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 8 毫秒
1.
2.
3.
The interaction of OH with Fe(TPP)+, Fe(TDCPP)+, Fe(TMP)+ and Fe(TFPP)+ in 1,2-dichloroethane was studied by titrating FeP solutions with aliquots of a solution of tetrabutylammonium hydroxide in acetonitrile. The number of OH ions (n) coordinated to the FeP and the stability constants (βn) for the FeP-OH complexes were calculated from UV-Vis absorbance data and iron spin states were determined through EPR spectroscopy. Fe(TMP)+ forms a high-spin mono-hydroxo complex, while Fe(TPP)+ and Fe(TDCPP)+ form high-spin bis-hydroxo complexes. To our knowledge, this is the first time that the formation of bis-hydroxo complexes from Fe(TPP)+ has been reported, and this was possible because the studies were carried out in basic organic media. In this same medium, FeIII-FeII reduction upon OH addition to Fe(TFPP)+ was observed, without concomitant formation of the μ-oxo dimeric species [Fe(TFPP)]2O.  相似文献   

4.
Novel mononuclear Fe(II) complexes of tris(pyrazol-1-yl)methane [Fe{HC(pz)3}2]2+ with and p-sulfonatothiacalix[4]arene (TCAS4−) as counterions were obtained. The compounds were characterized by magnetic susceptibility method, IR and UV-Vis spectroscopy. The structure of [Fe{HC(pz)3}2]SiF6 has been analyzed by single-crystal X-ray diffraction. The 1H NMR spectroscopy measurements of [Fe{HC(pz)3}2]2(TСAS) in aqueous solution reveal the outer sphere inclusion of [Fe{HC(pz)3}2]2+ into the cyclophanic cavity of TCAS4−. The temperature induced spin-crossover 1А1 ⇔ 5Т2, accompanied by thermochromism, has been revealed from the temperature dependence of μeff and IR spectra for both complexes. The comparative analysis of magnetochemical and spectroscopy data elucidates the effect of the cyclophanic counterion on the physico-chemical properties of Fe(II) complex.  相似文献   

5.
Lithioamidines {R′N(Li)C(R)NR′, I; R = CH3, R′ = C6H5, p-CH3,C6H4} react with iron(III) chloride
in monoglyme to produce navy-blue, high spin Fe{R′NC(R)NR′}3 complexes which are extremely air and moisture sensitive. The corresponding reaction when R = R′ = C6H5 produces a soluble red complex and an air-stable green complex, whereas when R = H, R′ = C6H5 and R = R′ = C6H5 and the reaction is started at ca. ?20°, red and green complexes respectively are formed. Though all the complexes are formulated Fe{R′NC(R)NR′}3, their properties reflect association through bridging amidino-groups. Iron(II) chloride reacts with I(R = CH3, R′ = p-CH3C6H4) to form two complexes, one crimson and soluble in organic solvents, and one brown and insoluble, which are fomulated [Fe{R′NC(R)NR′}2]n. The iron(III) complexes failed to react with, or were decomposed by, a variety of reducing, electrophilic and nucleophilic reagents, though blue Fe{p-CH3C6H4NC(CH3)N-p-CH3C6H4}3 reacts readily with nitric oxide to form a purple addition complex from which the N-nitroso-compound p-CH3C6H4NC(CH3)N(NO)-p-CH3C6H4 was obtained in high yield. Treatment of the corresponding brown iron(II) complex with nitric oxide gave no reaction.  相似文献   

6.
Iron(II)–dithiocarbamate complexes are used to trap nitrogen monoxide in biological samples, and the resulting nitrosyliron(II)–dithiocarbamate is detected and quantified by ESR. As the chemical properties of these compounds have been little studied, we investigated whether iron dithiocarbamate complexes can redox cycle. The electrode potentials of iron complexes of N-(dithiocarboxy)sarcosine (dtcs) and N-methyl-d-glucamine dithiocarbamate (mgd) are 56 and −25 mV at pH 7.4, respectively, as measured by cyclic voltammetry. The autoxidation and Fenton reaction of iron(II)–dtcs and iron(II)–mgd were studied by stopped-flow spectrophotometry with both iron(II) complexes and dioxygen or hydrogen peroxide in excess. In the case of excess iron(II)–dtcs and –mgd complexes, the rate constants of the autoxidation and the Fenton reaction are (1.6–3.2) × 104 and (0.7–1.1) × 105 M−1 s−1, respectively. In the presence of nitrogen monoxide, the oxidation of iron(II)–dtcs and iron(II)–mgd by hydrogen peroxide is significantly slower (ca. 10–15 M−1 s−1). The physiological reductants ascorbate, cysteine, and glutathione efficiently reduce iron(III)–dtcs and iron(III)–mgd. Therefore, iron bound to dtcs and mgd can redox cycle between iron(II) and iron(III). The ligands dtcs and mgd are slowly oxidized by hydrogen peroxide with rate constants of 5.0 and 3.8 M−1 s−1, respectively.  相似文献   

7.
Simulation of X- and Q-band electron paramagnetic resonance (EPR) spectra of an unsymmetric dinuclear [Mn(2)(II,III)L(mu-OAc)(2)]ClO(4) complex (1), (L is the dianion of 2-{[N,N-bis(2-pyridylmethyl)amino]methyl}-6-{[N-(3,5-di-tert-butyl-2-hydroxybenzyl)-N-(2-pyridylmethyl)amino]methyl}-4-methylphenol) was performed using one consistent set of simulation parameters. Rhombic g-tensors and hyperfine tensors were necessary to obtain satisfactory simulation of the EPR spectra. The anisotropy of the effective hyperfine tensors of each individual (55)Mn ion was further analyzed in terms of intrinsic hyperfine tensors. Detailed analysis shows that the hyperfine anisotropy of the Mn(III) ion is a result of the Jahn-Teller effect and thus an inherent character. In contrast, the anomalous hyperfine anisotropy of the Mn(II) ion is attributed as being transferred from the Mn(III) ion through the spin exchange interaction. The anisotropy parameter for the Mn(II) is deduced as D(II)=-1.26+/-0.2cm(-1). This is the first reported D(II) value for a Mn(II) ion in a weakly exchange coupled mixed-valence Mn(2)(II,III) complex with a bis-mu-acetato-bridge. The [see text] electronic configuration of the Mn(III) ion in 1 is revealed by the negative sign of its intrinsic hyperfine tensor anisotropy, Deltaa(III)=a(z)-a(x,y)=-46cm(-1). Lower spectral resolution of the Q-band EPR spectrum as compared to the X-band EPR spectrum is associated to large line width broadening of the x- and y-components in contrast to the z-component. The origins of the unequal distribution of line width between the z- and x-, y-components are discussed.  相似文献   

8.
Iron(II)-dithiocarbamate complexes are used to trap nitrogen monoxide in biological samples, and the resulting nitrosyliron(II)-dithiocarbamate is detected and quantified by ESR. As the chemical properties of these compounds have been little studied, we investigated whether iron dithiocarbamate complexes can redox cycle. The electrode potentials of iron complexes of N-(dithiocarboxy)sarcosine (dtcs) and N-methyl-d-glucamine dithiocarbamate (mgd) are 56 and -25 mV at pH 7.4, respectively, as measured by cyclic voltammetry. The autoxidation and Fenton reaction of iron(II)-dtcs and iron(II)-mgd were studied by stopped-flow spectrophotometry with both iron(II) complexes and dioxygen or hydrogen peroxide in excess. In the case of excess iron(II)-dtcs and -mgd complexes, the rate constants of the autoxidation and the Fenton reaction are (1.6-3.2) x 10(4) and (0.7-1.1) x 10(5) M(-1) s(-1), respectively. In the presence of nitrogen monoxide, the oxidation of iron(II)-dtcs and iron(II)-mgd by hydrogen peroxide is significantly slower (ca. 10-15 M(-1) s(-1)). The physiological reductants ascorbate, cysteine, and glutathione efficiently reduce iron(III)-dtcs and iron(III)-mgd. Therefore, iron bound to dtcs and mgd can redox cycle between iron(II) and iron(III). The ligands dtcs and mgd are slowly oxidized by hydrogen peroxide with rate constants of 5.0 and 3.8 M(-1) s(-1), respectively.  相似文献   

9.
The absorption and emission spectra of several complexes of neodymium(III) and europium(III) ions have been examined in order to obtain reliable information relating to coordination number, nature of bonding and symmetry around the lanthanide ion. It has been found that steric factors may force the polyhedron of coordination towards geometries less favourable by ligand–ligand repulsion. In general, no correlation has been found to exist between the low intensity of the hypersensitive transitions and high symmetry or low symmetry and high intensity. The results have pointed out the role of covalency in hypersensitivity.  相似文献   

10.
Here we report the formation of iron in hexavalent state, in ozonalysis of iron(III) in alkaline medium. The formation of tetrahedral ion is confirmed by UV-Visible and Mössbauer spectroscopic techniques. The value of isomer shift, δ, of the tetra-oxy anion is consistent with known δ values for various salts of iron(VI) ion.  相似文献   

11.
Catechol dioxygenases are mononuclear non-heme iron enzymes that catalyze the oxygenation of catechols to aliphatic acids via the cleavage of aromatic rings. In the last 20 years, a number of (catecholato)iron(III) complexes have been synthesized and characterized as structural and functional models for the catechol-bound iron(III) form of catechol dioxygenases. This review focuses on the structural and spectroscopic characteristics and oxygenation activity of the title complexes.  相似文献   

12.
13.
‘Intermediate’ spin iron(III) has been identified, using Mössbauer spectroscopy, in materials containing protoporphyrin(IX) iron(III). These materials were all precipitated from acid pH in the presence of a variety of ligands. The implications of the results are discussed both in comparison to other known ‘intermediate’ spin porphyrins and for their relevance to haem proteins.  相似文献   

14.
The aim of this study was to examine the effect of treating of chromium(III) and iron(III) and their combinations on Herpes Simplex Virus type 1 (HSV-1) and Bovine Viral Diarrhoea virus (BVDV) replication. The antiviral efficacies of chromium(III) and iron(III) on HSV-1 and BVDV were evaluated using Real Time PCR method. Moreover, the cytotoxicity of these microelements was examined using the MTT reduction assay. The IC50 (50% inhibiotory concentration) for the chromium chloride was 1100 μM for Hep-2 cells and 1400 μM for BT cells. The IC50 for the iron chloride was 1200 μM for Hep-2 cells and more than1400 μM for BT cells. The concentration-dependent antiviral activity of chromium chloride and iron chloride against HSV-1 and BVDV viruses was observed. In cultures simultaneously treated with (1) 200 μM of CrCl3 and 1000 μM of FeCl3, (2) 1000 μM of CrCl3 and 200 μM of FeCl3, (3) 400 μM of CrCl3 and 800 μM of FeCl3, (4) 800 μM of CrCl3 and 400 μM of FeCl3 a decrease in number of DNA or RNA copies was observed compared with control cells and cells incubated with chromium(III) and iron(III) used separately. The synergistic antiviral effects were observed for chromium(III) and iron(III) against HSV-1 and BVDV.  相似文献   

15.
The Fe3+-doxorubicin complex undergoes reactions that suggest that the complex self-reduces to a ferrous oxidized-doxorubicin free radical species. The Fe3+-doxorubicin system is observed to reduce ferricytochrome c, consume O2 and react with 2,2′-bipyridine. Bipyridine acts as a “ferrous ion scavenger” as it reacts with the ferrous ion produced by Fe3+-doxorubicin self-reduction. In the absence of O2, a ferrous doxorubicin complex accumulates. In the presence of oxygen, Fe2+ recycles back to Fe3+. The rates of these reactions were measured and the Fe3+-doxorubicin self-reduction was determined to be the rate-determining step. The Fe3+-doxorubicin induced inactivation of cytochrome c oxidase and NADH cytochrome c reductase on beef heart submitochondrial particles occurs at a rate similar to Fe3+-doxorubicin self-reduction. Thus the rate at which damage to these mitochondrial enzymes occurs may be controlled by a nonezymatic Fe3+-doxorubicin self-reduction.  相似文献   

16.
For the first time a 1-cyclopropyl substituted tetrazole (C3tz) has been used as a potential ligand for iron(II) spin-transition complexes. The complexation of 1-cyclopropyltetrazol with iron(II) tetrafluoroborate yielded a fine powdered product of [Fe(C3tz)6](BF4)2 being poorly soluble in most common solvents. Single crystals of complex were grown in situ from a solution of ligand and iron(II) hexafluorophosphate, which yielded a hexagonal prismatic crystalline product of [Fe(C3tz)6](PF6)2. A comparison of XRPD data of the homologues [Fe(C3tz)6](BF4)2 and [Fe(C3tz)6](PF6)2 proves them to be homeotypic. The thermally induced spin-crossover phenomenon of [Fe(C3tz)6](BF4)2 complex shows very abrupt spin transitions, with a spin-crossover temperature T1/2 ≈ 180 K which is found to be ≈50 K above the T1/2 of all known iron(II) complexes with n-alkyltetrazoles as ligands. The T1/2 was determined by temperature-dependent 57Fe-Mössbauer, far FT-IR and UV-Vis-NIR spectroscopy as well as temperature dependent magnetic susceptibility measurements (SQUID).  相似文献   

17.
We report here a complete physico-chemical study of the chelation of iron(III) by catechin (L1), an abundant polyphenol in green tea. Using a fruitful combination of electrospray mass spectrometry, absorption spectrophotometry and potentiometry, we have characterized three ferric complexes of catechin (L1Fe, and (L1)3Fe) as well as a ternary complex L1FeNTA when an exogenous ligand (nitrilotriacetic acid) is added to the medium. Thanks to this study, we discuss the influence of an exogenous tetradentate ligand in the ferric recognition processes by catecholate-type polyphenols.  相似文献   

18.

Abstract  

The purple acid phosphatases (PAPs) are the only binuclear metallohydrolases where the necessity for a heterovalent active site [Fe(III)–M(II) (M is Fe, Zn or Mn)] for catalysis has been established. The paradigm for the construction of PAP biomimetics, both structural and functional, is that the ligands possess characteristics which mimic those of the donor sites of the metalloenzyme and permit discrimination between trivalent and divalent metal ions. The donor atom set of the ligand 2-((2-hydroxy-5-methyl-3-((pyridin-2-ylmethylamino)methyl)benzyl)(2-hydroxybenzyl)amino)acetic acid (H3HPBA) mimics that of the active site of PAP although the iron(III) complex of this ligand has been characterized as the tetramer [Fe4(HPBA)2(μ-CH3COO)2(μ-O)(μ-OH)(OH2)2]ClO4·5H2O. The phosphoesterase-like activity of the complex in 1:1 acetonitrile/water has now been investigated using the substrate 2,4-bis(dinitrophenyl)phosphate. The pH dependence of the catalytic rate revealed a non-symmetric bell-shaped profile, with a finite but non-zero rate at high pH. Unlike the traditional approach usually employed to analyse these bell-shaped profiles, the approach used here involved incorporating additional species which contribute to the overall activity. Employing this approach, we show that the complex has a k cat of 1.6 (±0.2) × 10−3 s−1, three kinetically relevant pK a values of 5.3, 6.2 and 8.4, with K M of 7.4 ± 0.6 mM. The kinetic parameters are similar to those reported for heterovalent PAP biomimetics. Additionally, it is observed that, unlike the enzyme, the oxidation state is not the determining factor for catalytic activity.  相似文献   

19.
Rat liver was perfused by Hank's solution, containing desferal (deferoxamine). It was shown that in perfusion of the liver spectrum EPR a signal (g = 4.3; H = 63 G) appears. This signal belongs to desferal complexes, containing intracellular Fe/3/. Desferal transfer to the liver tissue and further formation of desferal complexes there takes place within first 5-10 min of liver perfusion by solution, containing 0.5 mM of desferal.  相似文献   

20.
The preparation of well-defined adducts of the M(guH)(2Cl3 (M = Cr, Fe) and VO(guH)Cl2 types (guH = neutral guanine), by refluxing ligand and metal chloride mixtures in ethanol-triethyl orthoformate, is reported. Characterization studies suggest that the new complexes are probably linear chain-like polymeric species, involving single bridges of bidentate guH ligands between adjacent metal ions. Bidentate bridging guH is most probably coordinated through the N(7) and N(9) imidazole nitrogens. The chloro ligands present in the adducts are exclusively terminal. Infrared evidence rules out the possibility of coordination of guanine through either of its exocyclic potential binding sites (i.e., CO oxygen and NH2 nitrogen) [1].  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