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.
The 16-electron, coordinatively unsaturated, dicationic ruthenium complex [Ru(P(OH)2(OMe))(dppe)2][OTf]2 (1a) brings about the heterolysis of the C-H bond in phenylacetylene to afford the phenylacetylide complex trans-[Ru(CCPh)(P(OH)2(OMe))(dppe)2][OTf] (2). The phenylacetylide complex undergoes hydrogenation to give a ruthenium hydride complex trans-[Ru(H)(P(OH)2(OMe))(dppe)2][OTf] (3) and phenylacetylene via the addition of H2 across the Ru-C bond. The 16-electron complex also reacts with HSiCl3 quite vigorously to yield a chloride complex trans-[Ru(Cl)(P(OH)2(OMe))(dppe)2][OTf] (4). On the other hand, the other coordinatively unsaturated ruthenium complex [Ru(P(OH)3)(dppe)2][OTf]2 (1b) reacts with a base N-benzylideneaniline to afford a phosphonate complex [Ru(P(O)(OH)2)(dppe)2][OTf] (5) via the abstraction of one of the protons of the P(OH)3 ligand by the base. The phenylacetylide, chloride, and the phosphonate complexes have been structurally characterized. The phosphonate complex reacts with H2 to afford the corresponding dihydrogen complex trans-[Ru(η2-H2)(P(O)(OH)2)(dppe)2][OTf] (5-H2). The intact nature of the H-H bond in this species was established using variable temperature 1H spin-lattice relaxation time measurements and the observation of a significant J(H,D) coupling in the HD isotopomer trans-[Ru(η2-HD)(P(O)(OH)2)(dppe)2][OTf] (5-HD).  相似文献   

2.
The synthesis and characterization of novel coordination polymers [Co(HCCB)(H2O)2]n (1), [Zn(HCCB)(H2O)2]n (2), {[Cd(HCCB)2]·0.5[Cd(μ-H2O)(H2O)4]2}n (3) and [Cu(HCCB)(H2O)2]n (4) based on 3-(carboxymethylamino)-4-chlorobenzoic acid (H3CCB) and mononuclear complexes [Cu(HBCCB)(H2O)]·H2O (5), [Co(HBCCB)(H2O)]·H2O (6), [Zn(HBCCB)(H2O)] (7) and [Cd(HBCCB)(H2O)] (8) containing 3-bis(carboxymethylamino)-4-chlorobenzoic acid (H3BCCB) have been described. The compounds under investigation have been characterized by elemental analyses, spectral studies and structures of 1-3 and 5 determined crystallographically. Structural data of 1 and 2 revealed that the deprotonated HCCB2− bridges metal centers leading to a double stranded 1D chain. On the other hand, the HCCB2− coordinated thorough carboxylate oxygen and amino nitrogen in 3 to afford a 1D chain whose charge neutrality is maintained by inclusion of aqua-bridged dimer [{Cd(μ-H2O)(H2O)4}2]4+. Strong Cu?Cl interaction (2.754 Å) in 5 imposes a coordination geometry that is half-way between the square planar and square pyramidal. The H3CCB, H3BCCB and 1-3 and 5 are fluorescent at rt. Thermal studies (TG and DSC) on 1-3 suggested higher stability of 2 relative to 1 and 3 [ΔHf (kcal/mol), ΔSf = 152.17, 0.60, 1; 195.56: 0.86, 2; 69.33:0.36, 3].  相似文献   

3.
The new pyridine-based NNN tridentate ligand 2,6-C5H3N(CMe2NH2)2 (1) was synthesized by the treatment of 2,6-pyridinedicarbonitrile with an excess of the organocerium reagent in situ generated from CeCl3 and methyllithium in THF. The reaction of 1 with [RuCl2(PPh3)3] in THF at ambient conditions afforded (OC-6-23)-[RuCl{2,6-C5H3N(CMe2NH2)2}(PPh3)2]Cl (2). The corresponding dimethyl sulfoxide complex [RuCl{2,6-C5H3N(CMe2NH2)2}{S(O)Me2}2]Cl (3) was isolated as a mixture of the (OC-6-23) and (OC-6-32) stereoisomers 3a and 3b from the reaction between 1 and (OC-6-22)-[RuCl2{S(O)Me2}3(OSMe2)] in toluene at 80 °C. A prolonged interaction in toluene at reflux temperature gave isomerically pure 3a. The metal trichloride hydrates MCl3 · xH2O (M = Ru, Rh, Ir; x ≅ 2-4) produced mer-[RuCl3{2,6-C5H3N(CMe2NH2)2}] (M = Ru: 4; Rh: 5; Ir: 6), when combined with 1 in refluxing ethanol. The crystal structures of the following compounds were determined: ligand 1 and complexes 2-5 as addition compounds 2 · CH2Cl2, 3a · C7H8, 4 · EtOH and .  相似文献   

4.
In a very acidic solution, potassium 1,3-propanediaminetetraacetate zinc chloride K2n[ZnCl2(1,3-H2pdta)ZnCl2]n (1) and its substituted iodide [ZnI2(H2O)(1,3-H4pdta)]n (2) (H4pdta = 1,3-propanediaminetetraacetatic acid, C11H18N2O8) were isolated. The former was obtained from the reaction of zinc chloride and H4pdta in pH ∼1.5. Further substitution of 1 results in the formation of iodide 2 with the addition of potassium iodide in acidic solution of pH 0.5. Complex 1 consists of a dimeric anionic unit [ZnCl2(1,3-H2pdta)ZnCl2]2− with strong intra-molecular hydrogen bonds [N1?O2 2.648(4); N1?O4 2.710(4) Å]. In neutral complex 2, an 1,3-pdta ligand links each monomeric unit [Zn(H2O)I2] to generate an infinite 1D chain, which extents into a 3D supramolecular structure by very strong inter-molecular hydrogen bonds the [O4?O2b 2.50 (1) Å, bx, y, z + 1]. 1 is soluble in water at room temperature, which is traced by 13C NMR experiment.  相似文献   

5.
Typically 2,2-diethylmalonate (dem) acts as a chelating ligand and binds to the metal in a η2 (dem-O, O′) mode. However, when cis,fac-[RuCl2(TMSO-S)4] is treated with K2(dem), it prefers to bind in an unusual bridging mode (μ-dem-O, O′) with the ruthenium (II) cation containing coordinated water, forming a strong hydrogen bond with the non-coordinated oxygen atoms of the 2,2-diethylmalonate ligand. The reaction products of cis,fac-[RuCl2(TMSO-S)4] (1) and cis,fac-[RuCl2(DMSO-S)3(DMSO-O)] (2) with dem are the dinuclear species with two bridging dem units, fac-[Ru(TMSO-S)3(H2O)(μ2-dem-O, O′)]2 (3) and fac-[Ru(DMSO-S)3(H2O)(μ2-dem-O, O′)]2 (4), respectively. The complex 3 was characterized by X-ray crystallography in which water ligands occupy anti positions with respect to each other. The NMR and X-ray study support each other with respect to dinuclear structure of 3 and 4, indicating that the dinuclear structure observed in the solid state is preserved in solution as well. The mononuclear anionic complex with chelating dem unit, K{fac-[RuCl(η2-dem-O, O′)(TMSO-S)3} (5), was also isolated from the reaction of 1 and K2(dem) demonstrating that 5 is an intermediate in the formation of 3.  相似文献   

6.
The reaction of [RuCl3(2mqn)NO] (H2mqn=2-methyl-8-quinolinol) with 2-chloro-8-quinolinol (H2cqn) afforded cis-1 [RuCl(2cqn)(2mqn)NO] (the oxygen of 2cqn is trans to the NO) (complex 1), cis-1 [RuCl(2cqn)(2mqn)NO] (the oxygen of 2mqn is trans to the NO) (complex 2) and a 1:1 mixture of cis-2 [RuCl(2cqn)(2mqn)NO] (the oxygen of 2mqn is trans to the NO) and cis-2 [RuCl(2cqn)(2mqn)NO] (the oxygen of 2cqn is trans to the NO) (complex 3). The reaction was compared with that of [RuCl3(2mqn)NO] with 8-quinolinol (Hqn) or 5-chloro-8-quinolinol (H5cqn). Photoirradiation reaction of complex 1 at room temperature in deaerated CH2Cl2 in the presence of NO gave trans-[RuCl(2cqn)(2mqn)NO] (the Cl is trans to the NO) and complex 2 with recovery of complex 1. The reaction was contrasted with that of cis-1 [RuCl(qn)(2mqn)NO] or cis-1 [RuCl(5cqn)(2mqn)NO]. The crystal structure of complex 1 was determined by X-ray diffraction. The reactions were examined under consideration of atomic charge of the phenolato oxygen in 8-quinolinol and its derivatives calculated at the restricted Hartree-Fock/6-311G** level.  相似文献   

7.
A new dinuclear ruthenium(II) catecholato complex [Cp*Ru(κ262-1,2-O2C6H4)RuCp*] (3; Cp* = η5-C5Me5) has been prepared by the reaction of [Cp*RuCl]4 with 2 equiv. of disodium catecholate in THF. Complex 3 has a dinuclear structure, in which one of the Cp*Ru fragments is κ2-bonded to the two oxygen atoms and the other is η6-bonded to the aromatic ring. Similar treatment of [Cp*RuCl]4 with disodium 2,3-naphthalenediolate affords an analogous κ26-bonded dinuclear complex [Cp*Ru(κ262-2,3-O2C10H6)RuCp*] (4) with selective π-complexation at the oxygen-substituted naphthalene ring. The molecular structure of 4 has been determined by X-ray crystallography. The oxygen-bound ruthenium atoms in complexes 3 and 4 are coordinatively unsaturated and readily uptake 1 equiv. of carbon monoxide to give the corresponding carbonyl adducts [Cp*Ru(CO)(κ262-1,2-O2C6H4)RuCp*] (5) and [Cp*Ru(CO)(κ262-2,3-O2C10H6)RuCp*] (6), respectively.  相似文献   

8.
Three new coordination complexes [Mn(L)(H2O)2](1,4-BDC)·2H2O (1), [Mn(L)0.5(1,4-BDC)]CH3OH·H2O (2) and [Mn(L)(H2O)2](1,2-HBDC)2·2H2O (3) were synthesized by solvothermal reactions of 1,2,4,5-tetrakis(imidazol-1-ylmethyl)benzene (L) and 1,4-benzenedicarboxylic acid (1,4-H2BDC) or 1,2-benzenedicarboxylic acid (1,2-H2BDC) with Mn(II) salt, and characterized by single crystal X-ray diffraction, IR, thermogravimetric and elemental analyses. In complexes 1 and 3, each ligand L links four Mn(II) atoms to form two-dimensional (2D) cationic network with non-coordinated 1,4-BDC2− and 1,2-HBDC anions lying in the voids between the two adjacent layers, respectively. The 2D layers are further connected together by hydrogen bonds to give three-dimensional (3D) supramolecular structures. However, the 1,4-BDC2− in 2 acts not only as counteranion, but also as bridging ligand leading to the formation of 2-fold interpenetrated 3D framework with pcu (primitive cubic unit) topology. The Mn(II) atoms bridged by carboxylate groups in 2 show antiferromagnetic interactions.  相似文献   

9.
Reaction between the dinuclear model hydrolases [M2(μ-OAc)2(OAc)2(μ-H2O)(tmen)2]; M = Ni (1); M = Co (2) and trimethylsilyltrifluoromethanesulphonate (TMS-OTf) under identical reaction conditions gives the mononuclear complex [Ni(OAc)(H2O)2(tmen)][OTf] · H2O (3) in the case of nickel and the dinuclear complex [Co2(μ-OAc)2(μ-H2O)2(tmen)2][OTf]2 (4) in the case of cobalt.Reaction of (3) with urea gives the previously reported [Ni(OAc)(urea)2(tmen)][OTf] (5), whereas (4) gives [Co2(OAc)3(urea)(tmen)2][OTf] (6) previously obtained by direct reaction of (2) with urea. Both (3) and (4) react with monohydroxamic acids (RHA) to give the dihydroxamate bridged dinuclear complexes [M2(μ-OAc)(μ-RA)2(tmen)2][OTf]; M = Ni (7); M = Co (8) previously obtained by the reaction of (1) and (2) with RHA, illustrating the greater ability of hydroxamic acids to stabilize dinuclear complexes over that of urea by means of their bridging mode, and offering a possible explanation for the inhibiting effect of hydroxamic acids by means of their displacing bridging urea in a possible intermediate invoked in the action of urease.  相似文献   

10.
Ten transition metal coordination complexes [Cu2(phen)(p-tpha)(μ-O)]n1, [Cu(m-tpha)(imH)2]n2, [Ni(5-Haipa)2(H2O)2]n3, [Ni(phen)2(H2O)2]·btc·[Ni(H2O)6]0.5·9H2O 4, [Co(2,5-pdc)(H2O)2]n·nH2O 5, [Co2(2,5-pdc)2(H2O)6]n·2nH2O 6, [Fe(2,5-Hpdc)2(H2O)2]·H2O 7, [Co(C6H4NO2)3]·H2O 8, [Fe22-btec)(μ2-H2btec)(bipy)2(H2O)2]n9, [Mn(phen)(2,5-pdc)(H2O)2]·H2O 10 (H4btec = 1,2,4,5-benzenetetracarboxylic acid, phen = 1,10-phenanthroline, 2,5-H2pdc = 2,5-pyridine-dicarboxylic acid, p-tpha = p-phthalic acid, m-tpha = m-phthalic acid, bipy = 2,2′-bipyridine, 5-H2aipa = 5-aminoisophthalic acid, imH = imidazole, H3btc = 1,3,5-benzenetricarboxylic acid) were synthesized through hydrothermal method. They were characterized by UV-Vis absorption spectra, single-crystal X-ray diffraction and surface photovoltage spectra (SPS). Structural analysis indicated that the complexes 1, 2, 3, 5, 6 and 9 were linked into infinite structures bridged by organic acid ligands. The other four complexes were molecular complexes and further connected to 2D or 3D structures by the hydrogen bonds. The SPS of complexes 1-10 indicate that there are positive response bands in the range of 300-800 nm showing different levels of photo-electric conversion properties. The intensity, position, shape and the number of the response bands in SPS are obviously different since the structure, species, valence, dn electrons configuration and coordinated environment of the center metals are different. There are good relationships between SPS and UV-Vis spectra.  相似文献   

11.
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   

12.
Cyclometalation of benzo[h]quinoline (bzqH) by [RuCl(μ-Cl)(η6-C6H6)]2 in acetonitrile occurs in a similar way to that of 2-phenylpyridine (phpyH) to afford [Ru(bzq)(MeCN)4]PF6 (3) in 52% yield. The properties of 3 containing ‘non-flexible’ benzo[h]quinoline were compared with the corresponding [Ru(phpy)(MeCN)4]PF6 (1) complex with ‘flexible’ 2-phenylpyridine. The [Ru(phpy)(MeCN)4]PF6 complex is known to react in MeCN solvent with ‘non-flexible’ diimine 1,10-phenanthroline to form [Ru(phpy)(phen)(MeCN)2]PF6, being unreactive toward ‘flexible’ 2,2′-bipyridine under the same conditions. In contrast, complex 3 reacts both with phen and bpy in MeCN to form [Ru(bzq)(LL)(MeCN)2]PF6 {LL = bpy (4) and phen (5)}. Similar reaction of 3 in methanol results in the substitution of all four MeCN ligands to form [Ru(bzq)(LL)2]PF6 {LL = bpy (6) and phen (7)}. Photosolvolysis of 4 and 5 in MeOH occurs similarly to afford [Ru(bzq)(LL)(MeCN)(MeOH)]PF6 as a major product. This contrasts with the behavior of [Ru(phpy)(LL)(MeCN)2]PF6, which lose one and two MeCN ligands for LL = bpy and phen, respectively. The results reported demonstrate a profound sensitivity of properties of octahedral compounds to the flexibility of cyclometalated ligand. Analogous to the 2-phenylpyridine counterparts, compounds 4-7 are involved in the electron exchange with reduced active site of glucose oxidase from Aspergillus niger. Structure of complexes 4 and 6 was confirmed by X-ray crystallography.  相似文献   

13.
The preparation and structural characterization of several new Ru(II) complexes in which four coordination positions are occupied by the sulfur atoms of a macrocycle, either 1,4,7,10-tetrathiacyclododecane ([12]aneS4) or 1,5,9,13-tetrathiacyclohexadecane ([16]aneS4), and the two others by relatively labile ligands (Cl, , H2O, dmso-S), are described:cis-[Ru([12]aneS4)(dmso-S)(H2O)](CF3SO3)2 (2a), cis-[Ru([12]aneS4)(dmso-S)(ONO2)](NO3) (2b), cis-[Ru([16]aneS4)Cl2] (4), and trans-[Ru([16]aneS4)(dmso-S)(H2O)](CF3SO3)2 (5).The complexes of the larger [16]aneS4 macrocycle have a flexible coordination geometry, either cis or trans, that makes them unsuited for being used as precursors in metal-driven self-assembly processes.On the contrary, the [12]aneS4 complexes cis-[Ru([12]aneS4)(dmso-S)Cl]Cl (1) and, above all, its chlorido free derivatives cis-[Ru([12]aneS4)(dmso-S)(H2O)](CF3SO3)2 (2a) and cis-[Ru([12]aneS4)(dmso-S)(ONO2)](NO3) (2b) are potential precursors of the geometrically stable 90° bis-acceptor fragment cis-[Ru([12]aneS4)]2+.Preliminary results of their reactivity towards the linear linker pyrazine (pyz) showed that the nature of the isolated product depends on that of the counter-anion.When treated with pyz 2b afforded the dinuclear complex [{Ru([12]aneS4)(ONO2)}2(μ-pyz)](NO3)2 (8), while 2a gave the molecular triangle [{cis-Ru([12]aneS4)(μ-pyz)}3](CF3SO3)6 (9), both in low yields.The X-ray structures of compounds 2a, 2b, 4, 5, [{Ru([12]aneS4)Cl}2(μ-pyz)]Cl2 (7), 9, and of the sandwich complex[Ru([12]aneS3-S)2](CF3SO3)2 (3), in which only three sulfur atoms of each macrocycle are bound to ruthenium, are also described.  相似文献   

14.
Reaction of cis-Ru(bisox)2Cl2, where bisox is 4,4,4′,4′-tetramethyl-2,2′-bisoxazoline, with excess of pyridine-2-carboxaldehyde (py-2-al) in 1:1 (v/v) methanol-water mixture under nitrogen atmosphere and subsequent addition of excess of NH4PF6 give [Ru(bisox)2(py-2-al)](PF6)2 · H2O (1). Refluxing of 1 in dehydrated methanol in presence of triethylamine yields the corresponding hemiacetalate complex: [Ru(bisox)2 (pyridine-2-(α-methoxymethanolato))]PF6 · 1.5H2O (2). Both the complexes have been characterised by single crystal X-ray crystallography, FTIR and NMR. In cyclic voltammetry in acetonitrile at a glassy carbon electrode, 2 displays a quasireversible Ru(II/III) couple at 1.08 V versus NHE which is not observed in 1. A tentative mechanism is proposed for the conversion of 1 to 2. DFT calculations with the LanL2DZ basis set have been performed to investigate these observations theoretically.  相似文献   

15.
By the reactions of a new ligand, 2,6-di(N,N′-5-ethyl-1,3,4-thia-diazole-2-formamide)pyridine (H2L), with Mn(II) and Cu(II) ions, two complexes namely [Mn33-O)(H2L)(L)2]·2DMF (1) and [Cu2(μ-H2O)(L)2]·DMF (2) were obtained, respectively. Compound 1 is a trinuclear complex containing triangle frames formed by three Mn(II) ions with a bridged μ3-O in the center; ligand H2L acts as a penta-dentate fashion. Compound 2 is a dinuclear complex, in which H2L coordinates two Cu(II) centers as a tetra-dentate coordination mode. Magnetic investigations show that an antiferromagnetic coupling between metal ions exists in 1, and a ferromagnetic coupling exists in 2. Thermal decomposition kinetic parameters of complex 1 are also obtained by employing non-isothermal (model-free) method.  相似文献   

16.
The bis(pyrazol-1-yl)azine ligands 2,3-bis(pyrazol-1-yl)quinoxaline (bpzqnx), 2,3-bis(pyrazol-1-yl)pyrazine (bpzprz) and 3,6-bis(3,5-dimethylpyrazol-1-yl)pyridazine (bpz*pdz) were prepared by the reaction of pyrazolate salts and the corresponding azine dichloride derivatives. The reaction of these ligands with Ru(arene) precursors led to the mononuclear complexes [RuCl(arene)(L)]BPh4 (arene = p-cymene, L = bpzqnx, 1, bpzprz, 5, bpz*pdz, 7; arene = C6H6, L = bpzqnx, 2, bpzprz, 6, bpz*pdz, 8) with the N-donor ligand coordinated in a bidentate chelate way. In general, the ligands coordinate through one pyrazole ring and the azine, except in the cases of 1 and 2 where the two pyrazolyl rings are coordinated to the metal in a symmetrical way. When the reactions between the ruthenium precursors and bpzqnx are carried out in MeOH, the complexes [RuCl(arene)(OMepzqnx)]BPh4 with partially methanolyzed ligands are isolated (arene = p-cymene, 3; C6H6, 4). In this process a methoxy group has replaced one of the pyrazole groups in the ligand. The X-ray structures of 6 and 7 have been determined. These compounds have a three-legged piano-stool structure with cations and anions packed through weak interactions. Complexes 1-8 are active in ketone hydrogenation transfer processes even in the absence of base.  相似文献   

17.
The interactions of π-arene-Ru(II)-chloroquine complexes with human serum albumin (HSA), apotransferrin and holotransferrin have been studied by circular dichroism (CD) and UV-Visible spectroscopies, together with isothermal titration calorimetry (ITC). The data for [Ru(η6-p-cymene)(CQ)(H2O)Cl]PF6 (1), [Ru(η6-benzene)(CQ)(H2O)Cl]PF6 (2), [Ru(η6-p-cymene)(CQ)(H2O)2][PF6]2 (3), [Ru(η6-p-cymene)(CQ)(en)][PF6]2 (4), [Ru(η6-p-cymene)(η6-CQDP)][BF4]2 (5) (CQ: chloroquine; DP: diphosphate; en: ethylenediamine), in comparison with CQDP and [Ru(η6-p-cymene)(en)Cl][PF6] (6) as controls demonstrate that 1, 2, 3, and 5, which contain exchangeable ligands, bind to HSA and to apotransferrin in a covalent manner. The interaction did not affect the α-helical content in apotransferrin but resulted in a loss of this type of structure in HSA. The binding was reversed in both cases by a decrease in pH and in the case of the Ru-HSA adducts, also by addition of chelating agents. A weaker interaction between complexes 4 and 6 and HSA was measured by ITC but was not detectable spectroscopically. No interactions were observed for complexes 4 and 6 with apotransferrin or for CQDP with either protein. The combined results suggest that the arene-Ru(II)-chloroquine complexes, known to be active against resistant malaria and several lines of cancer cells, also display a good transport behavior that makes them good candidates for drug development.  相似文献   

18.
A full account of half-sandwich complexes of ruthenium(II) having three-legged “piano-stool” geometry supported by tridentate (2-pyridyl)alkylamine ligands is presented. Reaction of the dimer [{(η6-C6H6)RuCl(μ-Cl)}2] with N-methyl-N,N-bis(2-pyridylmethyl)amine (MeL) in CH3OH in the presence of NH4PF6 affords the complex [(η6-C6H6)Ru(MeL)][PF6]2 (1). A similar reaction with N-methyl-N,N-bis(2-pyridylethyl)amine (MeL∗∗), however, affords a non-organometallic Ru(III)-dimeric complex (5) (the composition of this complex has been established by physicochemical method). Nucleophilic addition reaction on 1 with NaBH4 leads to the isolation of a cyclohexadienyl complex [(η5-C6H7)Ru(MeL)][PF6] (3). The molecular structure of 1 · 2CH3CN, 3, and previously reported cyclohexadienyl complex [(η5-C6H7)Ru(MeL)][PF6] (4) [MeL = N-methyl-[(2-pyridyl)ethyl(2-pyridyl)-methyl]amine], obtained from the reaction between NaBH4 and previously reported “piano-stool” complex [(η6-C6H6)Ru(MeL)][PF6]2 (2), has been confirmed by X-ray crystallography. Solution-state structure of new complexes 1 and 3 has been elucidated by their 1H NMR spectra in CD3CN. The behavior of complex 3 has been investigated with the aid of two-dimensional 1H NMR spectroscopy, as well. An attempt has been made to provide a rationale for the effect of supporting tridentate N-donor ligand [MeL, MeL, and MeL∗∗], varying in the chelate ring-size on (i) the relative stability of half-sandwich η6-benzene Ru(II) complexes and (ii) the electrophilicity of Ru(II)-coordinated benzene ring on the nucleophilic addition reactions.  相似文献   

19.
Reactions of zinc(II) ion with racemic malic acid (C4H6O5 = H3mal) result in the isolation of four new zinc(II) malato complexes: (NH4)[Zn(R-H2mal)3] · H2O (1), trans-[Zn(R-H2mal)(S-H2mal)(H2O)2] · 2H2O (2), (NH4)2[Zn(R-Hmal)(S-Hmal)] · 2H2O (3), and [Zn2(R-Hmal)(S-Hmal)(H2O)4]n · 2nH2O (4). Three R-malic acids in 1 act as bidentate ligands via their alcoholic and the central carboxy groups with Zn(II) ion, leaving the terminal carboxylic acid groups free. The R- and S-malates of 2 coordinate in a bidentate manner with zinc ion in trans-form. In 3, Zn(II) ion is coordinated by R- and S-malates in a tridentate fashion via their alcoholic and two carboxy groups. Complex 4 forms a two-dimensional layered structure through the links of a new dimeric unit [Zn2(R-Hmal)(S-Hmal)(H2O)4] with one of the oxygen atoms from the terminal carboxy group of malate ligand. The coordination of malates depends on pH variation, on Zn:malate ratio, and also on temperature. Tridentate chelation of malate in 3 is found between pH 4.5-9.0. The soluble monomeric species 1-3 have been investigated using 13C NMR spectra by long-time acquisition. The solution NMR spectra indicate that zinc malate complexes dissociate in H2O (D2O). Obvious downfield shifts of the central carboxy carbon atoms in 1-3 are observed compared with those of free malate, which indicate that these zinc malate complexes dissociate in aqueous solution.  相似文献   

20.
Halide abstraction from the 18 electron Ru(II) complex RuCl(CO)2[2,6-(CH2PtBu2)2C6H3] (2) with AgPF6 results in the exclusive formation of the cationic complex {Ru(CO)2[2,6-(CH2PtBu2)2C6H3]}+PF6 (3). The molecular structures of 2 and 3 were determined by complete single-crystal diffraction studies. X-ray crystallographic analysis of 3 reveals that the “open” coordination site is occupied by an agostic interaction between the metal center and an sp3 C-H bond of a tert-butyl substituent. DFT gas phase calculations (B97-1/SDD) show the necessity of two sterically demanding tert-butyl substituents on one P donor atom for the agostic interaction to occur. The reaction of 3 with H2 results in the quantitative conversion to {Ru(H)(CO)2[2,6-(CH2PtBu2)2C6H4]}+PF6 (4) where the aromatic Cipso-H bond is η2-coordinated to the metal center. Treatment of the agostic complex 4 with Et3N results in the formation of the neutral complex Ru(H)(CO)2[2,6-(CH2PtBu2)2C6H3] (5). The mechanistic details of 3 + H2 → 4 were investigated by DFT calculations at the B97-1/SDB-cc-pVDZ//B97-1/SDD level of theory.  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