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 15 毫秒
1.
 DNA binding by trans-[(H2O)(Pyr)(NH3)4RuII]2+ (Pyr=py, 3-phpy, 4-phpy, 3-bnpy, 4-bnpy) is highly selective for G7 with K G=1.1×104 to 2.8×104, with the more hydrophobic Pyr ligands exhibiting slightly higher binding. A strong dependence on ionic strength indicates that ion-pairing with DNA occurs prior to binding. At μ=0.05, d[RuII-DNA]/dt=k[RuII][DNA], where k=0.17–0.21 M–1 s–1 with the various Pyr ligands. The air oxidation of [(py)(NH3)4RuII] n -DNA to [(py)(NH3)4RuIII] n -DNA at pH 6 occurs with a pseudo-first-order rate constant of k obs=5.6×10–4 s–1 at μ=0.1, T=25  °C. Strand cleavage of plasmid DNA appears to occur by both Fenton/Haber-Weiss chemistry and by base-catalyzed routes, some of which are independent of oxygen. Base-catalyzed cleavage is more efficient than O2 activation at neutral pH and involves the disproportionation of covalently bound RuIII and, in the presence of O2, Ru-facilitated autoxidation to 8-oxoguanine. Disproportionation of [py(NH3)4RuIII] n -DNA occurs according to the rate law: d[RuII–GDNA]/dt=k 0[RuIII–GDNA]+k 1[RuIII–GDNA][OH], where k 0=5.4×10–4 s–1 and k 1=8.8 M–1 s–1 at 25  °C, μ=0.1. The appearance of [(Gua)(py)(NH3)4RuIII] under argon, which occurs according to the rate law: d[RuIII–G]/dt=k 0[RuIII–GDNA]+k 1[OH][RuIII–GDNA] (k 0=5.74×10–5 s–1, k 1=1.93×10–2 M–1 s–1 at T=25  °C, μ=0.1), is consistent with lysis of the N-glycosidic bond by RuIV-induced general acid hydrolysis. In air, the ratio of [Ru-8-OG]/[Ru-G] and their net rates of appearance are 1.7 at pH 11, 25  °C. Small amounts of phosphate glycolate indicate a minor oxidative pathway involving C4′ of the sugar. In air, a dynamic steady-state system arises in which reduction of RuIV produces additional RuII. Received: 11 November 1998 / Accepted: 3 March 1999  相似文献   

2.
Pentaammineruthenium(III) complexes of deoxyinosine (dIno) and xanthosine (Xao) ([RuIII(NH3)5(L)], L?is?dIno, Xao) in basic solution were studied by UV?Cvis spectroscopy, liquid chromatography/electrospray ionization mass spectrometry, and high-performance liquid chromatography. Both RuIII complexes disproportionate to RuII and RuIV. Disproportionation followed the rate law d[RuII]/dt?=?(k o?+?k 1[OH?])[RuIII]. k o and k 1 of disproportionation at 25?°C were 2.1 (±0.1)?×?10?3?s?1 and 21.4?±?3.2?M?1 s?1, respectively, for [RuIII(NH3)5(dIno)], and 3.5 (±0.7)?×?10?4?s?1 and 59.7?±?3.6?M?1?s?1, respectively, for [RuIII(NH3)5(Xao)]. The [RuIII(NH3)5(Xao)] complex disproportionates at a faster rate than [RuIII(NH3)5(dIno)] owing to the stronger electron-withdrawing effect of exocyclic oxygen in Xao. The activation parameters ??H ? and ??S ? for k 1 of [RuIII(NH3)5(dIno)] were 80.2?±?15.2?kJ?mol?1 and 47.6?±?9.8?J?K?1 mol?1, respectively, indicating that the disproportionation of RuIII to RuII and RuIV is favored owing to the positive entropy of activation. The final products of both complexes in basic solution under Ar were compared with those under O2. Under both conditions [Ru(NH3)5(8-oxo-L)] was produced, but via different mechanisms. In both aerobic and anaerobic conditions, the deprotonation of highly positively polarized C8-H of Ru-L by OH? initiates a two-electron redox reaction. For the next step, we propose a one-step two-electron redox reaction between L and RuIV under anaerobic conditions, which differentiates from Clarke??s mechanism of two consecutive one-electron redox reactions between L, RuIII, and O2.  相似文献   

3.
4.
A spectroelectrochemical study of [Ru2X9]n, X=Cl, Br; n=1, 2, 3, 4 has been undertaken. Stable solutions of n=4, 2, 1 can be formed by electrolysis at low temperatures. Analysis of the Vis-NIR spectra of the complexes indicate that the RuII---RIII dimers (n=4) have delocalised mixed valence and that the RuIII---RIII (n=3) dimers have a strong Ru---Ru bond. The more oxidised materials do not form a Ru---Ru bond; the spectroscopic data indicates the RuIII---RIV dimers have localised valency.  相似文献   

5.
We describe the synthesis, characterization, and reactivity of several Ru(II) complexes of the type cis-L2Ru(Z)n+, where L is an α-diimine [e.g. 2,2′-bipyridine (bpy) or 1,10-phenanthroline (phen)] ligand and Z is a bis-coordinated scorpionate ligand such as tris-(1-pyrazolyl)methane (HC(pz)3, PZ=1-pyrazolyl; n=2) or tetrakis-(1-pyrazolyl)borate anion (B(pz)4; n=1). The complexes each exhibit strong visible absorption assigned as a π*(L)←dπ(Ru) metal-to-ligand charge-transfer (MLCT) transition characteristic of the cis-L2Ru2+ kernel. A corresponding MLCT excited state emission is observed in room temperature CH3CN solution, although emission energies, lifetimes, and quantum yields are reduced relative to Ru(bpy)3 2+. Electronic spectra and cyclic voltammetry measurements indicate that the relative π-acceptor abilities of the coordinated Z are: Z=(1H-pyrazolyl)2(pz)2B(pz)2<(pyridine)2<(pz)2CH(pz). Uncoordinated pz groups of cis-(bpy)2Ru(pz)2B(pz)2 + can be reacted to form a sterically hindered, localized-valence (Kcom33 l mol−1) cis,cis-(bpy)2RuII(pz)2B(pz)2RuII(bpy)2 3+ dimer. The dimer properties are interpreted by comparison to the known cis-(bpy)2RuII(pz)2RuII(bpy)2 2+ analog. The dimer is photoreactive and undergoes an asymmetrical photocleavage in CH3CN (yielding cis-(bpy)2RuIII(pz)2B(pz)2 2+ and cis-(bpy)2RuII(CH3CN)2 2+), similar to the corresponding thermal reaction observed for the mixed-valence cis-(bpy)2RuII(pz)2RuIII(bpy)2 3+ system.  相似文献   

6.
Treatment of [Ru3(CO)9{P(C4H3S)3}(μ-dppm)] (1) [dppm = bis(diphenylphosphino)methane] with molecular oxygen in benzene at 60 °C affords oxo-capped [Ru3(CO)63-CO){P(C4H3S)3}(μ-dppm)(μ3-O)] (2), while with elemental sulfur and selenium related chalcogenide-capped clusters [Ru3(CO)63-CO){P(C4H3S)3}(μ-dppm)(μ3-E)] (3, E = S; 5, E = Se) and bis(chalcogenide) clusters [Ru3(CO)6{P(C4H3S)3}(μ-dppm)(μ3-E)2] (4, E = S; 6, E = Se) result. Reaction of 1 with H2S in refluxing THF affords the previously reported [(μ-H)2Ru3(CO)7(μ-dppm)(μ3-S)] (7) together with the new sulfido-capped dihydride [(μ-H)2Ru3(CO)6{P(C4H3S)3}(μ-dppm)(μ3-S)] (8). All new compounds have been characterized by spectroscopic data, and 2 and 8 by single-crystal X-ray diffraction analyses. Oxo-capped 2 consists of a triangular ruthenium framework capped on opposite sides by oxo and carbonyl groups, while 8 consists of a ruthenium triangle by a capping sulfido ligand and two inequivalent bridging hydride ligands.  相似文献   

7.
Mononuclear 5-(4-pyridyl)-10,15,20-triphenylporphyrin and 5-(3-pyridyl)-10,15,20-triphenylporphyrin as well as tetranuclear 5,10,15,20-tetra(4-pyridyl)porphyrin (tetra-4-pp) and 5,10,15,20-tetra(3-pyridyl)porphyrin) (tetra-3-pp) arene ruthenium(II) derivatives (arene is C6H5Me or p-Pr i C6H4Me) were prepared and evaluated as potential dual photosensitizers and chemotherapeutics in human Me300 melanoma cells. In the absence of light, all tetranuclear complexes were cytotoxic (IC50 ≤ 20 μM), while the mononuclear derivatives were not (IC50 ≥ 100 μM). Kinetic studies of tritiated thymidine and tritiated leucine incorporations in cells exposed to a low concentration (5 μM) of tetranuclear p-cymene derivatives demonstrated a rapid inhibition of DNA synthesis, while protein synthesis was inhibited only later, suggesting arene ruthenium–DNA interactions as the initial cytotoxic process. All complexes exhibited phototoxicities toward melanoma cells when exposed to laser light of 652 nm. At low concentration (5 μM), LD50 of the mononuclear derivatives was between 5 and 10 J/cm2, while for the tetranuclear derivatives LD50 was approximately 2.5 J/cm2 for the [Ru46-arene)4(tetra-4-pp)Cl8] complexes and less than 0.5 J/cm2 for the [Ru46-arene)4(tetra-3-pp)Cl8] complexes. Examination of cells under a fluorescence microscope revealed the [Ru46-arene)4(tetra-4-pp)Cl8] complexes as cytoplasmic aggregates, whereas the [Ru46-arene)4(tetra-3-pp)Cl8] complexes were homogenously dispersed in the cytoplasm. Thus, these complexes present a dual synergistic effect with good properties of both the arene ruthenium chemotherapeutics and the porphyrin photosensitizer.  相似文献   

8.
 The synthesis, spectroscopic, and electrochemical properties of trans-[L(Pyr)(NH3)4RuII/III] (Pyr=py, 3-phpy, 4-phpy, 3-bnpy, or 4-bnpy; L=H2O, Guo, dGuo, 1MeGuo, Gua, Ino, or G7-DNA) are reported. As expected, the Pyr ligand slows DNA binding by trans-[(H2O)(Pyr)(NH3)4RuII]2+ relative to [(H2O)(NH3)5RuII]2+ and favors reduction of RuIII by about 150 mV. The pyridine ligand also promotes the disproportionation of RuIII to afford the corresponding complexes of RuII and, presumably, RuIV. For L=Ino, disproportionation follows the rate law: d[RuII]/dt=k 0[RuIII]+k 1[OH][RuIII], k 0=(2.7±0.7)×10–4 s–1 and k 1=70±1 M–1 s–1. Received: 11 May 1998 / Accepted: 3 March 1999  相似文献   

9.
1 Cereals can be attacked severely by the grain aphid, Sitobion avenae (F.), and the bird cherry‐oat aphid, Rhopalosiphum padi (L.). The time of migration from winter hosts in spring is important regarding input to decision support systems concerning insecticide treatment of aphids. 2 The present study aimed to construct two separate migration models, which could be used immediately for advisors and farmers but also be part of a decision support system for the chemical control of aphids in winter wheat and spring barley. 3 Winter wheat (Triticum sativum Lam.) and spring barley (Hordeum vulgare L.) fields were monitored from 1991–2005 for the occurrence of grain aphids and bird cherry‐oat aphids, and the data were used to construct migration models. 4 The models were constructed based on all 9 years data and subsequently validated by using all 9 years data, excluding 1 year at a time. 5 The migration model for the grain aphid producing the best forecast was obtained with migration date M (number of days from 1 June), degree‐days (DD) of April (Ta) and DD of May (Tm), with the model being M= 265.0 ? 10.2 loge(Ta) ? 35.1 loge(Tm). 6 The migration model for the bird cherry‐oat aphid producing the best forecast was obtained with migration date M (number of days from 15 May), DD of April (Ta) and DD from 1–15 May (Tm1), with the model being M= 294.4 ? 34.7 loge(Ta) ? 22.5 loge(Tm1). 7 The models only worked well in areas where the grain aphid and the bird cherry‐oat aphid are entirely holocyclic.  相似文献   

10.
The reactions of the butterfly complex Ru4(CO)12(MeC2Ph) with several alkynes give the quasiplanar derivatives Ru4(CO)11(MeC2Ph)(Alkyne) in almost quantitative yields.The structure of Ru4(CO)11(MeC2Ph)2 has been determined by X-ray methods. Crystals are monoclinic, space group C2/c, with Z = 4 in a unit cell of dimensions a 22.383(16), b 9.048(8), c 18.268(12) Å, β = 127.25(4)°. The structure has been solved from diffractometer data by Patterson and Fourier methods and refined by full-matrix least-squares to R = 0.034 for 1420 observed reflections. The complex, having an imposed C2 symmetry, presents a tetranuclear metal cluster in which the Ru atoms are in a tetrahedrally-distorted square arrangement. Ten carbonyls are terminal and one symmetrically bridges an edge of the cluster. Each of the two alkyne ligands is σ-bonded to two Ru atoms on the opposite vertices of the cluster and π-bonded to the other two. The organometallic cluster has a Ru4C4 core in which the metal and carbon atoms occupy the vertices of a triangulated dodecahedron.  相似文献   

11.
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   

12.
《Inorganica chimica acta》1986,120(2):135-143
As known, on hydrated alumina support the Ru3(CO)12 cluster quickly decomposes into monometallic subcarbonyls. By FT-IR spectroscopy combined with data handling procedures, the structure and thermal behaviour of the bimetallic systems of Fe2Ru(CO)12/ Al2O3 and H2FeRu3(CO)13 together with that of Ru3(CO)12 have been studied. At the end of an interaction with the hydrated alumina surface, iron ruthenium bimetallic clusters decompose into identical ruthenium anchored surface species RuA  RuIII(CO)2, RuBRuII(CO)2 and RuCRu0(CO)2, like pure ruthenium clusters, and no CO bonded to iron has been detected RuB and RuC are stable in a wide temperature range (300–500 K) and they can be interconverted by oxidation and reduction. RuA is less stable (300–400 K). These main molecule-like species, anchored onto uniform sites of the surface, are accompanied by mobile subcarbonyls and stable monocarbonylic species, which occupy a large variety of different sites.  相似文献   

13.
The reaction of [RuIII(edta)(H2O)] (edta4− = ethylenediaminetetraacetate) and [RuIII(hedtra)(H2O)] (hedtra3− = N-hydroxyethylethylenediaminetriacetate) with various purine based 5′-nucleotides (Nu) viz. adenosin-5′-monophosphate (AMP), guanosin-5′-monophosphate (GMP), inosin-5′-monophosphate (IMP) was studied kinetically as a function of [Nu] at various temperatures (15-35 °C) at a fixed pH (4.5). Kinetic results suggest that the binding of 5′-nucleotides takes place in a rapid [Nu] dependent rate-determining step. Kinetic data and activation parameters are accounted for the operation of an associative mechanism. The antitumor activities of both [RuIII(edta)(H2O)] (1) and [RuIII(hedtra)(H2O] (2) have been evaluated using MCF-7 (breast cancer), NCI-H460 (lung cancer) and SF-268 (CNS) cell lines.  相似文献   

14.
Abstract

The noncovalent interactions of phytate (Phy6-) with biogenic amines were studied potentiometrically in aqueous solution, at t= 25°C. Several species are formed in the different H+-Phy6--amine (A) systems, which have the general formula Ap(Phy)Hq(12-q)-, with p ≤ 3 and 6 ≤ q ≤ 10. The stability of these species is strictly dependent on the charges involved in the formation equilibria. For the equilibrium pHiAi+ + Hj(Phy)(12-j)- = Ap(Phy)Hq(12q)-, (q = pi + j)we found the relationship logK= aζ (ζ is the charge product of reactants), where a= 0.35(0.03, valid for all the amines; this roughly indicates an average free energy contribution per bond -ΔG0 = 4.0 ± 0.2 kJ mol-1. A slightly more sophisticated equation is also proposed for predicting the stability of these species. Owing to the quite high (partially protonated) phytate charge, the stability of Ap(Phy)Hq(12-q)- species is quite high, making phytate a strong amine sequestering agent in a wide pH range.  相似文献   

15.
Analyses of how organisms are likely to respond to a changing climate have focused largely on the direct effects of warming temperatures, though changes in other variables may also be important, particularly the amount and timing of precipitation. Here, we develop a network of eight growth‐increment width chronologies for freshwater mussel species in the Pacific Northwest, United States and integrate them with tree‐ring data to evaluate how terrestrial and aquatic indicators respond to hydroclimatic variability, including river discharge and precipitation. Annual discharge averaged across water years (October 1–September 30) was highly synchronous among river systems and imparted a coherent pattern among mussel chronologies. The leading principal component of the five longest mussel chronologies (1982–2003; PC1mussel) accounted for 47% of the dataset variability and negatively correlated with the leading principal component of river discharge (PC1discharge; r = ?0.88; P < 0.0001). PC1mussel and PC1discharge were closely linked to regional wintertime precipitation patterns across the Pacific Northwest, the season in which the vast majority of annual precipitation arrives. Mussel growth was also indirectly related to tree radial growth, though the nature of the relationships varied across the landscape. Negative correlations occurred in forests where tree growth tends to be limited by drought while positive correlations occurred in forests where tree growth tends to be limited by deep or lingering snowpack. Overall, this diverse assemblage of chronologies illustrates the importance of winter precipitation to terrestrial and freshwater ecosystems and suggests that a complexity of climate responses must be considered when estimating the biological impacts of climate variability and change.  相似文献   

16.
The electrochromic properties of two new mixed valence ruthenium complexes: K[(NC5H4CH2PO3H2)RuIII(NH3)4(NC)RuII(CN)5] and K[(NC5H4PO3H2)RuIII(NH3)4(NC)RuII(CN)5], where phosphonic acid groups have been introduced at the pyridine ligand, have been studied in homogeneous solution and adsorbed on transparent nanocrystalline SnO2 electrodes. These species exhibit a superior stability with respect to the previously studied, K[(NC5H4CO2H)RuIII(NH3)4NCRuII(CN)5] complex, showing negligible optical density changes after cycling 20 000 times the electrodes between −0.5 and 0.5 V versus SCE.  相似文献   

17.
Reactions between Ru3(CO)12 and 1,8-bis(diphenylphosphino)naphthalene (dppn) have given the four complexes Ru3(μ-H){μ3-PPh2(nap)PPh(C6H4)}(CO)8 (1), Ru4(μ-H){μ3-PPh2(nap)PPh(C6H4)}(μ-CO)3(CO)7 (2) and Ru4(μ-H)(μ3-C6H4){μ-PPh(nap)PPh2}(CO)11 (3) (in refluxing thf), and Ru44-P(nap)PPh2}(μ4-C6H4)(μ-CO)(CO)9 (4) (in refluxing toluene) which have been characterised by single crystal X-ray studies. They have been formed by aryl C-H and aryl C-P bond cleavage reactions, presumably from an initial (unobserved) chelate dppn complex. The unchanged chelating ligand is found in Ru3(μ-dppm)(CO)8(dppn) (5), obtained from Ru3(μ-dppm)(CO)10 and dppn in refluxing thf.  相似文献   

18.
The structural and energetic perturbations caused by cavity-creating mutations (Leu-41 → Val and Leu-41 → Ala) in the dimeric 4-α-helical-bundle protein ROP have been characterized by CD spectroscopy and differential scanning calorimetry (DSC). Deconvolution of the CD spectra showed a decrease in α -helicity as a result of the amino acid exchanges that follows qualitatively the overall decrease in conformational stability. Transition enthalpies are sensitive probes of the energetic change associated with point mutations. ΔH0 values at the respective transition temperatures, T1/2 (71.0, 65.3, and 52.9°C at 0.5 mg/ml) decrease from 580 ± 20 to 461 ± 20 kJ/(mol of dimmer) and 335 ± 20 kJ/(mol of dimmer) for wildtype ROP (Steif, C., Weber, P., Hinz, H.-J., Flossdorf, J., Cesareni, G., Kokkinidis, M. Biochemistry 32:3867-3876, 1993), L41V, and L41A, respectively. The conformational stabilities at 25°C expressed by the standard Gibbs energies of denaturation, ΔG, are 71.7, 61.1, and 46.1 kJ/(mol of dimmer). The corresponding transition enthalpies have been obtained from extrapolation using the c(T)and c(T) functions. Their values at 25°C are 176.3, 101.9, and 141.7 kJ/(mol of dimmer) for wild-type ROP, L41V, and L41A, respectively. When the stability perturbation resulting from the cavity creating mutations is referred to the exchange of 1 mol of CH2 group, the average ΔΔG value is ?5.0 ± 1 kJ/(mol of CH2 group). This decrease in conformation stability suggests that dimeric ROP exhibits the same susceptibility to Leu → Yal and Leu → Ala exchanges as small monomeric proteins. Careful determinations of the partial specific heat capacities of wild-type and mutated protein solutions suggest that the mutational effects are predominantly manifested in the native rather than the unfolded state. © 1995 Wiley-Liss, Inc.  相似文献   

19.
  • 1 Shell growth in the freshwater pearl mussel, Margaritifera margaritifera, was investigated. Three non‐linear growth models (i.e. power, logistic and von Bertalanffy) were fitted to Scottish length‐at‐age data sets and compared.
  • 2 Overall, the von Bertalanffy model outperformed the other two approaches, generating the smallest residuals in eight out of 11 samples (the logistic model provided slightly better fits to the other three). It was concluded that individual M. margaritifera appear to grow in an approximately asymptotic fashion and that the von Bertalanffy equation is an appropriate growth model to fit to freshwater pearl mussel length‐at‐age data.
  • 3 The ranges in von Bertalanffy parameter estimates observed (k = 0.023–0.075 year‐1, L = 77–158 mm, to = ‐3.93–4.33 years) are typical of those reported in northern European populations.
  • 4 Most of the populations investigated had relatively low k‐values and high maximum age (Amax) estimates. This feature, which suggests high long‐term productivity and less vulnerability to decline (i.e. larger, longer‐living mussels produce more offspring), may be a reason why these populations have survived until now. The population which appears to be the most vulnerable (i.e. which has the highest k and lowest Amax) is probably not recruiting adequately at present.
  • 5 An index of absolute growth (mean shell length‐at‐age) was also used for comparing different populations. Observed between‐ and within‐river differences in mussel growth patterns may be associated with a number of environmental factors, particularly water temperature and productivity.
  • 6 A significant positive relationship between river length and mean mussel length‐at‐age was observed. In general, mussels grow large in large, cold rivers and vice versa, although there are exceptions which suggest that additional factors may be involved.
  相似文献   

20.
《Inorganica chimica acta》1986,120(2):145-152
Initial steps of the interaction of Ru3(CO)12, Fe3(CO)12, Fe2Ru(CO)12 and H2FeRu3(CO)13 clusters with hydrated alumina surfaces have been studied by FT-IR spectroscopy combined with data handling procedures. The first stage of the interaction is a pure physisorption. At the second stage the metal-metal bonds split producing a large variety of mobile subcarbonyls. In the case of Fe3(CO)12 the subcarbonyls form molecular Fe(CO)5, while at the bimetallic clusters they form molecular Fe- (CO)5 and Ru3(CO)12. Fe(CO)5 loses CO ligands producing Fe2+ and Fe3+ anchored ions. Ru3(CO)12, through further intermediate subcarbonyls, slowly decomposes into incipient anchored species RuO- (CO)2, RuII(CO)2 and RuIII(CO)2.  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