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 218 毫秒
1.
Single crystal X-ray structural characterizations are recorded for a wide range of adducts of the form MX:dppx (1:1)(n), M = silver(I) (predominantly), copper(I), X = simple (pseudo-) halide or oxy-anion (the latter spanning, where accessible, perchlorate, nitrate, carboxylate - a range of increasing basicity), dppx=bis(diphenylphosphino)alkane, Ph2P(CH2)xPPh2, x = 3-6. Adducts are defined of two binuclear forms: (i) [LM(μ-X)2L], with each ligand chelating a single metal atom, and (ii) [M(μ-X)2(μ-(P-L-P′))2M′] where both ligands L and halides bridge the two metal atoms; a few adducts are defined as polymers, the ligands connecting M(μ-X)2M′ kernels, this motif persisting in all forms. Synthetic procedures for all adducts have been reported. All compounds have been characterized both in solution (1H, 13C, 31P NMR, ESI MS) and in the solid state (IR).  相似文献   

2.
The reaction between nickel(II) nitrate and potassium phosphorus-1,1-dithiolates (di-sec-butyl and di-iso-butyl) in methanol yields 2:1 complexes which were characterized by FT-IR and NMR spectroscopy. 2:1 pyrazole adducts of both compounds were also obtained.The X-ray diffraction analysis of the compounds reveals square planar, four-coordination geometry for the homoleptic compounds and a six-coordinated distorted octahedral geometry for the adducts. In Ni[S2P(OBus)2]2 the molecules are associated through C-H?O hydrogen bonds (2.652 Å), and in Ni[S2P(OBui)2]2 the molecules are associated through C-H?S hydrogen bonds (2.948 Å). The pyrazole adducts are associated through N-H?O bonds and N-H?S bonds from the pyrazole nitrogen atoms, to form supramolecular assemblies. Thus, Ni[S2P(OBus)2(Pz)2]2 (Pz = pyrazole) forms bi-dimensional layers through N-H?O and N-H?S bonds (2.502 and 2.965 Å, respectively), whereas Ni[S2P(OBui)2(Pz)2]2 forms linear chains with N-H?S bonds 2.728 Å. The dithiophosphato groups behave as isobidentate chelating ligands.  相似文献   

3.
The uptake and degradation of nanomolar levels of [methyl-14C]choline in estuarine water samples and in seawater filtrate cultures composed mainly of natural free-living bacteria was studied. Uptake of [14C]choline exhibited Michaelis-Menten kinetics, with Kt + Sn values of 1.7 to 2.9 nM in filtrate cultures and 1.7 to 4.1 nM in estuarine-water samples. Vmax values ranged from 0.5 to 3.3 nM · h−1. The uptake system for choline in natural microbial assemblages therefore displays very high affinity and appears able to scavenge this compound at the concentrations expected in seawater. Uptake of choline was inhibited by some natural structural analogs and p-chloromercuribenzoate, indicating that the transporter may be multifunctional and may involve a thiol binding site. When 11 nM [14C]choline was added to water samples, a significant fraction (>50%) of the methyl carbon was respired to CO2 in incubations lasting 10 to 53 h. Cells taking up [14C]choline produced [14C]glycine betaine ([14C]GBT), and up to 80% of the radioactivity retained by cells was in the form of GBT, a well-known osmolyte. Alteration of the salinity in filtrate cultures affected the relative proportion of [14C]choline degraded or converted to [14C]GBT, without substantially affecting the total metabolism of choline. Increasing the salinity from 14 to 25 or 35 ppt caused more [14C]GBT to be produced from choline but less 14CO2 to be produced than in the controls. Lowering the salinity to 7 ppt decreased [14C]GBT production and increased 14CO2 production slightly. Intracellular accumulations of [14C]GBT in the salt-stressed cultures were osmotically significant (34 mM). Choline may be used as an energy substrate by estuarine bacteria and may also serve as a precursor of the osmoprotectant GBT, particularly as bacteria are mixed into higher-salinity waters.  相似文献   

4.
Pyrimido[1,2-a]-purin-10(3H)-one (M1G) is a secondary DNA damage product arising from primary reactive oxygen species (ROS) damage to membrane lipids or deoxyribose. The present study investigated conditions that might lead to artifactual formation or loss of M1G during DNA isolation. The addition of antioxidants, DNA isolation at low temperature or non-phenol extraction methods had no statistically significant effect on the number of M1G adducts measured in either control or positive control tissue samples. The number of M1G adducts in nuclear DNA isolated from brain, liver, kidney, pancreas, lung and heart of control male rats were 0.8, 1.1, 1.1, 1.1, 1.8 and 4.2 M1G/108 nt, respectively. In rat liver tissue, the mitochondrial DNA contained a 2-fold greater number of M1G adducts compared with nuclear DNA. Overall, the results from this study demonstrated that measuring M1G is a reliable way to assess oxidative DNA damage because the number of M1G adducts is significantly affected by the amount of ROS production, but not by DNA isolation procedures. In addition, this study confirmed that the background number of M1G adducts reported in genomic DNA could have been overestimated by one to three orders of magnitude in previous reports.  相似文献   

5.
32P-Postlabeling has emerged as a major tool for detecting DNA adducts resulting from exposure to complex carcinogen mixtures. An integral component of this assay is multi-directional PEI-cellulose TLC in which lipophilic 32P-adducts are resolved in high-salt, high-urea solvents following removal of the bulk of non-adduct radioactivity. This TLC system is very effective for adducts formed following exposure to individual carcinogens; however, adducts resulting from exposure to complex mixtures (e.g. cigarette smoke) generally appear in the form of the so-called diagonal radioactive zones. By using mixtures of polycyclic aromatic hydrocarbon- and aromatic amine-DNA adducts as well as adducts in mouse skin treated with cigarette smoke condensate, we have demonstrated that a combination of 0.3–0.4 M NH4 OH and isopropanol-4 M NH4OH (1-1.4:1) solvents can provide more sharply defined adduct spots than the commonly used urea solvents. The non-urea solvents also result in excellent resolution of many adducts which otherwise may remain buried in diagonal radioactive zones when using the urea solvents. In addition, the signal-to-noise ratio is increased 2- to 5-fold over the urea solvents enabling detection of discrete adducts at ≤3 adducts per 1010 nucleotides. These partition TLC solvents also involve fewer manipulations (e.g. no water washes to remove salt and urea), and are likely to be more informative with regards to the type of individual adducts detected in the biomonitoring of humans than has hitherto been possible.  相似文献   

6.
Silver carboxylates [Ag(O2CR): R=Me, tBu, 2,4,6-Me3C6H2], fluorocarboxlyates [Ag(O2CRf): Rf=C3F7, C6F13, C7F15] and their phosphine adducts [Ag(O2CR)·nPR3′: R=Me, tBu, 2,4,6-Me3C6H2, R′=Me, Ph, n=2; R=Me, R′=Me, n=3; Ag(O2CRf).2PPh3, Rf=C3F7, C6F13, C7F15] have been synthesised, characterised spectroscopically and used as precursors in the aerosol-assisted chemical vapour deposition of silver films. All the phosphine adducts produced films, though in general PMe3 adducts, proved more successful than PPh3 analogues. The fluoro-carboxylates and their PPh3 adducts all generated silver films, though the growth rate for the adducts was lower. All these latter films showed carbon impurities while fluorine was also evident in most cases. The X-ray structure of AgO2CC3F7·2PPh3 is also reported.  相似文献   

7.
《Inorganica chimica acta》1986,123(2):105-112
Adducts of adenosine (ado) with 3d metal perchlorates were synthesized by refluxing mixtures of ligand and salt in ethanol-triethyl orthoformate. Metal(III) perchlorates formed adducts involving 2:3 metal to ado molar ratio (MCr, Fe), i.e., M2(ado)3(ClO4)6·4H2O, whereas 1:1 adducts were produced by metal(II) perchlorates, as follows: M(ado)(ClO4)2·2H20 (MMn, Co, Ni, Cu); and M(ado)(Cl04)2 (MFe, Zn). All the new complexes seem to be polymeric, involving a linear chainlike backbone with single ado bridges between adjacent metal ions in most cases, i.e., −(-M-ado-)-n. The coordination sphere of each metal ion is completed by terminal aqua, ado and, with the exception of the Cu2+ complex, −OClO3 ligands, in the case of the hydrated new complexes. Ado would be binding through the N(1) and N(7) ring nitrogens, when functioning as bridging, bidentate. As regards the two water-free M(II) complexes (MFe, Zn), which are apparently distorted tetrahedral, the evidence available is interpreted in terms of the presence of tridentate bridging ado, binding through N(1), N(7) in the same fashion as above, and through one of the ribose hydroxyl oxygens, which form weaker bonds to M2+ ions located in a neighboring linear polymeric −(-M-ado-)-n unit; the coordination sphere in these complexes is completed by one −OClO3 ligand.  相似文献   

8.
Given the rise in drug-resistant Streptococcus pneumoniae, there is an urgent need to discover new antimicrobials targeting this pathogen and an equally urgent need to characterize new drug targets. A promising antibiotic target is dihydrodipicolinate synthase (DHDPS), which catalyzes the rate-limiting step in lysine biosynthesis. In this study, we firstly show by gene knock out studies that S. pneumoniae (sp) lacking the DHDPS gene is unable to grow unless supplemented with lysine-rich media. We subsequently set out to characterize the structure, function and stability of the enzyme drug target. Our studies show that sp-DHDPS is folded and active with a k cat = 22 s-1, K M PYR = 2.55 ± 0.05 mM and K M ASA = 0.044 ± 0.003 mM. Thermal denaturation experiments demonstrate sp-DHDPS exhibits an apparent melting temperature (T M app) of 72 °C, which is significantly greater than Escherichia coli DHDPS (Ec-DHDPS) (T M app = 59 °C). Sedimentation studies show that sp-DHDPS exists in a dimer-tetramer equilibrium with a K D 4→2 = 1.7 nM, which is considerably tighter than its E. coli ortholog (K D 4→2 = 76 nM). To further characterize the structure of the enzyme and probe its enhanced stability, we solved the high resolution (1.9 Å) crystal structure of sp-DHDPS (PDB ID 3VFL). The enzyme is tetrameric in the crystal state, consistent with biophysical measurements in solution. Although the sp-DHDPS and Ec-DHDPS active sites are almost identical, the tetramerization interface of the s. pneumoniae enzyme is significantly different in composition and has greater buried surface area (800 Å2) compared to its E. coli counterpart (500 Å2). This larger interface area is consistent with our solution studies demonstrating that sp-DHDPS is considerably more thermally and thermodynamically stable than Ec-DHDPS. Our study describe for the first time the knock-out phenotype, solution properties, stability and crystal structure of DHDPS from S. pneumoniae, a promising antimicrobial target.  相似文献   

9.
The value of ecosystems functions performed by forests in the climate change era has prompted increasing attention towards assessment of carbon stocks and fluxes in tropical forests. The aim of this study was to understand how forest management approaches and environmental controls impacted on soil CO2 efflux in a tropical Eastern Mau forest which is one of the blocks of the greater Mau complex in Kenya. Nested experimental design approach was employed where 32 plots were nested into four blocks (disturbed natural, undisturbed natural, plantation and glades). In 10 m2 plots, data were collected on soil CO2 efflux, soil temperature and soil moisture using soda lime methods, direct measurement and proxy techniques, respectively. There was significant forest management type effect (F3,127 = 3.01, p = 0.033) and seasonality effect (t test = 3.31, df = 1, p < 0.05) on mean soil CO2 efflux. The recorded mean soil CO2 efflux levels were as follows: plantation forest (9.219 ± 3.067 g C M?2 day?1), undisturbed natural forest (8.665 ± 4.818 g C M?2 day?1), glades (8.592 ± 3.253 g C M?2 day?1) and disturbed natural forest (7.198 ± 3.457 g C M?2 day?1). The study concludes that managing a forest in plantation form is primarily responsible for forest soil CO2 efflux levels due to aspects such as increased microbial activity and root respiration. However, further studies are required to understand the role and impact of soil CO2 efflux on the greater forest carbon budget.  相似文献   

10.
The rate theory for enzyme-triggered coagulation reactions, such as the clotting of fibrin or casein, is extended to the case of an arbitrary rate of production of the clotting species. It is shown that the general expression for the growth of the weight-average molecular weight of the clotting product, -Mw, is given by -Mw = M1{1 + ks {∫0tP(t)2 dt}/P(t)}, where M1 is the “monomer” molecular weight, ks the smoluchowskian flocculation rate constant and P(t) the total number of monomers produced by the enzyme in t. In the purely smoluchowskian case P(t) stands for the total number of monomers at the beginning of the clotting process. Numerical examples in which the rate of enzymic production is governed by complete Michaelis-Menten kinetics, are compared to cases in which this rate equals Vmax- It is shown that after exhaustion of the substrate the system continues to coagulate in a purely smoluchowskian way. Turbidimetric experiments on the clotting of micelles of whole and κ-casein are presented which suggest inactivation of the enzyme by non-productive binding in the flocs formed.  相似文献   

11.
Reactions of tBu3M (M = Al, Ga) with CO have been studied by density functional theory employing the B3PW91 functional. Calculations suggest that CO insertion into a M-C bond of tBu3M is thermodynamically favorable at room temperature, whereas CO coordination to tBu3M to form tBu3M·CO is unfavorable due to an unfavorable entropy change. These results are in agreement with experimental observations. Reaction of carbon monoxide with tBu3Ga at 50 °C and atmospheric pressure yields the dimeric tert-butylacyl complex [tBu2GaC(O)tBu]2 (1). Compound 1 has been characterized by X-ray crystallography and NMR and IR spectroscopy. Isotopic labeling with 13CO confirmed that the acyl carbon of 1 results from the CO starting material. CO insertion into a Ga-C bond does not occur for Me3Ga or nBu3Ga under a range of reaction conditions.  相似文献   

12.
Lung N-oxidase enzyme activity was about three times higher than liver N-oxidase at the pH optimum, about pH 8.9, whereas the activities were nearly the same at more physiological ranges of pH. The lung N-oxidase was also stimulated about 2-fold by 100 mM Mg2+ and by 0.1 mM Hg2+, whereas liver N-oxidase activity was inhibited by these concentrations of ions. The difference in response of liver and lung enzymes to Mg2+ and Hg2+ was not altered by preparing the microsomes in the presence of 50 mM ethylenediamine tetraacetic acid (EDTA) in 0.1 M Tris (hydroxymethyl) amino methane (Tris) buffer or 50 mM EDTA in 0.1 M KPO4 buffer, both at pH 7.6, indicating that the differences are probably not due to the presence of endogenous metals. The difference between the liver and lung N-oxidase systems may be due to the tissue environment rather than to the enzyme itself since mercury stimulation of lung N-oxidation began to disappear upon partial purification of the N-oxidase enzymes. In contrast to the effects of Hg2+ and Mg2+, 1 mM Ni2+ enhanced liver N-oxidase activity about 30% and 5 mM Ni2+ stimulated lung enzyme activity about 30% whereas concentrations above 10 mM were inhibitory to both N-oxidases. Both liver and lung demethylase activities were inhibited by these concentrations of Mg2+, Hg2+ and Ni2+.Various suifhydryl reagents were also tested for their effects on these enzymes. The mercurials, para-chloromercurybenzoate (pCMB) and phenylmercuryacetate (PMA) at concentrations of 0.1 mM had the same effect as HgCl2 inhibiting both demethylases and liver N-oxidase, but stimulating lung N-oxidase activity. However, 0.1 mM to 1 mMN-ethylmaleimide (NEM) and iodoacetamide had little if any effect on either liver or lung N-oxidase. It was also shown that Hg2+ effects on N-oxidase activity could be overcome by dilution.Changes in N,N-dimethyl aniline (DMA) metabolism with age were followed in rabbits from 4 days old to adult. There was a steady increase in lung demethylase activity and N-oxidase activity in the liver and lung to adult levels. However, the liver demethylase had a sharp increase in activity between 2 weeks and 1 month much like that seen with benzphetamine demethylase in rabbit liver.Activities of N-demethylase in liver and lung, and N-oxidr.se in liver from new-born rabbits were from 10 to 20 % of adult levels. However, in lung, N-oxidase activities in the newborn were about 50 % of adult levels. Microsomal N-oxidation in lungs from 2-day-old rabbits was stimulated by 0.1 mM mercury just as in the adult.  相似文献   

13.
The interaction of the radicals OH?, t-BuO?, eaq?, CO2XXX and O2XXX with the copper oxidase. laccase. from Polyporus, has been studied by the pulse-radiolysis technique. Each of these radicals formed transient adducts with a broad absorption maximum around 310 nm. Analysis of the optical properties and of the very fast rates of formation of these compounds shows that each radical interacts with a limited number of sites on the polypeplide part of the protein amongst R-S-S-R. histidine and aromatic residues. Interaction with the carbonyl group of some of the peptide bonds is also possible. The few target sites are probably hit simultaneously and electron transfer between these sites may also occur. In all cases, in a subsequent step, intramolecular electron transfer from the polypeptide radical adducts leads to a partial reduction of the blue type-1 Cu2+ with rates varying between 103 and 104 s?1. Further reduction of the type-1 Cu2+ occurs through a slow intermolecular reaction between two laccase radical transient adducts. In the case of COXXX2 and OXXX2, this slow reduction could alternatively be due to an intermolecular reaction between laccase and COXXX2 or OXXX2. The oxidant radicals OH?. BrXXX2 and (SCN)XXX2, which formed radical adducts with fully ascorbate-reduced laccase, did not induce any type-1 copper reoxidation.  相似文献   

14.
By spectrophotometry and 1H nmr, several of the stability constants of the thioether complexes between Mg2+, Ca2+, Mn2+, Cu2+, Zn2+, Cd2+, or Ag+ and d-biotin (Bio), tetrahydrothiophene (Tht), and dimethyl sulfide (Dms) have been measured in 50% aqueous ethanol, 96% N,N-dimethylformamide (DMF), 98% d6-dimethyl sulfoxide, or in D2O. With decreasing concentration of water, the thioether interaction increases with the biologically important metal ions, whereas, e.g., Ag+ behaves in the opposite way. The stability of these complexes is, in general, quite small: for example, with d-biotin in 96% DMF (I = 1.0; 25°C) log KM(Bio)M = 0.03 and 1.64 for Cu2+ and Ag+, respectively; in D2O (I = 0.5 for Ag+, all others 2–5; 27°C) log KM(Bio)M ? ?1.0, ?1.4, ?1.2, ?0.9, or 4.20 for Mg2+, Ca2+, Zn2+, Cd2+, or Ag+. In those cases where the difference log KM(Tht)M ? log KM(Bio)M can be calculated, it is in the order of 0.3 log units; this observation, as well as the chemical shifts measured, confirm the earlier suggestion that the interaction at the sulfur of biotin is stereoselective: the metal ion coordinates from “below” the tetrahydrothiophene ring of biotin to the sulfur atom, i.e., trans to the urea ring. It is emphasized that despite the low stability of these complexes with the biologically meaningful metal ions, the extent of the interaction is enough to create specific structures.  相似文献   

15.
Coupling of two diimidosulfinate units via a conjugated organic substituent to facilitate electronic communication between the two S(NR)2-moieties seems advantageous because polyimido sulfur compounds are exposed to SET processes. The coupling might facilitate electron transport. In this paper we present the syntheses and crystal structures of the heteroaromatic S-substituted metal diimidosulfinates [(thf)Li2{(H3CNC4H3)S(NtBu)2}2] (1), [(tmeda)Li{(SC8H5)S(NtBu)2}] (2), [Fe{(SC8H5)S(NtBu)2}2] (3), and [Cu{(SC8H5)S(NtBu)2}]2 (4), as potential starting materials. As the structural investigation shows that electronic communication between the S(NR)2-moiety and a S-bonded conjugated organic substituent is only feasible when the aromatic perimeter is arranged strictly in-plane to one S-N bond as observed in 1. Here the ring nitrogen atom is blocked against metal coordination. In the other complexes in-plane arrangement of the S-heteroaromatic substituent is precluded even by very weak metal coordination of the ring sulfur atom. The (N,N)-chelating lithium (2) and iron (3) coordination forces the ring in a more or less orthogonal arrangement relative to the SN2-plane. The ring-S?copper coordination in 4, which could have forced the heteroarene ring into the plane of one S-N bond, apparently is to weak to turn the ring about the S-C(sp2) bond.  相似文献   

16.
Deuterium-induced, 13C-isotope shifts are shown to vary considerably from the initially predicted values calculated for ordinary pyranose and furanose sugars, when minor structural changes are introduced into the carbohydrate ring. Both substitution of C-OH groups or reduction of C-OH to CH2 permitted the evaluation of γ effects of OD without the contribution of β-OD-induced shifting. The observed γ-shift values for these modified structures were twice as large as those previously noted. This difference is most probably due to favored salvation. Substitution of OH at C-6 led to the predicted loss of differential isotope-shift (d.i.s.) at C-6 because of its isolation from all β and γ OD groups. The 31P resonances of d-glucose 6-phosphate show downfield deuterium shifts. Based on d.i.s. values, new 13C-shift assignments are proposed for isomaltose and 2-amino-2-deoxy-α-d-glucose. A study of acidic carbohydrates has demonstrated that isotope shifts are somewhat larger for sp2-hybridized carbon atoms whose OH groups are acidic. Relaxation times for sp2 carbon atoms isolated from dipolar interaction with protons were very long in D2O relative to their relaxation time in the H2O environment.  相似文献   

17.
《Inorganica chimica acta》1986,116(2):109-117
Previously developed methods for the treatment of polyhedral boranes, carboranes, and metal clusters are extended to the treatment of gold clusters, which present a variety of new problems. In most cases gold atoms in such cluster compounds do not employ the usual 9-orbital sp3d5 spherical bonding orbital manifold. Instead almost all non-tetrahedral gold clusters consist of a center gold atom surrounded by a puckered polygonal belt of peripheral gold atoms generally with one or more additional peripheral gold atoms in distal positions above and/or below the belt. The peripheral gold atoms in such clusters use a 7-orbital spd5 cylindrical bonding orbital manifold, but their residual two orthogonal anti- bonding p orbitals can receive electron density from the filled d orbitals of adjacent peripheral gold atoms through dσ → pσ* and/or dπ → pπ* backbonding leading to bonding distances between adjacent peripheral gold atoms. Centered gold clusters can be classified into either spherical or toroidal clusters depending upon whether the center gold atom uses a 9-orbital sp3d5 spherical bonding orbital manifold or an 8-orbital sp2d5 toroidal bonding orbital manifold, respectively. The topology of the core bonding in gold clusters is generally not that of the Kn complete graph found in other clusters but instead mimics the topology of the polyhedron formed by the surface atoms. This apparently is a consequence of the poor lateral overlap of the cylindrical spd5 manifolds of the peripheral gold atoms. Examples of non-centered gold clusters treated in this paper include the squashed pentagonal bipyramidal Au7(PPh3)7+ and the edge-fused bitetrahedral (Ph3P)4Au6[Co(CO)4]2 which may be regarded as a ‘perauraethylene’ in which the six cluster gold atoms correspond to the six atoms of ethylene including a double bond between the two gold atoms corresponding to the two ethylene carbon atoms.  相似文献   

18.
Nickel and palladium paddlewheel complexes that feature 2-mercapto-1-t-butylimidazolyl (mimBut) bridging ligands, namely Ni2[mimBut]4 and Pd2[mimBut]4, have been synthesized and structurally characterized by X-ray diffraction. Since the mimBut ligand bridges in an asymmetric manner via a sulfur and nitrogen donor, paddlewheel compounds of the type M2[mimBut]4 may exist as isomers that are distinguished by the relative orientations of the ligands. In this regard, the (4,0)-Ni2[mimBut]4 and trans-(2,2)-Ni2[mimBut]4 isomers have been isolated for the nickel system, while the (4,0)-Pd2[mimBut]4 and (3,1)-Pd2[mimBut]4 isomers have been isolated for the palladium system.  相似文献   

19.
Cobalt(III) and rhodium(III) complexes of the series of [MIIICl3 − n(P)3 + n]n+ (M = Co or Rh; n = 0, 1, 2 or 3) have been prepared with the use of 1,1,1-tris(dimethylphosphinomethyl)ethane (tdmme) and mono- or didentate phosphines. The single-crystal X-ray analyses of both series of complexes revealed that the M-P and M-Cl bond lengths were dependent primarily on the strong trans influence of the phosphines, and secondarily on the steric congestion around the metal center resulting from the coordination of several phosphine groups. In fact, the M-P(tdmme) bonds became longer in the order of [MCl3(tdmme)] < [MCl2(tdmme)(PMe3)]+ < [MCl(tdmme)(dmpe)]2+ (dmpe = 1,2-bis(dimethylphosphino)ethane) < [M(tdmme)2]3+ for both CoIII and RhIII series of complexes, while the M-Cl bond lengths were shortened in this order (except for [M(tdmme)2]3+). Such a steric congestion around the metal center can also account for the structural and spectroscopic characteristics of the series of complexes, [MCl(tdmme)(dmpm, dmpe or dmpp)]2+ (dmpm = bis(dimethylphosphino)methane, dmpp = 1,3-bis(dimethylphosphino)propane). The X-ray analysis for [CoCl(tdmme)(dmpm or dmpe)](BF4)2 showed that all Co-P bonds in the dmpm complex were shorter by 0.03-0.04 Å than those in the dmpe complex. Furthermore, the first d-d transition energy of the CoIII complexes and the 1JRh-P(tdmme) coupling constants observed for the RhIII complexes indicated an unusual order in the coordination bond strengths of the didentate diphosphines, i.e., dmpm > dmpe > dmpp.  相似文献   

20.
Coordination compounds of chelating 8-methylthioquinoline (MTQ) with the complex fragments ReI(CO)3Cl, [RuII(bpy)2]2+, [RhIII(C5Me5)Cl]+, [IrIII(C5Me5)Cl]+, and PtIVMe4 were synthesized and structurally characterized. Whereas the ruthenium(II) complex displays the strongest preference of bonding to N versus S, the compound (MTQ)PtMe4 shows the most balanced metal-donor bonding within the chelate ring due to a relatively short bond to S (2.319 Å) versus N (2.150 Å). The complex fac-(MTQ)Re(CO)3Cl exhibits a particularly long metal-sulfur bond at 2.472 Å. Cyclic voltammetry of [(MTQ)Ru(bpy)2](PF6)2 reveals one reversible oxidation to RuIII and three closely spaced reduction waves for the coordinated ligands. In comparison with the imine/thioether chelate ligand 1-methyl-2-(methylthiomethyl)-1H-benzimidazole (mmb) the MTQ ligand with its more rigid chelate setting N(sp2)-C(sp2)-C(sp2)-S forms generally shorter M-S bonds and displays stronger π acceptor behaviour.  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