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 296 毫秒
1.
Complexes of group 13 metal halides with pyridine-type ligands (pyridine, pyrazine, and 4, 4´-bipyridine) have molecular, polymeric, or ionic structures containing metal atoms with a coordination number of 4, 5, or 6 depending on the component ratio, the acceptor ability of the halide, the donor ability and the coordination mode of the ligand. The strongest donor-acceptor bond is formed in the 1 : 1 molecular complexes, and their transition to the gas phase is energetically most favorable. The acceptor ability of Lewis acids in the complexes decreases in the series AlCl3 > AlBr3 > GaCl3 > GaBr3 > GaI3. The stability of the complexes with respect to homogeneous dissociation correlates with the donor proton affinity. Group 13 metal trihalides act as catalysts for the pyrolysis of ligands.  相似文献   

2.
The kinetics of the metal exchange reactions between open‐chain Gd(DTPA)2? and Gd(DTPA‐BMA), macrocyclic Gd(DOTA)? and Gd(HP‐DO3A) complexes, and Cu2+ ions were investigated in the presence of endogenous citrate, phosphate, carbonate and histidinate ligands in the pH range 6–8 in NaCl (0.15 M ) at 25 °C. The rates of the exchange reactions of Gd(DTPA)2? and Gd(DTPA‐BMA) are independent of the Cu2+ concentration in the presence of citrate and the reactions occur via the dissociation of Gd3+ complexes catalyzed by the citrate ions. The HCO3?/CO32? and H2PO4? ions also catalyze the dissociation of complexes. The rates of the dissociation of Gd(DTPA‐BMA), catalyzed by the endogenous ligands, are about two orders of magnitude higher than those of the Gd(DTPA)2?. In fact near to physiological conditions the bicarbonate and carbonate ions show the largest catalytic effect, that significantly increase the dissociation rate of Gd(DTPA‐BMA) and make the higher pH values (when the carbonate ion concentration is higher) a risk‐factor for the dissociation of complexes in body fluids. The exchange reactions of Gd(DOTA)? and Gd(HP‐DO3A) with Cu2+ occur through the proton assisted dissociation of complexes in the pH range 3.5–5 and the endogenous ligands do not affect the dissociation rates of complexes. More insights into the interaction scheme between Gd(DTPA‐BMA) and Gd(DTPA)2? and endogenous ligands have been obtained by acquiring the 13C NMR spectra of the corresponding diamagnetic Y(III)‐complexes, indicating the increase of the rates of the intramolecular rearrangements in the presence of carbonate and citrate ions. The herein reported results may have implications in the understanding of the etiology of nephrogenic systemic fibrosis, a rare disease that has been associated to the administration of Gd‐containing agents to patients with impaired renal function.  相似文献   

3.
The kinetics and dynamics of photocatalyzed dissociation of ethanol on TiO2(110) sur-face have been studied using the time-dependent and time-resolved femtosecond two-photon photoemission spectroscopy respectively, in order to unravel the photochemical properties of ethanol on this prototypical metal oxide surface. By monitoring the time evolution of the photoinduced excited state which is associated with the photocatalyzed dissociation of ethanol on Ti5c sites of TiO2(110), the fractal-like kinetics of this surface photocatalytic reaction has been obtained. The measured photocatalytic dissociation rate on reduced TiO2(110) is faster than that on the oxidized surface. This is attributed to the larger defect density on the reduced surface which lowers the reaction barrier of the photocatalytic reaction at least methodologically. Possible reasons associated with the defect electrons for the acceleration have been discussed. By performing the interferometric two-pulse corre-lation on ethanol/TiO2(110) interface, the ultrafast electron dynamics of the excited state has been measured. The analyzed lifetime (24 fs) of the excited state is similar to that on methanol/TiO2(110). The appearance of the excited state provides a channel to mediate the electron transfer between the TiO2 substrate and its environment. Therefore studying its ultrafast electron dynamics may lead to the understanding of the microscopic mechanism of photocatalysis and photoelectrochemical energy conversion on TiO2.  相似文献   

4.
The implementation of surface-induced dissociation (SID) to study the fast dissociation kinetics (sub-microsecond dissociation) of peptides in a MALDI TOF instrument has been reported previously. Silicon nanoparticle assisted laser desorption/ionization (SPALDI) now allows the study of small molecule dissociation kinetics for ions formed with low initial source internal energy and without MALDI matrix interference. The dissociation kinetics of N(CH3)4+ and N(CD3)4+ were chosen for investigation because the dissociation mechanisms of N(CH3)4+ have been studied extensively, providing well-characterized systems to investigate by collision with a surface. With changes in laboratory collision energy, changes in fragmentation timescale and dominant fragment ions were observed, verifying that these ions dissociate via unimolecular decay. At lower collision energies, methyl radical (CH3) loss with a sub-microsecond dissociation rate is dominant, but consecutive H loss after CH3 loss becomes dominant at higher collision energies. These observations are consistent with the known dissociation pathways. The dissociation rate of CH3 loss from N(CH3)4+ formed by SPALDI and dissociated by an SID lab collision energy of 15 eV corresponds to log k = 8.1, a value achieved by laser desorption ionization (LDI) and SID at 5 eV. The results obtained with SPALDI SID and LDI SID confirm that (1) the dissociation follows unimolecular decay as predicted by RRKM calculations; (2) the SPALDI process deposits less initial energy than LDI, which has advantages for kinetics studies; and (3) fluorinated self-assembled monolayers convert about 18% of laboratory collision energy into internal energy. SID TOF experiments combined with SPALDI and peak shape analysis enable the measurement of dissociation rates for fast dissociation of small molecules.  相似文献   

5.
Results of mass spectrometric studies are reported for the collisional dissociation of Group XI (Cu, Ag, Au) metal ion complexes with fatty acids (palmitic, oleic, linoleic and α-linolenic) and glycerolipids. Remarkably, the formation of M2H+ ions (M = Cu, Ag) is observed as a dissociation product of the ion complexes containing more than one metal cation and only if the lipid in the complex contains a double bond. Ag2H+ is formed as the main dissociation channel for all three of the fatty acids containing double bonds that were investigated while Cu2H+ is formed with one of the fatty acids and, although abundant, is not the dominant dissociation channel. Also, Cu(I) and Ag(I) ion complexes were observed with glycerolipids (including triacylglycerols and glycerophospholipids) containing either saturated or unsaturated fatty acid substituents. Interestingly, Ag2H+ ion is formed in a major fragmentation channel with the lipids that are able to form the complex with two metal cations (triacylglycerols and glycerophosphoglycerols), while lipids containing a fixed positive charge (glycerophospocholines) complex only with a single metal cation. The formation of Ag2H+ ion is a significant dissociation channel from the complex ion [Ag2(L–H)]+ where L = Glycerophospholipid (GP) (18:1/18:1). Cu(I) also forms complexes of two metal cations with glycerophospholipids but these do not produce Cu2H+ upon dissociation. Rather organic fragments, not containing Cu(I), are formed, perhaps due to different interactions of these metal cations with lipids resulting from the much smaller ionic radius of Cu(I) compared to Ag(I).  相似文献   

6.
The kinetics of formation of oxalatobis(biguanide)chromium(III) ion from diaquobis(biguanide)chromium(III) ion and oxalate has been studied in aqueous solution (pH ca. 3–4). The results indicate ion-pair formation between the reacting species followed by reaction by an essentially dissociative mechanism (SN1IP). The kinetics of dissociation of the oxalatobis(biguanide)chromium(III) ion in acid medium has also been studied. The reactions involved are the loss of the two biguanide ligands in stages and by proper choice of acid concentrations and temperature the rates of the two steps could be followed separately. The results indicate SN2CA mechanism.  相似文献   

7.
The interaction of [Pd(DAP)(H2O)2]2+ (DAP = 1,3-diaminopropane) with some selected bio-relevant ligands, containing different functional groups, were investigated. The ligands used are dicarboxylic acids, amino acids, peptides and DNA constituents. Stoichiometry and stability constants of the complexes formed are reported at 25°C and 0.1 M ionic strength. The results show the formation of 1:1 complexes with amino acids and dicarboxylic acids. The effect of chelate ring size of the dicarboxylic acid complexes on their stability constants is examined. Peptides form both 1:1 complexes and the corresponding deprotonated amide species. DNA constituents form 1:1 and 1:2 complexes. The effect of dioxane on the acid dissociation constants of CBDCA and the formation constant of its complex with Pd(DAP)2+ was reported. The kinetics of hydrolysis of glycine methyl ester bound to [Pd(DAP)(H2O)2]2+ was studied at 25°C and 0.1M ionic strength.   相似文献   

8.
For many years it has been known that the nine water molecules in [M3Q4(H2O)9]4+ cuboidal clusters (M=Mo, W; Q=S, Se) can be replaced by entering ligands, such as chloride or thiocyanate, and kinetic studies carried out mainly on the substitution of the first water molecule at each metal centre reveal that the reaction at the three metal centres occurs with statistical kinetics; that is, a single exponential with a rate constant corresponding to the reaction at the third centre is observed instead of the expected three‐exponential kinetic trace. Such simplification of the kinetic equations requires the simultaneous fulfilment of two conditions: first that the three consecutive rate constants are in statistical ratio, and second that the metal centres behave as independent chromophores. The validity of those simplifications has been checked for the case of the reaction of [Mo3S4(H2O)9]4+ with Cl? by using DFT and TD‐DFT theoretical calculations. The results of those calculations are in agreement with the available experimental information, which indicates that the H2O ligands trans to the μ‐S undergo substitution much faster than those trans to the μ3‐S. Moreover, the energy barriers for the substitution of the first water molecule at the three metal centres are close to each other, the differences being compatible with the small changes in the numerical values of the rate constants required for observation of statistical kinetics. TD‐DFT calculations lead to calculated electronic spectra, which are in reasonable agreement with those experimentally measured, but the calculations do not indicate that the three metal centres behave as independent chromophores, although the mathematical conditions required for simplification of the kinetic traces to a single exponential are reasonably well fulfilled at certain wavelengths. A re‐examination of the kinetics of the reaction by using global fitting procedures yields results, which are compatible with statistical kinetics, although an alternative interpretation in which substitution only occurs at a single metal centre under reversible conditions is also possible.  相似文献   

9.
The results of studying a unique synthesis method, the kinetics and mechanism of dissociation in acid media, and the stability of a new octacoordinated zirconium complex with two bidentate acetate ligands and one 5,10,15,20-tetraphenylporphine ligand, (AcO)2ZrTPP, are presented. The method of synthesis is distinguished by the use of stable ZrOCl2 in the complex formation reaction, instead of readily hydrolyzed zirconium tetrachloride, in boiling phenol, which exhibits the reduction properties. The UV, visible, IR, and 1H NMR spectral parameters and the values of chromatographic mobility and stability of the complex are obtained. A protolytic dissociation scheme, which is universal for mixed porphyrin-containing zirconium complexes, is substantiated using data on the dissociation rates of an (Cl)2ZrTPP analogue, depending on the temperature and acidity of the medium. The scheme includes a trimolecular rate-determining step (the complex ionized at one or both noncyclic ligands and two nonionized sulfuric acid molecules) and kinetically significant equilibria of dissociation of the noncyclic acido ligands and sulfuric acid.  相似文献   

10.
The extraction or metal ions with a series of aliphatic n-arsonic acids in chloroform is described. Homologous (C4 to C9) arsonic acids were examined for 20 metal ions. Results are discussed in relation to acid dissociation constants and are compared with results for analogous carboxylic acids. No true compounds are formed between the metal ions and the arsonic acids. Various separations of metal ions are possible.  相似文献   

11.
Equilibrium constants are deter mined for the protonation and metal complexation of the nickel(II) complexes with 4‐methyl‐4,7‐diazadecanediamide (4‐Me‐L‐2,2,2), 4,7‐dimethyl‐4,7‐diazadecanediamide (4,7‐N,N′‐Me2‐L‐2,2,2), 4‐ethyl‐4,7‐diazadecanediamide (4‐Et‐L‐2,2,2), and 4‐methyl‐4,8‐ diazaundecane diamide (4‐Me‐L‐2,3,2), in 0.10 M KCl at 25.0°C. The formation kinetics of these nickel (II) complexes have been studied under the same conditions with use of the stopped‐flow technique. The possible path ways for the complexation reaction of nickel (II) with these ligands are discussed. The first metal‐nitrogen bond formation is proposed as the rate‐determining step for the reactions of nickel (II) with the unprotonated ligands; proton loss is the rate‐limiting step in the reactions of nickel (II) with the monoprotonated ligands. Similarly, in dissociation reactions of these nickel (II) complexes, the rate‐determining step for the water dissociation pathway is the break age of the second nickel‐nitrogen bond; the rate‐determining step for the proton‐assisted path way is the protonation of the released amino group. The important factors determining the reactivity of these complexes are considered. The kinetic results of the formation and dissociation reactions of these complexes are consistent with dissociative mechanism.  相似文献   

12.
Functionalization of the inert C? H bonds of unsaturated molecules by transition metal complex is an important means to form new C? C bonds. The functionalization is usually initiated by the ligand dissociation of a complex. In this paper we employ both ab initio and density functional methods to explore the influence of central metals, conformation, solvent and protonation on the ligand dissociation of the (hfac‐O,O)2M(L)(py) complexes [M=Rh(III) or Ir(III), hfac‐O,O=k2‐O,O‐1,1,1,5,5,5‐hexafluoroacetylacetonato, L=CH3, CH3CO2, (CH3CO)2CH, CH3O or OH, py=pyridine]. We demonstrate that ligand pyridine dissociates more easily than the "L" ligands under study in aprotic solvent and gas phase and the dissociation of pyridine is more facile in the trans‐conformation than in the cis‐isomer. These phenomena are rationalized based on electronic structure and molecular orbital interactions. We show that solvation only slightly stabilizes the complexes and does not change the ligand dissociation ordering. In particular, we show that pyridine is no longer the labile ligand in protic media. Instead, the oxygen‐containing ligands (apart from those like hfac that form a cyclic structure with the central metal) that coordinate to the central metal via oxygen atom become the labile ones. Finally our calculations indicate that hfac is a stable ligand, even in protic media.  相似文献   

13.
Drug‐binding kinetics could play important roles in determining the efficacy of drugs and has caught the attention of more drug designers. Using the dissociation of 1H‐pyrrolo[2,3‐b]‐pyridines from the focal adhesion kinase as an example, this work finds that steered molecular dynamics simulations could help screen compounds with long‐residence times. It also reveals a two‐step mechanism of ligand dissociation resembling the release of ADP from protein kinase A reported earlier. A phenyl group attaching to the pyrrole prolongs residence time by creating a large activation barrier for transition from the bound to the intermediate state when it becomes exposed to the solvent. Principal component analysis shows that ligand dissociation does not couple with large‐scale collective motions of the protein involving many of its amino acids. Rather, a small subset of amino acids dominates. Some of these amino acids do not contact the ligands directly along the dissociation pathways and could exert long‐range allosteric effects. © 2018 Wiley Periodicals, Inc.  相似文献   

14.
Salt effects on the kinetics of acid hydrolysis of some novel hydrophilic iron(II) complexes have been investigated in aqueous medium. The ligands are derived from the condensation of amino acids (glycine, L-alanine, L-leucine, L-isoleucine, DL-methionine, DL-serine or L-phenylalanine) and sodium 2-hydroxybenzaldehyde-5-sulfonate. The reaction was studied under conditions of pseudo first order kinetics. The general rate equation was suggested as follows: rate = k obs[complex], where k obs = k 2[H+]. The reaction rate decreases with an increase of the salt concentration.  相似文献   

15.
New Ni(II) complexes with bioactive bishydrazones ligands based on (pyridine-2-carboxaldhyde)-3-isatin, (2-acetylpyridine)-3-isatin, and (2-benzoylpyridine)-3-isatin have been synthesized and characterized by elemental analysis, conductivity measurements, IR and UV-Vis spectroscopy, and thermal analysis. The complexes stoichiometry and formation constants have been determined. The results suggest that isatinbishydrazones act as neutral tridentate ligands with ONN sites coordinating to the metal ion via isatin C=O, azomethine CR=N, and pyridine C=N groups to give [Ni(L)H2O]Cl2·2H2O, (L = neutral tridentate isatin hydrazone ligand). Kinetics and thermodynamic parameters of the complexes thermal decomposition have been elucidated from the thermal data using Coats and Redfern method, which has confirmed the first order kinetics.  相似文献   

16.
The specific features revealed in the structure of germanium(IV) compounds with ligands in the form of anions of polybasic acids (monoamine, diamine, and triamine complexones, i.e., hydroxyethylidene-diphosphonic and carboxylic acids) have been considered. The influence of the individuality of specific acids on the structure type (mononuclear, binuclear, trinuclear, hexanuclear, polynuclear), the coordination mode of monodentate ligands and donor atoms of polydentate ligands (terminal, bridging, chelating, chelating-bridging), and variants of the coordination of polydentate ligands, i.e., anions of polybasic acids, with metal atoms (germanium, rare-earth elements, copper, barium), as well as on the dependence of the Ge-O bond length on the individual nature of ligands (OH, H2O, O(oxo)) and donor atoms of polydentate ligands α- and β-O(carb), O(hydr), O(P)] and their function in the structure (terminal, bridging, chelating, chelating-bridging), has been analyzed using 28 homometallic and heterometallic complexes as an example.  相似文献   

17.
Summary In order to investigate the kinetics of CO2dissociation on supported nickel catalysts, a novel technique, which can give the surface reaction rate constants with no information on the number of active sites, was developed. It was revealed that CO2dissociation was more enhanced on TiO2support than on other metal oxide ones. The activity pattern and activation energies were in good agreement with those obtained by a conventional pulse technique using the number of active sites, suggesting the validity of the present technique for investigating the kinetics of the surface reaction.</o:p>  相似文献   

18.
Reactions of 3-acetyl-2,5-dimethylthiophene with thiosemicarbazide and semicarbazide hydrochloride resulted in the formation of new heterocyclic ketimines, 3-acetyl-2,5-dimethylthiophene thiosemicarbazone (C9H13N3OS2 or L1H) and 3-acetyl-2,5- dimethylthiophene semicarbazone (C9H13N3OS or L2H), respectively. The Pd(II) and Pt(II) complexes have been synthesized by mixing metal salts in 1:2 molar ratios with these ligands by using microwave as well as conventional heating method for comparison purposes. The authenticity of these ligands and their complexes has been established on the basis of elemental analysis, melting point determinations, molecular weight determinations, IR, 1H NMR and UV spectral studies. These studies showed that the ligands coordinate to the metal atom in a monobasic bidentate manner and square planar environment around the metal atoms has been proposed to the complexes. Both the ligands and their complexes have been screened for their antimicrobial activities. The antiamoebic activity of both the ligands and their palladium compounds against the protozoan parasite Entamoeba histolytica has been tested.  相似文献   

19.
Borges F  Guimarães C  Lima JL  Pinto I  Reis S 《Talanta》2005,66(3):670-673
Studies on the complexation of copper(II) by phenolic acids, as ligand models of humic substances were done by potentiometry. The acids under study were: 3,4-dihydroxyhydrocinnamic acid or hydrocaffeic acid (1), 3,4-dihydroxyphenylacetic acid (2) and 3,4-dihydroxybenzoic acid or protocatechuic acid (3). Acidity constants of the ligands and the formation constants of metal-ligand complexes were evaluated by computer programs. The carboxylic group of the phenolic acids has different pKa1 values, being the dissociation constants intrinsically related with the distance between the function and the aromatic nucleus. The results obtained allow concluding that acidity constants of the catechol moiety of the compounds are similar with pKa2 and pKa3 values between 9.47-9.41 and 11.55-11.70. The complexation properties of the three ligands towards copper(II) ion are quite similar, being the species found not different either in nature or stability. Although the model ligands have some structural differences no significant differences were found in their complexation properties towards copper(II). So, it can be postulated that complexation process is intrinsically related with the presence of a catechol group.  相似文献   

20.
The kinetics and mechanism of dissociation of copper(II) complexes with octaethylporphyrin and of manganese(III) with tetraphenylporphyrin were studied in relation to the composition of an H2SO4-AcOH binary solvent. Concentration ranges of H2SO4 in the composition of the binary solvent, where the kinetics of dissociation of metal porphyrins obeys common kinetic regularities, were found using data for porphyrin complexes of transition metals and indium. The reaction order with respect to proton and the nature of the active form of this latter change in the range of solvent compositions, where the H 0-c0 H 2 SO 4 dependence is nonlinear.  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