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 31 毫秒
1.
Two dinuclear manganese complexes, [Mn(2)BPMP(mu-OAc)(2)].ClO(4) (1, where BPMP is the anion of 2,6-bis([N,N-di(2-pyridinemethyl)amino]methyl)-4-methylphenol) and [Mn(2)L(mu-OAc)(2)].ClO(4) (2, where L is the trianion of 2,6-bis([N-(2-hydroxy-3,5-di-tert-butylbenzyl)-N-(2-pyridinemethyl)amino]methyl)-4-methylphenol), undergo several oxidations by laser flash photolysis, using ruthenium(II)-tris-bipyridine (tris(2,2-bipyridyl)dichloro-ruthenium(II) hexahydrate) as photo-sensitizer and penta-amminechlorocobalt(III) chloride as external electron acceptor. In both complexes stepwise electron transfer was observed. In 1, four Mn-valence states from the initial Mn(2)(II,II) to the Mn(2)(III,IV) state are available. In 2, three oxidation steps are possible from the initial Mn(2)(III,III)state. The last step is accomplished in the Mn(2)(IV,IV) state, which results in a phenolate radical. For the first time we provide firm spectral evidence for formation of the first intermediate state, Mn(2)(II,III), in 1 during the stepwise light-induced oxidation. Observation of Mn(2)(II,III) is dependent on conditions that sustain the mu-acetato bridges in the complex, i.e., by forming Mn(2)(II,III) in dry acetonitrile, or by addition of high concentrations of acetate in aqueous solutions. We maintain that the presence of water is necessary for the transition to higher oxidation states, e.g., Mn(2)(III,III) and Mn(2)(III,IV) in 1, due to a bridging ligand exchange reaction which takes place in the Mn(2)(II,III) state in water solution. Water is also found to be necessary for reaching the Mn(2)(IV,IV) state in 2, which explains why this state was not reached by electrolysis in our earlier work (Eur. J. Inorg. Chem (2002) 2965). In 2, the extra coordinating oxygen atoms facilitate the stabilization of higher Mn valence states than in 1, resulting in formation of a stable Mn(2)(IV,IV) without disintegration of 2. In addition, further oxidation of 2, led to the formation of a phenolate radical (g = 2.0046) due to ligand oxidation. Its spectral width (8 mT) and very fast relaxation at 15 K indicates that this radical is magnetically coupled to the Mn(2)(IV,IV) center.  相似文献   

2.
Aluminum ions are highly soluble in acidic environments. Toxicity of aluminum ions for heterotrophic, facultatively and obligately chemolithoautotrophic acidophilic bacteria was examined. Acidiphilium cryptum grew in glucose-mineral medium, pH 3, containing 300 mM aluminum sulfate [Al(2)(SO(4))(3)] after a lag phase of about 120 h with a doubling time of 7.6 h, as compared to 5.2 h of growth without aluminum. Precultivation with 1 mM Al(2)(SO(4))(3) and transfer to a medium with 300 mM Al(2)(SO(4))(3) reduced the lag phase from 120 to 60 h, and immediate growth was observed when A. cryptum was precultivated with 50 mM Al(2)(SO(4))(3), suggesting an aluminum-induced resistance. Aluminum resistance was not induced by Fe(3+) ions and divalent cations. Upon exposure of A. cryptum to 300 mM Al(2)(SO(4))(3), the protein profile changed significantly as determined by SDS-PAGE. When other acidophiles were cultivated with 50-200 mM aluminum sulfate, no lag phase was observed while the growth rates and the cellular yields were significantly reduced. This growth response was observed with Acidobacterium capsulatum, Acidiphilium acidophilum, Acidithiobacillus ferrooxidans, and Acidithiobacillus thiooxidans. Precultivation of these strains with aluminum ions did not alter the growth response caused by aluminum. The content of A. cryptum cultivated with 300 mM Al(2)(SO(4))(3)was 0.44 microg Al/mg cell dry weight, while that of the other strains cultivated with 50 mM Al(2)(SO(4))(3) ranged from 0.30 to 3.47 microg Al/mg cell dry weight.  相似文献   

3.
The new homodinuclear complexes, [Cu(2)(II)(HLdtb)(mu-OCH(3))](ClO(4))(2) (1) and [Cu(2)(II)(Ldtb)(mu-OCH(3))](BPh(4)) (2), with the unsymmetrical N(5)O(2) donor ligand (H(2)Ldtb) - {2-[N,N-Bis(2-pyridylmethyl)aminomethyl]-6-[N',N'-(3,5-di-tert-butylbenzyl-2-hydroxy)(2-pyridylmethyl)]aminomethyl}-4-methylphenol have been synthesized and characterized in the solid state by X-ray crystallography.In both cases the structure reveals that the complexes have a common {Cu(II)(mu-phenoxo)(mu-OCH(3))Cu(II)} structural unit.Magnetic susceptibility studies of 1 and 2 reveal J values of -38.3 cm(-1) and -2.02 cm(-1), respectively, and that the degree of antiferromagnetic coupling is strongly dependent on the coordination geometries of the copper centers within the dinuclear {Cu(II)(mu-OCH(3))(mu-phenolate)Cu(II)} structural unit.Solution studies in dichloromethane, using UV-Visible spectroscopy and electrochemistry, indicate that under these experimental conditions the first coordination spheres of the Cu(II) centers are maintained as observed in the solid state structures, and that both forms can be brought into equilibrium ([Cu(2)(HLdtb)(mu-OCH(3))](2+)=[Cu(2)(Ldtb)(mu-OCH(3))](+)+H(+)) by adjusting the pH with Et(3)N (Ldtb(2-) is the deprotonated form of the ligand).On the other hand, potentiometric titration studies of 1 in an ethanol/water mixture (70:30 V/V; I=0.1M KCl) show three titrable protons, indicating the dissociation of the bridging CH(3)O(-) group.The catecholase activity of 1 and 2 in methanol/water buffer (30:1 V/V) demonstrates that the deprotonated form is the active species in the oxidation of 3,5-di-tert-butylcatechol and that the reaction follows Michaelis-Menten behavior with k(cat)=5.33 x 10(-3)s(-1) and K(M)=3.96 x 10(-3)M. Interestingly, 2 can be electrochemically oxidized with E(1/2)=0.27 V vs.Fc(+)/Fc (Fc(+)/Fc is the redox pair ferrocinium/ferrocene), a redox potential which is believed to be related to the formation of a phenoxyl radical.Since these complexes are redox active species, we analyzed their activity toward the nucleic acid DNA, a macromolecule prone to oxidative damage.Interestingly these complexes promoted DNA cleavage following an oxygen dependent pathway.  相似文献   

4.
A range of [PtR(2)(chxn)] (R=C(6)F(5), o-HC(6)F(4), p-HC(6)F(4), p-MeOC(6)F(4) or 3,5-H(2)C(6)F(3); chxn=cyclohexane-1,2-diamine) and cis-[PtR(2)(dmso)(2)] (R=C(6)F(5), p-HC(6)F(4) or p-MeOC(6)F(4); dmso=dimethyl sulfoxide) complexes have been prepared from the corresponding [PtR(2)(diene)] (diene=cis,cis-cycloocta-1,5-diene (cod), hexa-1,5-diene (hex), norbornadiene (nbd) or dicyclopentadiene (dcy)) derivatives and have been spectroscopically characterized. A representative crystal structure of [Pt(C(6)F(5))(2)(cis-chxn)] was determined and shows a slightly distorted square planar geometry for platinum with chxn virtually perpendicular to the coordination plane. The biological activity against L1210 and L1210/DDP cell lines of these compounds together with the behaviour of other organoplatinum complexes, [PtR(2)L(2)] (L(2)=ethane-1,2-diamine (en) or cis-(NH(3))(2)) have been determined. Despite the use of relatively inert fluorocarbon anions as leaving groups, moderate-high cell growth inhibitory activity is observed. None of the fluorocarbon complexes displayed any cross resistance with cisplatin.  相似文献   

5.
Three new complexes [Cu(L)(2)(NO(3))](NO(3))(H(2)O)(1/2)(CH(3)OH)(1/2) (1), [Cd(L)(2)(NO(3))(2)](H(2)O)(3) (2) and [Cd(L)(2)(ClO(4))(CH(3)OH)](ClO(4))(H(2)O)(1/4)(CH(3)OH) (3) (L=1-[3-(2-pyridyl)pyrazol-1-ylmethyl]naphthalene) were synthesized and characterized by elemental analyses, IR and X-ray diffraction analysis. Among them, the Cu(II) and Cd(II) ions were both coordinated by four N donors from two distinct L ligands via N,N-bidentate chelating coordination mode. Additional weak interactions, such as the face-to-face pi-pi stacking and C-Hcdots, three dots, centeredO H-bonding interactions, linked the mononuclear unit into 1D chain and further into 2D network. Complexes 1-3 were subjected to biological assays in vitro against six different cancer cell lines. All of them exhibited cytotoxic specificity and notable cancer cell inhibitory rate. The interactions of 1-3 with calf thymus DNA were investigated by thermal denaturation, viscosity measurements, spectrophotometric and electrophoresis methods. The results indicate that these complexes bound to DNA by intercalation mode via the ligand L and had different nuclease activities, which were in good agreement with their DNA-binding strength. Moreover, the central metal ions of 1-3 played a vital role in DNA-binding behaviors, DNA-cleavage activities and cytotoxicities, whereas the contribution of the different counter anions to their bioactivities also should not be ignored.  相似文献   

6.
The synthesis and crystallographic characterization of a series of diiron(II) complexes with sterically hindered terphenyl carboxylate ligands and alkyl amine donors are presented. The compounds [Fe(2)(mu-O(2)CAr(Tol))(4)(L)(2)] (L=NH(2)(CH(2))(2)SBn (1); NH(2)(CH(2))(3)SMe (2); NH(2)(CH(2))(3)CCH (3)), where (-)O(2)CAr(Tol) is 2,6-di(p-tolyl)benzoate, and [Fe(2)(mu-O(2)CAr(Xyl))(2)(O(2)CAr(Xyl))(2)(L)(2)] (L=NH(2)(CH(2))(3)SMe (4); NH(2)(CH(2))(3)CCH (5)), where (-)O(2)CAr(Xyl) is 2,6-di(3,5-dimethylphenyl)benzoate, were prepared as small molecule mimics of the catalytic sites of carboxylate-bridged non-heme diiron enzymes. The compounds with the (-)O(2)CAr(Tol) carboxylate form tetrabridged structures, but those containing the more sterically demanding (-)O(2)CAr(Xyl) ligand have only two bridging ligands. The ancillary nitrogen ligands in these carboxylate-rich complexes incorporate potential substrates for the reactive metal centers. Their oxygenation chemistry was studied by product analysis of the organic fragments following decomposition. Compound 1 reacts with dioxygen to afford PhCHO in approximately 30% yield, attributed to oxidative dealkylation of the pendant benzyl group. Compound 3 decomposes to form Fe(II)Fe(III) and Fe(III)Fe(IV) mixed-valence species by established bimolecular pathways upon exposure to dioxygen at low temperatures. Upon decomposition, the alkyne-substituted amine ligand was recovered quantitatively. When the (-)O(2)CAr(Tol) carboxylate was replaced by the (-)O(2)CAr(Xyl) ligand in 5, different behavior was observed. The six-coordinate iron(III) complex with one bidentate and two monodentate carboxylate ligands, [Fe(O(2)CAr(Xyl))(3)(NH(2)(CH(2))(3)CCH)(2)] (6), was isolated from the reaction mixture following oxidation.  相似文献   

7.
Nucleoside phosphorylation by phosphate minerals   总被引:1,自引:0,他引:1  
In the presence of formamide, crystal phosphate minerals may act as phosphate donors to nucleosides, yielding both 5'- and, to a lesser extent, 3'-phosphorylated forms. With the mineral Libethenite the formation of 5'-AMP can be as high as 6% of the adenosine input and last for at least 10(3) h. At high concentrations, soluble non-mineral phosphate donors (KH(2)PO(4) or 5'-CMP) afford 2'- and 2':3'-cyclic AMP in addition to 5'-and 3'-AMP. The phosphate minerals analyzed were Herderite Ca[BePO(4)F], Hureaulite Mn(2+)(5)(PO(3)(OH)(2)(PO(4))(2)(H(2)O)(4), Libethenite Cu(2+)(2)(PO(4))(OH), Pyromorphite Pb(5)(PO(4))(3)Cl, Turquoise Cu(2+)Al(6)(PO(4))(4)(OH)(8)(H(2)O)(4), Fluorapatite Ca(5)(PO(4))(3)F, Hydroxylapatite Ca(5)(PO(4))(3)OH, Vivianite Fe(2+)(3)(PO(4))(2)(H(2)O)(8), Cornetite Cu(2+)(3)(PO(4))(OH)(3), Pseudomalachite Cu(2+)(5)(PO(4))(2)(OH)(4), Reichenbachite Cu(2+)(5)(PO(4))(2)(OH)(4), and Ludjibaite Cu(2+)(5)(PO(4))(2)(OH)(4)). Based on their behavior in the formamide-driven nucleoside phosphorylation reaction, these minerals can be characterized as: 1) inactive, 2) low level phosphorylating agents, or 3) active phosphorylating agents. Instances were detected (Libethenite and Hydroxylapatite) in which phosphorylation occurs on the mineral surface, followed by release of the phosphorylated compounds. Libethenite and Cornetite markedly protect the beta-glycosidic bond. Thus, activated nucleic monomers can form in a liquid non-aqueous environment in conditions compatible with the thermodynamics of polymerization, providing a solution to the standard-state Gibbs free energy change (DeltaG degrees ') problem, the major obstacle for polymerizations in the liquid phase in plausible prebiotic scenarios.  相似文献   

8.
The aim of this work is to study the binding of nickel ions to hexahistidine (His(6)) combining potentiometric titrations and spectroscopic (UV-Vis and circular dichroism) determinations in order to establish the species distribution as a function of the pH, their stoichiometry, stability and geometry. For comparative purposes, the same procedure was applied to the Ni-histidine (His) system. His behaves as a tridentate ligand, coordinating the carboxyl group, the imidazole and the amino nitrogen atoms to Ni(II) ions in an octahedral coordination and a bis(histidine) complex is formed at pH higher than 5. For the Ni-His(6) system, the complex formation starts at pH 4 and five different species (Ni(His(6))H, Ni(His(6)), Ni(n)(His(6))(n), Ni(n)(His(6))(n)H(-n/2), Ni(n)(His(6))(n)H(-n)) are formed as a function of the pH. Ni(His(6))H involves the coordination of the imidazole nitrogen and a deprotonated amide nitrogen (N(Im), N(-)) resulting in an octahedral geometry. In Ni(His(6)), an imidazole nitrogen is deprotonated and coordinated (2N(Im), N(-)) to the metal ion with a square planar geometry. The aggregated forms result from the extra Ni-N(Im) coordination, resulting in a 4N square planar geometry that is stabilized by inter/intramolecular hydrogen bonds. This coordination mode is not altered during the deprotonation steps from Ni(n)(His(6))(n).  相似文献   

9.
Four imidazolate-bridged binuclear copper(II)-copper(II) and copper(II)-zinc(II) complexes viz., [(Bipy)(2)Cu-Im-Cu(Bipy)(2)](ClO(4))(3).CH(3)OH, [(Phen)(2)Cu-Im-Cu(Phen)(2)](BF(4))(3).2CH(3)OH, [(Bipy)(2)Cu-Im-Zn(Bipy)(2)](BF(4))(3), and [(Phen)(2)Cu-Im-Zn(Phen)(2)](BF(4))(3), (Bipy=2,2'-Bipyridyl, Phen=1-10-Phenanthroline and Im=imidazolate ion) were synthesized as a possible models for superoxide dismutase (SOD). Complex [(Bipy)(2)Cu-Im-Cu(Bipy)(2)](ClO(4))(3).CH(3)OH has been structurally characterized. This complex crystallizes in the triclinic space group P1, with the unit parameters a=8.88(5) A, b=13.79(17) A, c=20.18(18) A, alpha=76.424(8)(o), beta=85.888(6)(o), gamma=82.213(7). The metal-nitrogen bond length from 1.972-2.273 A and the distance Cu-Cu is 5.92 A. The five-coordinate geometry about the copper(II) ion is square pyramidal. Magnetic moment and electron paramagnetic resonance (e.p.r.) spectral measurements of the homobinuclear complexes have shown an antiferromagnetic exchange interaction. From the e.p.r. and UV-Vis spectral measurement studies, these complexes have been found to be stable (pH 8.5-10.5 for 1, 10.5 for 2,3 and 8.5 for 4). These complexes catalyse the dismutation of superoxide radical (O(2)(-)) at biological pH. All the observations indicate that these complexes act as good possible models for superoxide dismutase.  相似文献   

10.
Reaction of Mn(II) with phenoxyalkanoic acids and di-2-pyridyl ketone oxime (Hpko) leads to neutral tetranuclear complexes of the general formula Mn(4)(O)(pko)(4)(phenoxyalkanoato)(4) (phenoxyalkanoic acids: H-mcpa=2-methyl-4-chloro-phenoxy-acetic acid, H-2,4,5-T=2,4,5-trichloro-phenoxy-acetic acid or H3,4-D=3,4-dichloro-phenoxy-acetic acid). The compounds were synthesized by adding di-2-pyridyl ketone oxime to MnCl(2) in the presence of the sodium salts of the alkanoic acids in methanol. The crystal structure of Mn(4)(II/II/II/IV)(O)(pko)(4)(2,4,5-T)(4).2.5CH(3)OH.0.25H(2)O 1 shows that the complex consists of a [Mn(4)(mu(4)-O)](8+) core with a Mn(IV) and 3 Mn(II) ions in octahedral environment and a mu(4)-O atom bridging the four manganese ions. Spectroscopic studies of the interaction of these tetranuclear clusters with DNA showed that these compounds bind to dsDNA. The binding strength of the Mn(4)(II/II/II/IV)(O)(pko)(4)(2,4,5-T)(4) complex for calf thymus DNA is equal to 1.1x10(4)M(-1). Among the deoxyribonucleotides they bind preferentially to deoxyguanylic acid (dGMP). Competitive studies with ethidium bromide (EthBr) showed that the Mn(4)(II/II/II/IV)(O)(pko)(4)(2,4,5-T)(4) complex exhibited the ability to displace the DNA-bound EthBr indicating that the complex binds to DNA via intercalation in strong competition with EthBr for the intercalative binding site. Additionally, DNA electrophoretic mobility experiments showed that all three complexes, at low cluster concentration, are obviously capable of binding to pDNA causing its cleavage (relaxation) at physiological pH and temperature. At higher cluster concentration, catenated dimer forms of pDNA was formed.  相似文献   

11.
The compound [Cu(2)(II)(D(1))(H(2)O)(2)](ClO(4))(4).2H(2)O [D(1)=binucleating ligand with tris(2-pyridylmethyl)amine (TMPA) moieties linked in the 5-pyridyl position by a -CH(2)CH(2)- bridge] mediated efficient oxidative cleavage of pBR322 plasmid DNA under reducing conditions. A mononuclear analogue, [Cu(TMPA)(H(2)O)](ClO(4))(2), was less effective at linearizing supercoiled (Form I) plasmid DNA as compared to the binuclear complex. A new method for quenching the copper-dependent reactions has been developed to avoid plasmid scission by the binuclear complex and the standard gel loading buffer. EDTA was not sufficient for retarding copper reaction, but diethyldithiocarbamic acid was capable of inhibiting all reactivity. Investigation of oxidative cleavage of double-helical oligonucleotides by [Cu(2)(II)(D(1))(H(2)O)(2)](ClO(4))(4) confirmed the enhanced reactivity of the binuclear over the mononuclear complex and provided mechanistic insights into the nature of the reaction. Cleavage of DNA required both the binuclear complex and a reductant and likely proceeded through an O(2)-derived intermediate that does not include a diffusible hydroxyl radical. The greater efficiency of the binuclear complex relative to the mononuclear analogue is consistent with their relative abilities to activate dioxygen.  相似文献   

12.
Ce(4+) (Ce(NH(4))(2)(NO(3))(6)) at 1mM induces apoptosis of suspension cultures of Taxus cuspidata cells; however, the underlying signal mechanisms are unknown. We show here that a 46-kDa ERK (extracellular signal-regulated kinase)-like MAPK appears to be down-regulated at 4h, and remains at low levels for up to 48 h. An inhibitor of superoxide anions (O(2)(-)) generation, diphenyl iodonium (DPI) successfully blocks down-regulation of ERK-like MAPK and degradation of DNA. Moreover, a 41-kDa p38-like MAPK activity remains unchanged from 0.5 to 48 h. The p38 inhibitor SB202190 effectively inhibits p38-like MAPK activity, however, SB202190 fails to modify the apoptotic rate at concentrations up to 100 microM. Three nuclease (34-kDa, 22-kDa and 20-kDa) activities are profoundly enhanced in Ce(4+)-induced T. cuspidata cells. They have an optimum pH at 6.8, and are stimulated by Ca(2+)/Mg(2+). Caspase-3 inhibitor, Ac-DEVD-CHO, does not attenuate the 34-kDa nuclease activity, but inhibits the 22-kDa and the 20-kDa nuclease activities. In addition, inhibition of O(2)(-) generation by DPI significantly reduces the three nuclease activities. In conclusion, the present study suggests that down-regulation of ERK-like MAPK, burst of O(2)(-), activation of caspase-3-like and induction of three nucleases as the key signaling events mediating apoptosis in Ce(4+)-induced cultured T. cuspidata cells.  相似文献   

13.
Horton TE  DeRose VJ 《Biochemistry》2000,39(37):11408-11416
The effects of Co(NH(3))(6)(3+) on the hammerhead ribozyme are analyzed using several techniques, including activity measurements, electron paramagnetic resonance (EPR), and circular dichroism (CD) spectroscopies and thermal denaturation studies. Co(NH(3))(6)(3+) efficiently displaces Mn(2+) bound to the ribozyme with an apparent dissociation constant of K(d app) = 22 +/- 4.2 microM in 500 microM Mn(2+) (0.1 M NaCl). Displacement of Mn(2+) coincides with Co(NH(3))(6)(3+) inhibition of hammerhead activity in 500 microM Mn(2+), reducing the activity of the WT hammerhead by approximately 15-fold with an inhibition constant of K(i) = 30.9 +/- 2.3 microM. A residual 'slow' activity is observed in the presence of Co(NH(3))(6)(3+) and low concentrations of Mn(2+). Under these conditions, a single Mn(2+) ion remains bound and has a low-temperature EPR spectrum identical to that observed previously for the highest affinity Mn(2+) site in the hammerhead ribozyme in 1 M NaCl, tentatively attributed to the A9/G10.1 site [Morrissey, S. R. , Horton, T. E., and DeRose, V. J. (2000) J. Am. Chem. Soc. 122, 3473-3481]. Circular dichroism and thermal denaturation experiments also reveal structural effects that accompany the observed inhibition of cleavage and Mn(2+) displacement induced by addition of Co(NH(3))(6)(3+). Taken together, the data indicate that a high-affinity Co(NH(3))(6)(3+) site is responsible for significant inhibition accompanied by structural changes in the hammerhead ribozyme. In addition, the results support a model in which at least two types of metal sites, one of which requires inner-sphere coordination, support hammerhead activity.  相似文献   

14.
Glasner ME  Bergman NH  Bartel DP 《Biochemistry》2002,41(25):8103-8112
The class I ligase, a ribozyme previously isolated from random sequence, catalyzes a reaction similar to RNA polymerization, positioning its 5'-nucleotide via a Watson-Crick base pair, forming a 3',5'-phosphodiester bond between its 5'-nucleotide and the substrate, and releasing pyrophosphate. Like most ribozymes, it requires metal ions for structure and catalysis. Here, we report the ionic requirements of this self-ligating ribozyme. The ligase requires at least five Mg(2+) for activity and has a [Mg(2+)](1/2) of 70-100 mM. It has an unusual specificity for Mg(2+); there is only marginal activity in Mn(2+) and no detectable activity in Ca(2+), Sr(2+), Ba(2+), Zn(2+), Co(2+), Cd(2+), Pb(2+), Co(NH(3))(6)(3+), or spermine. All tested cations other than Mg(2+), including Mn(2+), inhibit the ribozyme. Hill analysis in the presence of inhibitory cations suggested that Ca(2+) and Co(NH(3))(6)(3+) inhibit by binding at least two sites, but they appear to productively fill a subset of the required sites. Inhibition is not the result of a significant structural change, since the ribozyme assumes a nativelike structure when folded in the presence of Ca(2+) or Co(NH(3))(6)(3+), as observed by hydroxyl-radical mapping. As further support for a nativelike fold in Ca(2+), ribozyme that has been prefolded in Ca(2+) can carry out the self-ligation very quickly upon the addition of Mg(2+). Ligation rates of the prefolded ribozyme were directly measured and proceed at 800 min(-1) at pH 9.0.  相似文献   

15.
5-Fluoroorotic acid (H(3)FOro) is a potent inhibitor for some metalloproteins such as dihydroorotase and dihydroorotate dehydrogenase and for thymidylate synthase (nonmetalloprotein) in the human malaria parasite Plasmodium falciparum. To study the coordination chemistry of H(3)Foro, the ammonium salt [NH(4)(+)][H(2)FOro(-)].1H(2)O (1) and the first coordination compounds of H(3)FOro with transition metals [Ni(HFOro(2-))(H(2)O)(4)].1H(2)O (2), [Cu(HFOro(2-))(NH(3))(H(2)O)](n) (3) and [Cu(3)(FOro(3-))(2)(NH(3))(6)(H(2)O)(2)] (4) have been synthesised and characterised by single-crystal X-ray diffraction, IR spectroscopy and by thermogravimetry. Three different coordination modes of 5-fluoroorotic acid have been established. In all cases the ligand is chelated to the metal via an amido-nitrogen and a carboxylate-oxygen but for (3), there is also a carboxylate oxygen from another HFOro(2-) ligand resulting in a polymeric structure and for (4), the second amido-nitrogen in the ororotic acid ring coordinates to give a trinuclear complex. The metal coordination polyhedra are octahedral in (2), square-pyramidal in (3) and square-planar and approximately square-pyramidal in (4). An octahedral coordination geometry including a N(1)/O(61)-chelating HFOro(2-) ligand with four aqua ligands is proposed for the Zn complex [Zn(HFOro(2-)) (H(2)O)(4)].0.5H(2)O (5), based on IR and thermogravimetric data. Extensive hydrogen bonded networks and some ring-ring stacking interactions are observed in each of the structures.  相似文献   

16.
Cu in blood has been believed to transport into cell via albumin and some amino acids. To shed light on the Cu transport process we studied the reaction of the Cu(II)-peptide with the amino acid by absorption and CD spectra. Albumin mimic peptides GlyGly-L-HisGly (GGHG) and penta-Gly(G5) formed stable 4N coordinated Cu(II) complexes, but in the reaction with histidine (His) and penicillamine (Pes) the ternary Cu(II) complex formations were observed different by the kinetic study. Cu(II)-G5 complexes reacted with Pes to form the ternary complex Cu(H(-1)G5)(Pes(-)) which was subsequently transformed to the binary complex Cu(Pes(-))(2). In the system with GGHG the Cu(II) was also transported from GGHG to Pes, but the ternary Cu(H(-1)GGHG)(Pes(-)) complex as the intermediate was detected a trace. The ternary complex would be spontaneously transformed to Cu(Pes(-))(2) upon forming, because the rate constant of the ternary complex formation k(1+)= approximately 2M(-1)s(-1) was less than k(2+)= approximately 5 x 10(2)M(-1)s(-1) for the Cu(Pes(-))(2) formation at physiological pH. In the Cu(II)-GGHG-His system the ternary Cu(H(-1)GGHG)(His) complex was also hardly identified because the formation constant K(1) and k(1+) were very small and the equilibrium existed between Cu(H(-2)GGHG) and Cu(His)(2) and its overall equilibrium constant beta(2) for Cu(His)(2) was very small to be 1.00+/-0.05 M(-1) at pH 9.0. These results indicated that the ternary complex is formed in the Cu transport process from the albumin to the amino acid, but His imidazole nitrogen in the fourth-binding site of Cu(II) strongly resists the replacement by the incoming ligand.  相似文献   

17.
The effects of varying extracellular pH on the rates of uptake of titratable anions by human erythrocytes under conditions of constant intracellular pH have been determined for a series of highly related anions, the phosphate "analogs." These compounds are simply substituted phosphorus oxyacids, differing in the number and acidity of titratable protons: phosphate (HPO4(2-), pKa 6.8); phosphite (HPO3(2-), pKa 6.4); hypophosphite (H2PO2-); methylphosphonate ((CH3)PO3(2-), pKa 7.4); dimethylphosphinate ((CH3)2PO2-); fluorophosphate [PO3F2-, pKa 4.7); and thiophosphate (HSPO3(2-), pKa 5.5). Suspensions of intact, Cl(-)-loaded erythrocytes (intracellular pH, 7.2) were incubated at 37 degrees C in isotonic buffers (pH 4-8) containing 60 mM phosphate analog for specified time intervals, whereupon influx was halted by the addition of 1 mM 4-acetamido-4'-isothiocyanatostilbene-2,2'-disulfonic acid (SITS), an inhibitor of anion exchange. The intracellular anion concentrations were determined from 31P or 19F nuclear magnetic resonance spectra from the erythrocyte suspensions. The influx rates for the titratable phosphate analogs exhibited bimodal pH dependence, reaching maximal levels at pH values that increased with increasing anion pK. This pH-dependent behavior is consistent with a transport channel that contains a titratable regulatory site which interacts with the translocated anion. Based upon the Henderson-Hasselbalch equation, the probability that a titratable anion will have an electric charge of equal magnitude to that of the titratable carrier is highest at a pH value exactly midway between the pK of the regulatory site and that of the anion. The pH maxima observed for the phosphate analogs indicate a pK for this site of 5.5 at 37 degrees C. Intracellular pH changes associated with influx indicated that transport of the "fast" anion phosphite is largely in monoionized form. Intracellular pH changes associated with transport of slow anions were predominantly determined by partial ionic equilibrium effects and did not indicate the ionization state of the transported anion.  相似文献   

18.
Mononuclear complexes of Cu(II), Ni(II), and Mn(II) with a new Schiff base ligand derived from indoline-2,3-dione and 2-hydroxybenzohydrazide, [Cu(II)(L)(2)], [Ni(II)(L)(2)], and [Mn(II)L.(AcO).2C(2)H(5)OH] [HL=(Z)-2-hydroxy-N'-(2-oxoindolin-3-ylidene)benzohydrazide], have been prepared. The complexes have been structurally characterized by X-ray crystallography. Among the three complexes, the Cu(II) complex had the novel highest antitumor activity.  相似文献   

19.
Twelve zinc(II) complexes with thiosemicarbazone and semicarbazone ligands were prepared and characterized by elemental analysis, thermogravimetric and differential thermal analysis (TG/DTA), FT-IR and 1H and 13C NMR spectroscopy. Seven three-dimensional structures of zinc(II) complexes were determined by single-crystal X-ray analysis. Their antimicrobial activities were evaluated by MIC against four bacteria (B. subtilis, S. aureus, E. coli and P. aeruginosa), two yeasts (C. albicans and S. cerevisiae) and two molds (A. niger and P. citrinum). The 5- and 6-coordinate zinc(II) complexes with a tridentate thiosemicarbazone ligand (Hatsc), ([Zn(atsc)(OAc)](n) 1, [Zn(Hatsc)(2)](NO(3))(2).0.3H(2)O 2, [ZnCl(2)(Hatsc)] 3 and [Zn(SO(4))(Hatsc)(H(2)O)].H(2)O 4 [Hatsc=2-acetylpyridine(thiosemicarbazone)]), showed antimicrobial activities against test organisms, which were different from those of free ligands or the starting zinc(II) compounds. Especially, complex 2 showed effective activities against P. aeruginosa, C. albicans and moderate activities against S. cerevisiae and two molds. These facts are in contrast to the results that the 5- or 6-coordinate zinc(II) complexes with a tridentate 2-acetylpyridine-4N-morpholinethiosemicarbazone, ([Zn(mtsc)(2)].0.2EtOH 5, the previously reported catena-poly [Zn(mtsc)-mu-(OAc-O,O')](n) and [Zn(NO(3))(2)(Hmtsc)] [Hmtsc=2-acetylpyridine (4N-morpholyl thiosemicarbazone)]), showed no activities against the test microorganisms. The 5- and 6-coordinate zinc(II) complexes with a tridentate 2-acetylpyridinesemicarbazone, ([Zn(OAc)(2)(Hasc)] 6 and [Zn(Hasc)(2)](NO(3))(2) 7 [Hasc=2-acetylpyridine(semicarbazone)]), showed no antimicrobial activities against bacteria, yeasts and molds. Complex [ZnCl(2)(Hasc)] 8, which was isostructural to complex 3, showed modest activity against Gram-positive bacterium, B. subtilis. The 1:1 complexes of zinc(II) with pentadentate thiosemicarbazone ligands, ([Zn(dmtsc)](n) 9 and [Zn(datsc)](n) 10 [H(2)dmtsc=2,6-diacetylpyridine bis(4N-morpholyl thiosemicarbazone) and H(2)datsc=2,6-diacetylpyridine bis(thiosemicarbazone)]), did not inhibit the growth of the test organisms. On the contrary, 7-coordinate zinc(II) complexes with one pentadentate semicarbazone ligand and two water molecules, ([Zn(H(2)dasc)(H(2)O)(2)](OAc)(2).5.3H(2)O 11 and [Zn(H(2)dasc)(H(2)O)(2)](NO(3))(2).H(2)O 12 [H(2)dasc=2,6-diacetylpyridine bis(semicarbazone)]), showed modest to moderate activities against bacteria. Based on the X-ray structures, the structure-activity correlation for the antimicrobial activities was elucidated. The zinc(II) complexes with 4N-substituted ligands showed no antimicrobial activities. In contrast to the previously reported nickel(II) complexes, properties of the ligands such as the ability to form hydrogen bonding with a counter anion or hydrated water molecules or the less bulkiness of the 4N moiety would be a more important factor for antimicrobial activities than the coordination number of the metal ion for the zinc(II) complexes.  相似文献   

20.
Anion inhibition of the proton pump in rat liver multivesicular bodies   总被引:3,自引:0,他引:3  
Rat liver multivesicular bodies (MVB), as well as other hepatic subcellular organelles, are acidified by an electrogenic ATP-dependent proton pump that requires Cl- for maximal acidification (Van Dyke, R. W., Hornick, C. A., Belcher, J., Scharschmidt, B. F., and Havel, R.J. (1985) J. Biol. Chem. 260, 11021-11026), suggesting that Cl- serves as a permeable charge-compensating anion. However, we have observed that NO3- is unable to substitute for Cl-. This study was undertaken therefore to examine more closely the effects of Cl- on MVB acidification and to determine whether NO3- and other anions interact with the proton pump. ATP-dependent vesicle acidification and membrane potential (psi) were measured using the fluorescent dyes acridine orange and Oxonol V (bis(3-phenyl-5-oxoisoxasol-4-yl)pentamethine oxonol), respectively. Cl- both stimulated acidification (Km = 23.2 +/- 4.2 mM) and decreased psi (IC50 = 3.4 +/- 0.6 mM) in a concentration-dependent, nonlinear fashion. In the presence of saturating Cl- (100 mM), however, NO3- (shown to be more permeable than Cl-) and the impermeant anions SO4(2-) and PO4(2-), inhibited both ATP-dependent acidification and psi in a concentration-dependent manner. Other anions, including gluconate and HCO3-, had no effect. The inhibitory effect of NO3- was reversible. Neither SO4(2-) nor PO4(2-) appeared to block Cl- movement across the vesicle membrane as assessed by the ability of Cl- to decrease an established psi. In additional experiments, the effects of anions on relaxation of a previously established pH gradient were measured. Compared to Cl- or gluconate, NO3- had no significant effect on pH gradient relaxation, even when MVB were preloaded with NO3-, indicating that rapid cycling of NO3-/HNO3 across the MVB membrane does not occur. The organic nitrate, isosorbide dinitrate, also inhibited both acidification and psi and, similar to NO3-, had no effect on pH gradient relaxation. By contrast, NO2- potently inhibited both MVB acidification and psi but also rapidly relaxed a pre-established pH gradient, suggesting that NO2- increases MVB membrane proton permeability. Finally, MVB exhibited N-ethylmaleimide-sensitive ATPase activity that was inhibited 23.9% by NO3- (100 mM). In conclusion, although MVB are permeable to a variety of anions (Cl-, Br-, NO3-, NO2-), only Cl- and Br- support maximal rates of acidification by the proton pump.(ABSTRACT TRUNCATED AT 400 WORDS)  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