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 593 毫秒
1.
Base-assisted reduction of [Ru(CO)3Cl2]2 in the presence of NP-Me2 (2,7-dimethyl-1,8-naphthyridine) in thf provides an unsupported diruthenium(I) complex [Ru2(CO)4Cl2(NP-Me2)2] (1). Two NP-Me2 and four carbonyls bind at equatorial positions and two chlorides occupy sites trans to the Ru-Ru single bond. Reaction of [Ru(CO)3Cl2]2, TlOTf, KOH and NP-Me2 in acetonitrile, in a sealed container, affords a bicarbonate bridged diruthenium(I) complex [Ru2(CO)2(μ-CO)2(μ-O2COH)(NP-Me2)2](OTf) (2). The in situ generated CO2 is the source for bicarbonate under basic reaction medium. Isolation of 2 validates the decarboxylation step in the base-assisted reduction of [RuII(CO)3Cl2]2 → [RuI2(CO)4]2+.  相似文献   

2.
The synthesis of the diruthenium(II) compounds [Ru2(μ-O2CR)4(MeOH)2] [R = Me (1), Ph (2), CMePh2 (3) C6H4-p-OMe (4), C6H4-p-CMe3 (5)] by reaction of with hydroquinone, under a nitrogen atmosphere, in the presence of a base is described. This reaction constitutes an easy via to the preparation of diruthenium(II) compounds. The structure of the complexes [Ru2(μ-O2CMe)4(MeOH)2] (1) and [Ru2(μ-O2CPh)4(thf)2] (2b) is established by single crystal X-ray diffraction. These compounds show a diruthenium(II) unit bridged by four carboxylate ligands with the axial positions occupied by methanol and tetrahydrofuran molecules for 1 and 2b, respectively. Complex 1 shows, in the solid state, polymeric chains in which the molecules [Ru2(μ-O2CMe)4(MeOH)2] are linked by hydrogen bonds.  相似文献   

3.
A series of complexes containing the bulky carboxylate ligand 2,4,6-triisopropylbenzoate (TiPB) of type trans-[Ru2(TiPB)2(O2CCH3)2X] [X = Cl (1), PF6 (2)] and [Ru2(TiPB)4X] [X = Cl (3), PF6 (4)] have been synthesised. The corresponding complexes trans-[Ru2(TiPB)2(O2CCH3)2] (5) and [Ru2(TiPB)4] (6) were also isolated. Magnetic susceptibility measurements indicate that the diruthenium cores have the expected three (1-4) or two (5 and 6) unpaired electrons consistent with σ2π4δ2π)3 and σ2π4δ2δ∗2π∗2 electronic configurations. Compounds 1-4 and 6 were structurally characterised by X-ray crystallography, and show the expected paddlewheel arrangement of carboxylate ligands around the diruthenium core. The diruthenium cores of complexes 3, 4 and 6 are all distorted to minimise steric interactions between the bulky carboxylate ligands. The Ru-Ru bond length in the complex 6 [2.2425(6) Å] is the shortest observed for a diruthenium tetracarboxylate and, surprisingly, is 0.014 Å shorter than in the analogous complex 4, despite an increase in the formal Ru-Ru bond order from 2.0 (6) to 2.5 (4). This is rationalised in terms of the extent of internal rotation, or distortion, about the diruthenium core. This was supported by density functional theory calculations on the model complexes [Ru2(O2CH)4] and [Ru2(O2CH)4]+, that demonstrate the relationship between Ru-Ru bond length and internal rotation. Electrochemical and electronic absorption data were recorded for all complexes in solution. Comparison of the data for the ‘bis-bis’ (1, 2 and 5) and tetra-substituted (3, 4 and 6) complexes indicates that the shortening of the Ru-Ru bond length results in a small increase in energy of the near-degenerate δ and π orbitals.  相似文献   

4.
The activity of homobimetallic ruthenium alkylidene complexes, [(p-cymene)Ru(Cl)(μ-Cl)2Ru(Cl)(CHPh)(PCy3)] [Ru-I] and [(p-cymene)Ru(Cl)(μ-Cl)2Ru(Cl)(CHPh)(IPr)] [Ru-II], on intermolecular [2+2+2] cyclotrimerisation reactions of monoynes has been investigated for the first time. It was found that these complexes can catalyse the chemo and regioselective cyclotrimerisation reactions of alkynes at both 25 and 50 °C in polar, aprotic solvents. The catalytic activity of [Ru-I] and [Ru-II] was compared to other well-known ruthenium catalysts such as Grubbs first generation catalyst [RuCl2(CHPh)(PCy3)2] [Ru-III], [RuCl(μ-Cl)(p-cymene)]2 [Ru-IV] and [RuCl2(p-cymene)PCy3] [Ru-V] complexes. To examine the effect of the steric hinderance of substrates on the regioselectivity of the reaction, a series of sterically hindered silicon containing alkynes (1a, 1b, 1c) were used. It was shown that the isomeric product distribution of the reaction shifts from 1,2,4-trisubstituted arenes to 1,3,5-trisubstituted arenes as the steric hinderance on the substrates increases. These homobimetallic ruthenium alkylidene complexes also catalysed regio- and chemo-selective cross-cyclotrimerisation reactions between silicon-containing alkynes (1a, 1b, 1c) and aliphatic alkynes (1d-g).  相似文献   

5.
Treatment of NaO2CCHC(Me)Fc with cadmium acetate and iron(II) sulfate in the presence of 2,2′-bipy yielded [Cd2Fe(μ-O2CCHC(Me)Fc)22-O2CCHC(Me)Fc)222-O2CCHC(Me)Fc)2(2,2′-bipy)2] · 2H2O (1); while from NaO2CC6H4{C(O)Fc-o}, cadmium acetate, and pbbm the product was {[Cd(η2-O2CC6H4{C(O)Fc-o})2(pbbm)] · 0.5H2O}n (2) [Fc = (η5-C5H5)Fe(C5H45); 2,2′-bipy = 2,2′-bipyridyl; pbbm = 1,1′-(1,5-pentamethylene)bis-1H-benzimidazole]. Compounds 1 and 2 have been characterized by elemental analysis, IR spectroscopy and single crystal X-ray diffraction. In centro-symmetric crystalline 1, the Fe and the two flanking atoms are six-coordinate; the three carboxylato ligands between the Fe and a Cd atom have different coordination modes. Crystalline 2 consists of an infinite polymeric chain, in which adjacent [Cd(η2-O2CC6H4{C(O)Fc-o})2] units are linked by pbbm ligands; thus each Cd atom is six-coordinate. Some electrochemical properties of the two complexes are reported.  相似文献   

6.
Seven diiridium(II) complexes were synthesized by ligand substitution reactions of [Ir2(μ-O2CMe)2Cl2(CO)2] (1) and [Ir2(μ-O2CMe)2Cl2(CO)2(py)2] (2).The reaction of 2 with the silver salt of a less coordinating anion, AgSbF6, gave a cationic complex [Ir2(μ-O2CMe)2Cl(CO)2(py)3]SbF6 (3).A tricarbonyl cationic complex [Ir2(μ-O2CMe)2(CO)3Cl(py)2]SbF6 (4) was obtained under a CO atmosphere.Complex 2 reacted with AgO2CCF3 to give [Ir2(μ-O2CMe)2Cl(O2CCF3)(CO)2(py)2] (5) in toluene.[Ir2(μ-hiq)2(CO)2Cl2] (Hhiq = 1-hydroxyisoquinoline, 6) was synthesized by the bridging-ligand substitution of 1 with Hhiq.Its axial adducts [Ir2(μ-hiq)2Cl2(CO)2L2] (L = Mepy (4-methylpyridine), 7 or PPh3, 8) were synthesized by addition of the ligands to a suspension of 6.In the structures of 7 and 8, two iridium atoms are bridged by two hiq ligands in a head-to-tail arrangement.The reaction of 1 with Hmhp (2-hydroxy-4-methylpyridine) led to triply bridged [Ir2(μ-mhp)3(CO)2Cl(Hmhp)] (9).In complex 9, all the mhp ligands bridge between the Ir atoms in a head-to-head manner.The Ir-Ir distances of 3, 4, 5, 7 and 8 are 2.6047(7), 2.6216(9), 2.5899(9), 2.5933(5) and 2.634(2) Å, respectively, which are similar to those observed in[Ir2(μ-O2CMe)2Cl2(CO)2L2]. The Ir-Ir distance of 2.5512(4) Å in 9 is shorter than in the other complexes.  相似文献   

7.
Subsequent addition of 1,2-benzenedithiol (S2-H2) and nBuLi to a solution of [Ru(NO)Cl3 · xMeOH] in THF afforded exclusively the monomeric species NBu4[RuII(NO)(S2)2] (1). Formation of dimeric (NBu4)2[RuII(NO)(S2)2]2 (2) has been confirmed when the deprotonated ligand S2-Li2 was added to [Ru(NO)Cl3 · xMeOH] and allowed to stir for 30 h. The monomer 1 undergoes aerial oxidation to give (NBu4)2[RuIV(S2)3] (3). The reaction between RuCl3 · xH2O and S2-H2 in the presence of NaOMe, afforded the dinulear RuIII species (NMe4)2[RuIII(S2)2]2 (4). A modified method for the preparation of 1 is being employed to synthesize the osmium analogue NBu4[Os(NO)(S2)2] (5) effectively. The solid state structures of 1, 2 and 3 were determined by X-ray crystal structure analysis. A comparison of relevant bond distance data suggests that 1,2-benzenedithiolate acts as an “innocent” ligand.  相似文献   

8.
Crystallisation of simple cyanoruthenate complex anions [Ru(NN)(CN)4]2− (NN = 2,2′-bipyridine or 1,10-phenanthroline) in the presence of Lewis-acidic cations such as Ln(III) or guanidinium cations results, in addition to the expected [Ru(NN)(CN)4]2− salts, in the formation of small amounts of salts of the dinuclear species [Ru2(NN)2(CN)7]3−. These cyanide-bridged anions have arisen from the combination of two monomer units [Ru(NN)(CN)4]2− following the loss of one cyanide, presumably as HCN. The crystal structures of [Nd(H2O)5.5][Ru2(bipy)2(CN)7] · 11H2O and [Pr(H2O)6][Ru2(phen)2(CN)7] · 9H2O show that the cyanoruthenate anions form Ru-CN-Ln bridges to the Ln(III) cations, resulting in infinite coordination polymers consisting of fused Ru2Ln2(μ-CN)4 squares and Ru4Ln2(μ-CN)6 hexagons, which alternate to form a one-dimensional chain. In [CH6N3]3[Ru2(bipy)2(CN)7] · 2H2O in contrast the discrete complex anions are involved in an extensive network of hydrogen-bonding involving terminal cyanide ligands, water molecules, and guanidinium cations. In the [Ru2(NN)2(CN)7]3− anions themselves the two NN ligands are approximately eclipsed, lying on the same side of the central Ru-CN-Ru axis, such that their peripheries are in close contact. Consequently, when NN = 4,4′-tBu2-2,2′-bipyridine the steric bulk of the t-butyl groups prevents the formation of the dinuclear anions, and the only product is the simple salt of the monomer, [CH6N3]2[Ru(tBu2bipy)(CN)4] · 2H2O. We demonstrated by electrospray mass spectrometry that the dinuclear by-product [Ru2(phen)2(CN)7]3− could be formed in significant amounts during the synthesis of monomeric [Ru(phen)(CN)4]2− if the reaction time was too long or the medium too acidic. In the solid state the luminescence properties of [Ru2(bipy)2(CN)7]3− (as its guanidinium salt) are comparable to those of monomeric [Ru(bipy)(CN)4]2−, with a 3MLCT emission at 581 nm.  相似文献   

9.
The organotin complex [Ph3SnS(CH2)3SSnPh3] (1) was synthesized by PdCl2 catalyzed reaction between Ph3SnCl and disodium-1,3-propanedithiolate which in turn was prepared from 1,2-propanedithiol and sodium in refluxing THF. Reaction of 1 with Ru3(CO)12 in refluxing THF affords the mononuclear complex trans-[Ru(CO)4(SnPh3)2] (2) and the dinuclear complex [Ru2(CO)6(μ-κ2-SCH2CH2CH2S)] (3) in 20 and 11% yields, respectively, formed by cleavage of Sn-S bond of the ligand and Ru-Ru bonds of the cluster. Treatment of pymSSnPPh3 (pymS = pyrimidine-2-thiolate) with Ru3(CO)12 at 55-60 °C also gives 2 in 38% yield. Both 1 and 2 have been characterized by a combination of spectroscopic data and single crystal X-ray diffraction analysis.  相似文献   

10.
The dihydroxo-bridged dinuclear copper(II) compound [Cu2(dpyam)2(μ-OH)2]I2 (1) and the triply bridged dinuclear copper(II) compounds with a formato bridge [Cu2(dpyam)2(μ-O2CH)(μ-OH)(μ-OMe)](ClO4) (2) and [Cu2(dpyam)2(μ-O2CH)(μ-OH)(μ-Cl)](ClO4) · 0.5H2O (3) (in which dpyam=di-2-pyridylamine) have been synthesized and their crystal structures determined by X-ray crystallographic methods. All three compounds are either centrosymmetric, or have a symmetry plane in the molecule. Compound 1 contains the [Cu2(dpyam)2(μ-OH)2]+ unit and iodide anions. Each copper(II) ion is in a slightly tetrahedrally distorted square planar coordination with the square plane consisting of two nitrogen atoms of the dpyam ligand and two bridging hydroxo groups. The Cu-I distances of 3.321 Å are quite long and only involve a weak semi-coordination. Compound 2 contains a triply bridged dinuclear copper(II) species, the coordination environment around each copper(II) ion involves a distorted trigonal-bipyramidal CuN2O3 chromophore. In the dinuclear unit of compound 3, the triply bridged copper(II) ions show a distorted trigonal-bipyramidal coordination of the CuN2O2Cl chromophore. The Cu-Cu distances are 2.933(2), 3.023(1) and 3.036(1) Å for compounds 1, 2 and 3, respectively.The magnetic susceptibility measurements, measured from 5 to 280 K, revealed a weak antiferromagnetic interaction between the Cu(II) atoms for compound 1 with a singlet-triplet energy gap (J) of −15.3 cm−1, whereas compounds 2 and 3 are ferromagnetic with J=62.5 and 79.1 cm−1, respectively.  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.
The dinuclear bis(6-X-pyridin-2-olato) ruthenium complexes [Ru2(μ-XpyO)2(CO)4(PPh3)2] (X = Cl (4B) and Br (5B)), [Ru2(μ-XpyO)2(CO)4(CH3CN)2] (X = Cl (6B), Br (7B) and F (8B)) and [Ru2(μ-ClpyO)2(CO)4(PhCN)2] (9B) were prepared from the corresponding tetranuclear coordination dimers [Ru2(μ-XpyO)2(CO)4]2 (1: X = Cl; 2: X = Br) and [Ru2(μ-FpyO)2(CO)6]2 (3) by treatment with an excess of triphenylphosphane, acetonitrile and benzonitrile, respectively. In the solid state, complexes 4B-9B all have a head-to-tail arrangement of the two pyridonate ligands, as evidenced by X-ray crystal structure analyses of 4B, 6B and 9B, in contrast to the head-to-head arrangement in the precursors 1-3. A temperature- and solvent-dependent equilibrium between the yellow head-to-tail complexes and the red head-to-head complexes 4A-7A and 9A, bearing an axial ligand only at the O,O-substituted ruthenium atom, exists in solution and was studied by NMR spectroscopy. Full 1H and 13C NMR assignments were made in each case. Treatment of 1 and 2 with the N-heterocyclic carbene (NHC) 1-butyl-3-methylimidazolin-2-ylidene provided the complexes [Ru2(μ-XpyO)2(CO)4(NHC)], X = Cl (11A) or Br (12A). An XRD analysis revealed the head-to-head arrangement of the pyridonate ligands and axial coordination of the carbene ligand at the O,O-substituted ruthenium atom. The conversion of 11A and 12A into the corresponding head-to-tail complexes was not possible.  相似文献   

13.
The first complexes that contain the 2,6-bis(dicyclohexylphosphinomethyl)pyridine ligand (PNP) have been isolated and characterized. The reactions of K4Mo2Cl8, (n-Bu4N)2Re2Cl8 and PdBr2(1,5-COD) afford Mo2Cl4(PNP)(HPCy2) (1), ReCl3(PNP) (2) and PdBr2(PNP) (4), respectively, while from the reaction of PNP with cis-Re2(μ-O2CCH3)2Cl4(H2O)2 the heteromacrocylic dication [Cy2P{CH2pyCH2}2PCy2]2+ has been isolated as its mixed [Cl]/[ReO4] salt (3). The reaction of cis-Re2(μ-O2CCH3)2Cl4(H2O)2 with bis(diphenylphosphinomethyl)sulfide (PSP) gives the mononuclear Re(V) complex ReO(OEt)Cl2(PSP) (5) in which the S atom is not coordinated. The structures of 1-5 have been established by X-ray crystallography, that of 5 being the first for a complex of this ligand.  相似文献   

14.
The nuclearity, bonding and H-bonded networks of copper(I) halide complexes with thiophene-2-carbaldehyde thiosemicarbazones {(C4H3S)HC2N3-N(H)-C1(S)N1HR} are influenced by R substituents at N1 atom. Thiophene-2-carbaldehyde-N1-methyl thiosemicarbazone (HttscMe) or thiophene-2-carbaldehyde-N1-ethyl thiosemicarbazone (HttscEt) have yielded halogen-bridged dinuclear complexes, [Cu2(μ-X)21-S-Htsc)2(Ph3P)2] (Htsc, X: HttscMe, I, 1; Br, 2; Cl, 3; HttscEt, I, 4; Br, 5; Cl, 6), while thiophene-2-carbaldehyde-N1-phenyl thiosemicarbazone (HttscPh) has yielded mononuclear complexes, [CuX(η1-S-HttscPh)2] (X, I, 7a; Br 8; Cl, 9) and a sulfur bridged dinuclear complex, [Cu2(μ-S-HttscPh)21-S-HttscPh)2I2] 7b co-existing with 7a in the same unit cell. These results are in contrast to S-bridged dimers [Cu2(μ-S-Httsc)21-Br)2(Ph3P)2] · 2H2O and [Cu2(μ-S-Httsc)21-Cl)2(Ph3P)2] · 2CH3CN obtained for R = H and X = Cl, Br (Httsc = thiophene-2-carbaldehyde thiosemicarbazone) as reported earlier. The intermolecular CHPh?π interaction in 1-3 (2.797 Å, 1; 3.264 Å, 2; 3.257 Å, 3) have formed linear polymers, whereas the CHPh?X and N3?HCH interactions in 4-6 (2.791, 2.69 Å, 5; 2.776, 2.745 Å, 6, respectively) have led to the formation of H-bonded 2D polymer. The PhN1H?π, interactions (2.547 Å, 8, 2.599 Å, 9) have formed H-bonded dimers only. The Cu?Cu separations are 3.221-3.404 Å (1-6).  相似文献   

15.
Three kinds of crystalline compounds containing the nitrosylpentaamminechromium complexes [Cr(NO)(NH3)5]2+(A) were obtained: chloride ACl2 (red-orange), chloride perchlorate ACl(ClO4) (brown), and perchlorate A(ClO4)2 (green). The cause of the color change of the complex A with the change of outer sphere anions was sought using X-ray structural data of ACl2, ACl(ClO4), and A(ClO4)2. Crystal data: ACl2, orthorhombic, space group Cmcm, a=10.0236 (9) Å, b=9.098 (3) Å, c=10.357(1) Å, V=944.5 (5) Å3, Z=4; ACl(ClO4), tetragonal, space group P4/nmm, a=7.6986 (8) Å, c=9.9566(8) Å, V=590.1 (1) Å3,Z=2; A(ClO4)2, orthorhombic, space group Pnma, a=15.760 (2) Å, b=11.480(2) Å, c=7.920 (2) Å, V=1432.9 (4) Å3, Z=4. The complex cation in ACl2 has a distorted octahedral structure with a linear CrNO moiety. The short CrN (nitrosyl) distance of 1.692 (7) Å indicates the presence of multiple bonding between the chromium atom and the nitrogen atom in the nitrosyl group. The interatomic distances and angles within the complex cations hardly change with the change of the counter anions, while the distances between the complex cations in each crystal increase in the order ACl2<ACl(ClO4)<A(ClO4)2. The bulky perchlorate anions seems to separate the complex cations, while smaller chloride anions are not large enough to separate them. The distance (3.213(5) Å) between O(NO) and N(NH3 in the adjacent complex cation) is rather short in the crystal of ACl2, and there are six hydrogen bonds, where the NO group is surrounded by four NH3 ligands. The distance (4.002(5) Å) between O(NO) and N(NH3) is much longer in the crystal of A(ClO4)2, indicating the presence of no hydrogen bonding. In the crystal of ACl(ClO4) the distance (3.452(4) Å) between O(NO) and N(NH3) is in between those of ACl2 and A(ClO4)2. The presence of hydrogen bonding between O(NO) and N(NH3 in the adjacent complex cation) seems to cause the color change with the change of outer sphere anions.  相似文献   

16.
The SS bond-activation of diorganyl disulfide by the anionic metal carbonyl fragment [Mn(CO)5] gives rise to an extensive chemistry. Oxidative decarbonylation addition of 2,2′-dithiobis(pyridine-N-oxide) to [Mn(CO)5], followed by chelation and metal-center oxidation, led to the formation of [MnII(SC5H4NO)3] (1). The effective magnetic moment in solid state by SQUID magnetometer was 5.88 μB for complex 1, which is consistent with the MnII having a high-spin d5 electronic configuration in an octahedral ligand field. The average Mn(II)S, SC and NO bond lengths of 2.581(1), 1.692(4) and 1.326(4) Å, respectively, indicate that the negative charge of the bidentate 1-oxo-2-thiopyridinato [SC5H4NO] ligand in complex 1 is mainly localized on the oxygen atom. The results are consistent with thiolate-donor [SC5H4NO] stabilization of the lower oxidation state of manganese (Mn(I)), while the O,S-chelating [SC5H4NO] ligand enhances the stability of manganese in the higher oxidation state (Mn(II)). Activation of SS bond as well as OH bond of 2,2′-dithiosalicylic acid by [Mn(CO)5] yielded [(CO)3Mn(μ-SC6H4C(O)O)2Mn(CO)3]2− (4). Oxidative addition of bis(o-benzamidophenyl) disulfide to [Mn(CO)5] resulted in the formation of cis-[Mn(CO)4(SR)2] (R=C6H4NHCOPh) which was employed as a chelating metallo ligand to synthesize heterotrinuclear [(CO)3Mn(μ-SR)3Co(μ-SR)3Mn(CO)3] (8) possessing a homoleptic hexathiolatocobalt(III) core.  相似文献   

17.
Molybdenum tetramers: Mo43-O)4[μ-O2P(CH2Cl)2]4O4 (1), Mo43-O)4(μ-O2P(CH2OH)2)4O4 (2), Mo43-O)4[μ-O2P(PhOMe)2]4O4 (3), and Mo43-O)4[μ-O2P(o-C6H4(CH2)2)]4O4 (4) have been synthesized and characterized by IR, UV-Vis, and 31P NMR spectroscopy. Molybdenum tetramers 1 and 4 along with the ligands L2A and L4 were structurally characterized by single crystal X-ray crystallography. An infinite 2D polymeric sheet was formed via inter and intra hydrogen bonds in the crystals of L2A. The crystals of L4 consist of infinite polymeric chains formed through hydrogen bonding. All molybdenum tetramers were tested as catalysts for the epoxidation of cis-cyclooctene in the presence of H2O2. Compounds 1 and 2 resulted in more than 80% epoxide after 24 hours at 70 °C, and displayed superior catalytic activities over compounds 3 and 4 under identical conditions. The superior catalytic activities of compounds 1 and 2 may be attributed to their better solubility in the ethanol/H2O2 system.  相似文献   

18.
Trirutheniumdodecacarbonyl (Ru3(CO)12) reacts with 2-hydroxy-6-methylpyridine and with 2-hydroxy-5,6,7,8-tetrahydroquinoline in toluene to form centrosymmetric tetranuclear complexes of the type [Ru(η2, μ-L)(CO)23-L)Ru(CO)2]2, where L is the respective (N,O)-pyridonate ligand (2 and 3). The structures of these complexes, which are almost insoluble in all common solvents, could be determined by single-crystal X-ray diffraction. Reaction of Ru3(CO)12 with 2-hydroxy-4,6-diphenylpyridine in methanol includes ortho-metallation at the phenyl ring, furnishing the dinuclear complex [Ru(κ2N,C-L)(CO)2(μ-OCH3)2Ru(CO)22N,C-L)] (4), where L = (2-(6-hydroxy-4-phenylpyridin-2-yl)phenyl), according to an X-ray crystal structure determination.  相似文献   

19.
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   

20.
We have used the elimination of AuX(PR3) (X = halide, R = Ph, tol) that occurs in reactions of alkynylgold(I)-phosphine complexes with M3(μ-H)33-CBr) (CO)9 (M = Ru, Os) to prepare the complexes M3(μ-H)33-CCCR)(CO)9 [M = Ru, R = Ph 2, CCSiMe33, Fc 4, CCFc 6-Ru, CC[Ru(PPh3)2Cp] 8; M = Os, R = CCFc 6-Os, CCCCFc 7], Fc′{(μ3-CCC)Ru3(μ-H)3(CO)9}25, and bis-cluster-capped carbon chain complexes {M3(μ-H)3(CO)9}233-C(CC)nC} (M = Ru, n = 2 9, 3 10-Ru; M = Os, n = 3 10-Os) and {(L)(OC)8(μ-H)3M3}C(CC)nC{Co3(μ-dppm)(CO)7} (n = 1, M = Ru, L = CO 11, PPh312-Ru/P; n = 2, L = CO 12-Ru, PPh313; M = Os, L = CO 12-Os) in good to excellent yields. X-ray structural determinations of 2-5, 6-Ru, 6-Os, 7, 9, 11, 12-Ru, 12-Os and 12-Ru/P are reported.  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