was also observed that, with a nominal oxygen coverage exceeding ca. 3 mL, rather high CO/CO2 conversion probabilities were achieved1. In the case of coadsorption of CO and oxygen on Ru(0001) surface under UHV conditions, a model comprising two CO molecules in an (22)-O unit cel…  相似文献   

10.
The reactions of Ru3(CO)12with 4-phenylbut-3-an-2-ine (1a), 3-phenyl-1-p-tolylprop-2-an-1-ine (1b), and 1,3-diferrocenylprop-2-an-1-ine (1c) afforded the Ru2(CO)6(-H)(O=C(R1)C(H)=C(R2)) (2) and Ru3(CO)8(O=C(R1)C(H)=C(R2))2(3) complexes. Dissolution of these complexes in CHCl3or CH2Cl2gave rise to the Ru2(CO)4(-Cl)2(O=C(R1)C(H)=C(R2)) complexes (4). The thermal transformations of complexes 2and 3in the presence of an excess of the ligand yielded the Ru2O2(CO)4(3-OC(R1)C(H)(CH2R2)C(R2)C(H)C(R1))2(5) and Ru(CO)2(O=C(R1)C(H)=C(R2))2(6) complexes. Analogous complexes were obtained upon more prolonged heating of the starting reaction mixtures. The structures of complexes 4a, 5a, and 6cwere established by X-ray diffraction analysis and confirmed by spectroscopic data.  相似文献   

11.
Following up on an earlier theoretical report by Knapp-Mohammady (Phys Lett A 372:1881–1884, 2008) on the ground state of the neutral Ru complex NAMI-A (trans-imidazoledimethylsulfoxide-tetrachlororuthenate), we first report here a quantum-chemical study of the effect of both oxidation and reduction of the parent molecule to form the anionic and cationic species. The new structures are compared with the equilibrium nuclear structure reported earlier for the neutral complex. We anticipate that one such Ru cluster, with potential as an anti-cancer drug, will interact via an appropriate receptor, rather than directly with DNA. A receptor for NAMI-A binding in here proposed to be the steroid hormone, estrogen, C18H24O2. The biomolecular structure of the dicomplex is predicted from restricted Hartree–Fock theory and density functional theory (DFT) calculations. The vibrational frequencies of NAMI-A and the dicomplex with estrogen are also reported. Some maps of the ground-state electron-density for the three neutral biomolecular species are finally presented. The use of vibrational spectroscopy, vibrational absorption (VA) and vibrational circular dichorism (VCD) are advocated to be measured, simulated and be used to understand the nature of the interaction of the Ru complex NAMI-A in complex with estrogen. Our aim in presenting these spectral simulations is to motivate the measurement of the VA and VCD spectra of estrogen, the Ru complex NAMI-A and finally of the estrogen–Ru NAMI-A complex. It should also be instructive to measure the VA and VCD spectra of estrogen and the estrogen receptor, both alone, together and finally together in the presence of the Ru NAMI-A complex to substantiate our claim that the Ru complex NAMI-A ties up estrogen, and hence prevents estrogen binding to the estrogen receptor.  相似文献   

12.
The reaction of the iron metalate [Fe{Si(OMe)3}(CO)3(dppm-P)] (1b) with [Ru(CO)3Cl(μ-Cl)]2 afforded the heterodinuclear complex [(OC)3{(MeO)3Si}Fe(μ-dppm)Ru(CO)3Cl)] (Fe–Ru) (3) in which a long Fe–Ru separation of 2.956(1) ? has been crystallographically evidenced. It was shown by density functional theory (DFT) calculations to correspond to the minimum-energy structure. Upon treatment of the corresponding hydrido complex [HFe{Si(OMe)3}(CO)3(dppm-P)] (1a) with [Ru(CO)3Cl(μ-Cl)]2, the chloride-bridged tetranuclear hydrido complex [(OC)3{(MeO)3Si}Fe(H)(μ-dppm)Ru(CO)2Cl(μ-Cl)]2 (4) was formed in which the Fe and Ru centers are only linked via bridging dppm ligands. Electronic supplementary material  The online version of this article (doi:) contains supplementary material, which is available to authorized users.
Pierre Braunstein (Corresponding author)Email:
  相似文献   

13.
《Solid State Sciences》2012,14(8):1050-1054
In this study, we successfully fabricated PtRu/Ru core–shell nanowires (NWs) prepared from as-spun Ru/Pt core–shell NWs via a co-electrospinning method. Their formation mechanism together with the structural characteristics, morphology, and composition of the resulting PtRu/Ru core–shell NWs was elucidated by means of scanning electron microscopy (SEM), transmission electron microscopy (TEM) with energy dispersive X-ray spectroscopy (EDS), and X-ray diffraction (XRD). PtRu/Ru core–shell NWs fabricated from as-spun Ru/Pt core–shell NWs were formed as a result of interdiffusion between Ru atoms and Pt atoms during calcination after co-electrospinning.  相似文献   

14.
Protonation of RuRhH2Cl(COD)(dppm)2 (1) (COD = 1,5 cyclooctadiene, dppm = bisdiphenylphosphinomethane) with HBF4 · Et2O gives [RuRhHCl(COD)-(dppm)2]BF4 (2), which has been shown to contain two chelating dppm ligands on ruthenium and a bridging hydride (RuH 2.08(7) Å; RhH 1.64(8) Å). Complex 2 reacts with CO to give [RuRhHCl(CO)3(dppm)2]BF4 (3) containing two bridging dppm groups. Reaction of 1 with 0.5 molar equivalents of [RhCl(COD)]2 at 80°C affords the trinuclear RuRh2H2Cl(PhPCH2PPh2)(COD)2(dppm) (4) in low yield (25%), and that with CuCl at room temperature gives RuRhCuH2Cl2 (COD)(dppm)2 (5) in high yield. Complex 5 is not very stable in solution and is converted into RuCuH2Cl(dppm)2 (6), a typical adduct between a Lewis acid and a hydride complex, which can be more easily obtained from RuH2(dppm)2 and CuCl in toluene at 80°C.  相似文献   

15.
A ruthenium-catalyzed asymmetric arylation of aliphatic aldehydes and α-ketoesters with arylboronic acids has been developed, giving chiral alkyl(aryl)methanols and α-hydroxy esters in good yields. The use of a chiral bidentate phosphoramidite ligand (Me-BIPAM) achieved excellent enantioselectivities.  相似文献   

16.
The photochemistry of the tris-substituted clusters Ru3(CO)9(PR3)3 (R=Ph or OMe) with no added ligands, with CO, C2H4, alkynes and H2 is compared and contrasted with results obtained for analogous thermal reactions. Photolysis of a CH2Cl2 solution of Ru3(CO)9(PPh3)3 leads to the metallated complex HRu3(CO)8(PPh3)2(PPh2C6H4). In CCl4, Ru(CO)3(PR3)Cl2 is formed on photolysis of Ru3(CO)9(PR3)3. Photolysis of CO saturated solutions of Ru3(CO)9(PR3)3 leads to Ru(CO)4(PR3). C2H4 saturated solutions of Ru3(CO)9(PR3)3 generate the novel Ru(CO)3(PR3)(2-C2H4) complexes upon photolysis. With C2H2, photolysis of solutions of Ru3(CO)9(PR3)3 leads to the novel complexes Ru(CO)3(PR3)(2-C2H2). Substituted alkyne complexes have been prepared. Thermolysis of Ru3(CO)9(PR3)3 with HCCPh leads to the novel acetylide clusters HRu3(CO)6(PR3)3(3-2-C2Ph). With PhC CPh, only Ru3(CO)9{P(OMe)3}3 reacts, yielding the novel alkyne cluster Ru3(CO)6{P(OMe)3}3(3-2-C2Ph2). With H2, photolysis of CH2Cl2 solutions of Ru3(CO)9(PR3)3 leads to H2Ru(CO)2(PR3)2. Irradiating a 4:1 CH2Cl2 to EtOAc solution of Ru3(CO)9(PR3)3 under an atmosphere of H2 leads to the novel dihydrido species H2Ru3(CO)7(PR3)3. Thermolysis of H2 saturated solutions of Ru3(CO)9(PR3)3 leads to H4Ru4(CO)8(PR3)4.  相似文献   

17.
Addition of ruthenium compounds (Ru(acac)3, Ru3(CO)12) to cobalt catalysts for pent-3-ene nitrile alkoxycarbonylation increases both activity and selectivity in the production of cyanoesters by (i) reducing the amount of Co(II) and facilitating the generation of the active carbonylation species, HCo(CO)n, and (ii) improving the isomerisation of pentene nitriles, providing significant amounts of pent-4-ene nitrile for the formation of the ω-ester.  相似文献   

18.
The reaction of Ru3(CO)12 and [Ir(CO)4]- (as [PPh4]+ or [N(PPh3)2]+ salts) yields the anion [Ru3Ir2(CO)14]2- (1) which has been found to derive from the intermediate [Ru3Ir(CO)13]- anion. Treatment of (1) with acids gives the conjugated hydrido species [Ru3Ir2(CO)14H]- (2). The two anions were characterized by single-crystal X-ray diffraction of their [PPh4]+ salts. [PPh4]2[Ru3Ir2(CO)14]: space group C2/c, Z=4, a=22.121(5) Å, b=10.546(5) Å, c=25.931(5) Å, =103.870(5)°, R=0.052 and Rw=0.130 for 3128 independent reflections with I>2(I ). [PPh4][Ru3Ir2(CO)14H]: space group P21/c, Z=8, a=22.833(5) Å, b=13.893(5) Å, c=25.810(5) Å, =92.650(5)°, R=0.070 and Rw=0.150 for 12141 independent reflections with I>2(I). Both anions 1 and 2 have a trigonal bipyramidal metal frame. There are two independent anions in the asymmetric unit of 2 differing in their ligand stereochemistry.  相似文献   

19.
The free energies interconnecting nine tungsten complexes have been determined from chemical equilibria and electrochemical data in MeCN solution (T = 22 °C). Homolytic W-H bond dissociation free energies are 59.3(3) kcal mol(-1) for CpW(CO)(2)(IMes)H and 59(1) kcal mol(-1) for the dihydride [CpW(CO)(2)(IMes)(H)(2)](+) (where IMes = 1,3-bis(2,4,6-trimethylphenyl)imidazol-2-ylidene), indicating that the bonds are the same within experimental uncertainty for the neutral hydride and the cationic dihydride. For the radical cation, [CpW(CO)(2)(IMes)H](?+), W-H bond homolysis to generate the 16-electron cation [CpW(CO)(2)(IMes)](+) is followed by MeCN uptake, with free energies for these steps being 51(1) and -16.9(5) kcal mol(-1), respectively. Based on these two steps, the free energy change for the net conversion of [CpW(CO)(2)(IMes)H](?+) to [CpW(CO)(2)(IMes)(MeCN)](+) in MeCN is 34(1) kcal mol(-1), indicating a much lower bond strength for the 17-electron radical cation of the metal hydride compared to the 18-electron hydride or dihydride. The pK(a) of CpW(CO)(2)(IMes)H in MeCN was determined to be 31.9(1), significantly higher than the 26.6 reported for the related phosphine complex, CpW(CO)(2)(PMe(3))H. This difference is attributed to the electron donor strength of IMes greatly exceeding that of PMe(3). The pK(a) values for [CpW(CO)(2)(IMes)H](?+) and [CpW(CO)(2)(IMes)(H)(2)](+) were determined to be 6.3(5) and 6.3(8), much closer to the pK(a) values reported for the PMe(3) analogues. The free energy of hydride abstraction from CpW(CO)(2)(IMes)H is 74(1) kcal mol(-1), and the resultant [CpW(CO)(2)(IMes)](+) cation is significantly stabilized by binding MeCN to form [CpW(CO)(2)(IMes)(MeCN)](+), giving an effective hydride donor ability of 57(1) kcal mol(-1) in MeCN. Electrochemical oxidation of [CpW(CO)(2)(IMes)](-) is fully reversible at all observed scan rates in cyclic voltammetry experiments (E° = -1.65 V vs Cp(2)Fe(+/0) in MeCN), whereas CpW(CO)(2)(IMes)H is reversibly oxidized (E° = -0.13(3) V) only at high scan rates (800 V s(-1)). For [CpW(CO)(2)(IMes)(MeCN)](+), high-pressure NMR experiments provide an estimate of ΔG° = 10.3(4) kcal mol(-1) for the displacement of MeCN by H(2) to give [CpW(CO)(2)(IMes)(H)(2)](+).  相似文献   

20.
The title complex [Cp*Ru(η6-C6H5BPh3)] has been synthesized by the reaction of [Cp*Ru(H2O)(NBD)]BF4 with H2 and NaBPh4, and its crystal structure was determined by singlecrystal X-ray diffraction analysis. It crystallizes in triclinic, space group P(1) with a = 11.0610(10), b = 11.2317(10), c = 12.3633(11) (A), α = 81.419(2), β = 67.8370(10), γ = 88.370(2)°, V= 1405.9(2) (A)3,Z= 2, C34H35BRu, Mr= 555.50, Dc = 1.312 g/cm3, F(000) = 576 and μ(MoKa) = 0.577 mm-1. The final R and wR are 0.0559 and 0.1483, respectively for 4365 observed reflections with I > 2σ(Ⅰ). In the title complex, the four phenyl rings bonded to the B atom are deposited in a tetrahedral geometry,and one of the phenyl rings is η6-bonded to ruthenium.  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