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.
We report the reactivity of three binuclear non-heme Fe(III) compounds, namely [Fe2(bbppnol)(μ-AcO)(H2O)2](ClO4)2 (1), [Fe2(bbppnol)(μ-AcO)2](PF6) (2), and [Fe2(bbppnol)(μ-OH)(Cl)2]·6H2O (3), where H3bbppnol = N,N′-bis(2-hydroxybenzyl)-N,N′-bis(2-methylpyridyl)–1,3-propanediamine-2-ol, toward the hydrolysis of bis-(2,4-dinitrophenyl)phosphate as models for phosphoesterase activity. The synthesis and characterization of the new complexes 1 and 3 was also described. The reactivity differences observed for these complexes show that the accessibility of the substrate to the reaction site is one of the key steps that determinate the hydrolysis efficiency.  相似文献   

2.
The pseudobinary systems ACl?FeCl3 (A=Na, K, Rb, Cs) were reinvestigated by means of differential thermal analysis and X-ray powder diffraction. The existence of the compounds AFeCl4 (A=Na?Cs) and Cs3Fe2Cl9 could be confirmed; Cs3Fe2Cl9 is a stable compound which decomposes to CsCl and CsFeCl4 above 270°C. Additionally, two Rb-compounds—Rb3FeCl6 and Rb3Fe2Cl9—were found, which decompose, when heated, in the solid state. Rb3Fe2Cl9 is isotypic with the analogous Cs-compound; Rb3FeCl6 has the Cs3BiCl6 structure. Cs3FeCl6 is isotypic with Cs3CrCl6, a recently found orthorhombic variant of the elpasolite type.  相似文献   

3.
Six new μ-terephthalato iron(III) binuclear complexes have been prepared and identified: [Fe2(TPHA)(L)4]-(ClO4)4 [L = 2,2′-bipyridine (bpy); 1,10-phenanthroline (phen); 4,4′-dimethyl-2,2′-bipyridine (Me2bpy); 5-methyl-1,10-phenanthroline (Me-phen); 5-chloro-1,10-phenanthroline (Cl-phen) and 5-nitro-1,10-phenanthroline (NO2-phen)]; where TPHA = the terephthalate dianion. Based on the elemental analyses, molar conductance and magnetic moments of room-temperature measurements, and spectroscopic studies, extended TPHA-bridged structures consisting of two iron(III) ions, each in an octahedral environment, are proposed for these complexes. The [Fe2(TPHA)(Me-phen)4](ClO4)4 (1) and [Fe2(TPHA)(phen)4](ClO4)4 (2) complexes were characterized by variable temperature magnetic susceptibility (4–300 K) measurements and the observed data were successfully simulated by the equation based on the spin Hamiltonian operator, Ĥ = −2 1 Ŝ 2, giving the exchange integrals J = −1.05 cm−1 for (1) and J = −9.28 cm−1 for (2). This result indicates the presence of a weak antiferromagnetic spin-exchange interaction between the metal ions within each molecule. The influence of the terminal ligand methyl substituents on magnetic interactions between the metals is also discussed. This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

4.
5.
The rate of reaction of NO 2 ion with various FeIII porphyrins in the presence of PPh3 is shown to depend on the redox potential of the FeIII center. There is a linear relationship between the ease of reduction of the FeIII to FeII and the kinetics for the formation of the FeII porphyrin nitrosyl adduct, with concomitant oxidation of PPh3 to PPh3O. Cyclic voltammograms show reversible one-electron reductions that can be ascribed to the FeIII/FeII couple ranging from E1/2 = −343 to −145 mV (versus Ag/AgCl). The order of increasing half-wave reduction potentials for the FeIII/FeII porphyrin redox centers studied is octaethylporphyrin > etioporphyrin I > deuteroporphyrin IX dimethyl ester > protoporphyrin IX dimethyl ester > α,β,γ,δ-tetraphenylporphyrin. This sequence of redox potentials complements the pseudo first-order kinetics ( to m s −1) for the oxidation of PPh3 and subsequent FeII porphyrin nitrosyl adduct formation. The rates of reaction of biomimetic FeIII porphyrins with NO 2 ion demonstrate how metal center redox properties are influenced by the surrounding ligand. In this paper we have elucidated a possible mechanistic control for the rate of this reaction.  相似文献   

6.
Threshold collision-induced dissociation techniques are employed to determine bond dissociation energies (BDEs) of mono- and bis-complexes of alkali metal cations, Li+, Na+, K+, Rb+, and Cs+, with indole, C8H7N. The primary and lowest energy dissociation pathway in all cases is endothermic loss of an intact indole ligand. Sequential loss of a second indole ligand is observed at elevated energies for the bis-complexes. Density functional theory calculations at the B3LYP/6-31G level of theory are used to determine the structures, vibrational frequencies, and rotational constants of these complexes. Theoretical BDEs are determined from single point energy calculations at the MP2(full)/6-311+G(2d,2p) level using the B3LYP/6-31G* geometries. The agreement between theory and experiment is very good for all complexes except Li+ (C8H7N), where theory underestimates the strength of the binding. The trends in the BDEs of these alkali metal cation-indole complexes are compared with the analogous benzene and naphthalene complexes to examine the influence of the extended pi network and heteroatom on the strength of the cation-pi interaction. The Na+ and K+ binding affinities of benzene, phenol, and indole are also compared to those of the aromatic amino acids, phenylalanine, tyrosine, and tryptophan to elucidate the factors that contribute to the binding in complexes to the aromatic amino acids. The nature of the binding and trends in the BDEs of cation-pi complexes between alkali metal cations and benzene, phenol, and indole are examined to help understand nature's preference for engaging tryptophan over phenylalanine and tyrosine in cation-pi interactions in biological systems.  相似文献   

7.
Bromobutide, a novel herbicide, was labeled with carbon-14 independently at the carbonyl group and the phenyl ring for use in metabolic studies. 14C-Carbonation of neopentylmagnesium chloride (3) gave 3,3-dimethyl[1-14C]butyric acid (4a) quantitatively. Chlorination of 4a with thionyl chloride followed by alpha-bromination with bromine yielded 2-bromo-3,3-dimethyl[1-14C]-butyryl halide (5a), which was subsequently condensed with alpha, alpha-dimethylbenzylamine (6a) to afford [carbonyl-14C]bromobutide (1a). The overall yield of 1a was 76% from barium [14C]-carbonate (2). Similarly, condensation of alpha, alpha-dimethyl[phenyl-14C]benzylamine (6b), which was prepared from alpha-methyl[phenyl-U-14C]styrene (7) in three steps, with 2-bromo-3,3-dimethylbutyryl halide (5b) gave [phenyl-14C]bromobutide (1b) in 67% yield after purification. The specific activities of 1a and 1b were 1.38 and 0.781 GBq/mmol (37.2 and 21.1 mCi/mmol), respectively.  相似文献   

8.
Li T  Li F  Lü J  Guo Z  Gao S  Cao R 《Inorganic chemistry》2008,47(13):5612-5615
A novel pentadecatungstate, [H 6Ce 2(H 2O)Cl(W 5O 18) 3] (7-) ( 1), constructed by a dinuclear cerium unit and 15-member ring WO 6 units was prepared and characterized by single-crystal X-ray diffraction. Polyanion 1 exhibits blue photoluminescence with an emission maximum at 488 nm, which is characteristic of cerium(III) transitions from 5d to (2)F 5/2 states. Furthermore, the study of the electrochemical property investigation of 1 shows two reversible redox peaks ascribing to two-electron processes.  相似文献   

9.
Dinuclear Cd(II), Cu(II), and Zn(II) complexes of L2OH (L2OH = 1,3-bis(1,4,7-triazacyclonon-1-yl)-2-hydroxypropane) are compared as catalysts for cleavage of the RNA analogue HpPNP (HpPNP = 2-hydroxypropyl 4-nitrophenyl phosphate) at 25 degrees C, I = 0.10 M (NaNO(3)). Zn(II) and Cu(II) readily form dinuclear complexes at millimolar concentrations and a 2:1 ratio of metal ion to L2OH at neutral pH. The dinuclear Zn(2)(L2O) and Cu(2)(L2O) complexes have a bridging alkoxide group that brings together the two cations in close proximity to facilitate cooperative catalysis. Under similar conditions, the dinuclear complex of Cd(II) is a minor species in solution; only at high pH values (pH 10.4) does the Cd(2)(L2O) complex become the predominant species in solution. Analysis of the second-order rate constants for cleavage of HpPNP by Zn(2)(L2O) is straightforward because a linear dependence of pseudo-first-order rate constant on dinuclear complex is observed over a wide pH range. In contrast, plots of pseudo-first-order rate constants for cleavage of HpPNP by solutions containing a 2:1 ratio of Cd(II) to L2OH as a function of increasing L2OH are curved, and second-order rate constants are obtained by fitting the kinetic data to an equation for the formation of the dinuclear Cd(II) complex as a function of pH and [L2OH]. Second-order rate constants for cleavage of HpPNP by these dinuclear complexes at pH 9.3 and 25 degrees C vary by 3 orders of magnitude in the order Cd(2)(L2O) (2.8 M(-)(1) s(-)(1)) > Zn(2)(L2O) (0.68 M(-)(1) s(-)(1)) > Cu(2)(L2O) (0.0041 M(-1) s(-1)). The relative reactivity of these complexes is discussed in terms of the different geometric preferences and Lewis acidity of the dinuclear Zn(II), Cu(II), and Cd(II) complexes, giving insight into the importance of these catalyst properties in the cleavage of phosphate diesters resembling RNA.  相似文献   

10.
The luminescence lifetime of TbL2 with a pendant 15-crown-5 increased by 65% to 2.95 ms with an [Na+] concentration of 0.13 M in aqueous solution; the maximum amplification of the luminescence intensity of TbL1 containing aza-15-crown-5 reached a factor of 47 upon addition of the aromatic antenna p-chlorobenzoate 1.  相似文献   

11.
Zhou  Yuming  Ge  Yuhua  Gou  Shaohua  Ju  Huangxian  Liu  Juzheng 《Transition Metal Chemistry》1997,22(4):347-349
A new macrocyclic dinuclear iron(III) complex, prepared by condensing 2,6-diformylpyridine N-oxide with 1,3-diaminopropane in the presence of FeCl2·4H2O, and formulated as Fe2LCl6C2H5OH·6H2O, was characterized by its elemental analyses, by i.r., Mo¨ssbauer and e.p.r. spectra, and was investigated electrochemically. The results show that two iron ions are situated in the same chemical environment in the compound and that the electrode reaction can be considered as a double electron-transfer process. The diffusion coefficient is 2.16×10–6 cm2 s–1 and E° –0.169 V (versus s.c.e.) and = 0.37. The average electron transfer rate constant ko is 2.8×10–3 cm s–1.  相似文献   

12.
A simple triacetylide complex of chromium(III) is synthesized for use as a potential precursor to metal-dicarbide clusters. Reaction of Me(3)SiCCLi with [(Me(3)tacn)Cr(CF(3)SO(3))(3)] (Me(3)tacn = N,N',N"-trimethyl-1,4,7-triazacyclononane) in THF generates [(Me(3)tacn)Cr(CCSiMe(3))(3)], which subsequently reacts with Bu(4)NF to supply [(Me(3)tacn)Cr(CCH)(3)] as an air-stable orange solid. The crystal structure of this unprecedented triacetylide complex reveals octahedral coordination of the chromium center, with linear Cr-C(triple bond)C bond angles and C(triple bond)C bond distances essentially identical to the corresponding distance in acetylene. Crystallization of the complex from a DMF solution containing K(CF(3)SO(3)) leads to the sandwich complex ([(Me(3)tacn)Cr(CCH)(3)](2)K)(+), in which the K(+) ion is coordinated in a side-on fashion by each of the six C(triple bond)C units. With the larger Cs(+) cation, a triangular ([(Me(3)tacn)Cr(CCH)(3)](3)Cs)(+) complex is instead observed. The magnetic properties of these alkali metal complexes are indicative of weak antiferromagnetic exchange between Cr(III) centers, with J = -0.8 and -0.3 cm(-1), respectively.  相似文献   

13.
14.
Crystal structures of three fluorides (NH4)2NaFeF6, (Fe), (NH4)2NaGaF6, (Ga), and (NH4)2NaCrF6, (Cr), as well as a substituted compound [(NH4)1−xKx]2KAlF6 (x≈0.17), (Al), have been refined using single-crystal and powder X-ray diffraction techniques. All these four ammonium hexafluorides have a cubic elpasolite-type structure and crystallize in the space group Fmm with lattice constants a=8.483(3), 8.450 (3), 8.4472(2) and 8.724(3) Å for compounds (Fe), (Ga), (Cr) and (Al), respectively. The effective ionic radius of the ammonium ion calculated from those compounds has a mean value of R=1.729 Å for CN=12. An ultraviolet-visible absorption spectrum of (NH4)2NaCrF6, measured at room temperature, gives a crystal field (Dq=1575 cm−1) and Racah parameters (B=758 cm−1 and C=3374 cm−1). Abnormal anisotropic thermal parameters of fluorine atoms have been observed in the compound (Al), and interpreted to arise from four strong hydrogen bonds (F…H-N) that are distributed in a square form around each fluorine atom.  相似文献   

15.
The thermal decomposition of thiosulphatobismuthates(III) of alkali metals was investigated. The general formulae of the thiosulphatobismuthates are M3[Bi(S2O3)3]·H2O where M = Na, K, Rb or Cs, and M2Na[Bi(S2O3)3]·H2O where M = K or Cs.Typical thermal curves for thiosulphatobismuthates(III) and the results obtained in thermal, X-ray, chemical and spectrophotometrical analyses of the decomposition products are shown. The results were used to determine three stages of the thermal decomposition. At the first stage, at about 200°C, hydrated compounds are dehydrated. At the second stage, above 200°C, there is a rapid decrease in mass which is caused by evolving sulphur dioxide; bismuth sulphide and an intermediate decomposition product are formed. At about 320°C the thermal decomposition products are bismuth sulphide and alkali metal sulphate.  相似文献   

16.
Two calix[4]arenes with four 2-pyridyhnethyl-l-oxide pendant groups at the lower rim have been synthesized, and their Tb(III) and Eu(III) complexes are fluorescent upon UV light excitation at 312 nm. The complexes are not stable in aqueous solution, completely losing their luminescent properties.  相似文献   

17.
The equilibria and the kinetics of the binding of Iron(III) to salicylhydroxamic (SHA) and benzohydroxamic (BHA) acids have been investigated in aqueous solution (I = 1 M (HClO(4)/NaClO(4)), T = 298 K) using spectrophotometric and stopped-flow methods. Whereas Iron(III) forms a 1:1 complex (ML) with BHA, it forms both ML and M(2)L complexes with SHA. The presence of M(2)L in aqueous medium is corroborated by FTIR measurements. The reactive form of Iron(III) is the hydrolyzed species FeOH(2+), which binds to the O,O site in ML and to the O,O and O(P),N (P = phenolate) sites in M(2)L, inducing full deprotonation of the latter. The reaction pathway is discussed in terms of a multistep mechanistic scheme in which the metal-ligand interaction is coupled to hydrolysis and self-aggregation steps of Iron(III). The observation and characterization of M(2)L as a stable species is important because it contains the -Fe-O-N-Fe- sequence, which constitutes the repetitive motif of the SHA-based metallacrown ring and provides the rationale for 12-MC-4 metallacrowns. In the framework of this study, the kinetics of the Iron(III) dimerization and trimerization have also been investigated using the stopped-flow method to perform dilution jumps. The reaction scheme put forward involves two parallel steps (FeOH(2+) + FeOH(2+) and Fe(3+) + FeOH(2+)) that lead to formation of the Fe(2)(OH)(2)(4+) dimer and a slower step (FeOH(2+) + Fe(2)(OH)(2)(4+)) to form the trimer species. The kinetics of the last step have been investigated here for the first time, and the results deduced indicate that, of the two possible trimer structures reported in the literature, Fe(3)(OH)(3)(6+) and Fe(3)(OH)(4)(5+), the latter prevails by far.  相似文献   

18.
Matrix-assisted laser desorption ionization mass spectrometry (MALDI-MS) and electrospray ionization mass spectrometry (ESI-MS) are used to evaluate the alkali metal ion binding selectivities of a series of calixarenes. Each calixarene of interest is mixed with one or more alkali metal salts (1:100 ratio of calixarene to metal), either in the ESI solution or on the MALDI probe surface, and the relative binding selectivities are directly determined from the intensities of the calixarene/metal complexes in the mass spectra. For t-butylcalix[4]arene-tetraacetic acid tetraethyl ester (calixarene 1), complexation of Na+ is favored over complexation of K+, in agreement with prior solution results obtained by conventional methods. For the three calixarenes that do not have t-butyl groups on the upper rims, the calixarenes preferentially bind K+ over Na+, thus demonstrating that size selective complexation can be probed with both the ESI and MALDI methods. Collision-activated dissociation results indicate that the phenyl oxygens, but not necessarily the ethoxy ethyl oxygens of the lower rims, are the primary binding sites for the alkali metal ions.  相似文献   

19.
The luminescence and circularly polarized luminescence (CPL) spectra of M(I)[Eu((+)-hfbc)(4)] show a similar behavior to the exciton CD in the intraligand π-π* transitions when the alkali metal ions and solvents are manipulated. There is a difference in susceptibility in solvation toward the alkali metal ions but not toward the Eu(III) ion, as in the case of axially symmetric DOTA-type compounds. The remarkable CPL in the 4f-4f transitions provide much more information on the stereospecific formation of chiral Eu(III) complexes, since CPL spectroscopy is limited to luminescent species and reflects selectively toward helicity of the local structural environment around the lanthanide(III). While in comparison, exciton CD reveals the chiral structural information from the helical arrangement of the four bladed chelates. Of special importance, the observation of the highest CPL activities measured to date for lanthanide(III)-containing compounds (i.e., Eu and Sm) in solution supports the theory that the chirality of lanthanide(III) in the excited state corresponds to that in the ground state, which was derived from the exciton CD.  相似文献   

20.
Relevant to mixtures studied in solution and utilised in styrene polymerisation but hitherto not characterised in the solid, two crystal structures of alkali metal alkoxide-magnesium bis(alkyl) co-complexes have been determined, revealing that in binding to the C2O tripodal face of a common organomagnesiate anion, Na+ prefers O-coordination, whereas K+ prefers C-coordination.  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