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 9 毫秒
1.
A dinuclear Zn(II) complex with hexaaza macrocyclic ligand bearing two 2-hydroxypropyl pendants, 3,6,9,16,19,22-hexaaza-6,19-bis(2-hydroxypropyl)-tricyclo [22,2,2,2(11,14)]triaconta-11,13,24,26,27,29-hexane (L) was synthesized and studied as a catalyst of the cleavage of 4-nitrophenyl acetate (NA). X-ray diffraction analysis of [Zn(2)LCl(2)]Cl(2)(.)6H(2)O revealed that Zn(II) adopts a trigonal-bipyramidal geometry. The complexation constants of L with Zn(II) have been determined at 298 K by means of potentiometric titration. [Zn(2)H(-2)L](2+) is the dominant species in aqueous solution around pH 8. The Zn(2)L-promoted hydrolysis of NA showed a second-order rate constant of 0.33 M(-1)s(-1) at pH 9.0, and the main promoter species are concluded to be the deprotonated species [Zn(2)H(-2)L](2+).  相似文献   

2.
A new copper(II) complex of santonic acid [Cu(2)(sant)(4)(H(2)O)(2)].2(1/2)H(2)O has been prepared and characterized by electronic, vibrational, EPR spectral studies, and stability determinations in solution. The presence of two antiferrromagnetically coupled copper centers in the solid state was detected by EPR. The dinuclear Cu(II) complex crystallizes in the tetragonal P4(3)2(1)2 space group, with a=b=14.498(3), c=64.07(1)A. Biological studies indicate that the complex displays interesting potential antitumoral actions.  相似文献   

3.
Polarographic and UV-spectrophotometric investigations of Pb(II) complex formation with beta-cyclodextrin have showed that the complexation of Pb(II) ions begins at pH >10. The formation of lead(II) 1:1 complex with the beta-cyclodextrin anion was observed at pH 10-11.5. The logarithm of the stability constant of this complex compound is 15.9+/-0.3 (20 degrees C, ionic strength 1.0), and the molar extinction coefficient value is ca. 5500 (lambda(max)=260 nm). With further increase in solution pH the Pb-beta-cyclodextrin complex decomposes and converts to Pb(OH)(2) or Pb(OH)(3)(-) hydroxy-complexes. This process occurs with a decrease in Pb(II) complexation degree. The latter result could be explained by a decrease in the beta-cyclodextrin anion activity. Neither Pb(OH)(2) nor Pb(OH)(3)(-) encapsulation into beta-CD cavity was observed.  相似文献   

4.
The kinetics of the decomposition of H(2)O(2) catalyzed by Cu(II) has been studied by the initial-rate method in aqueous phosphate media at near physiological pH. The activity of the catalyst is increased by [Fe(CN)(6)](3-) and decreased by VO(3)(-), CrO(4)(2-) and Zn(II). Three reaction pathways are involved in the Cu(II)-H(2)O(2) reaction, the kinetic orders of the catalyst being 1 (rate constant k1), 2 (rate constant k2) and 3 (rate constant k3). The three pathways present fractional apparent orders (>1) in H(2)O(2) and base catalysis. The apparent activation energies associated to rate constants k1, k2 and k3 are 102+/-4, 65+/-8 and 61+/-5 kJ mol(-1). Free-radical chain mechanisms are proposed for the three pathways.  相似文献   

5.
Co(II) interacts with bleomycin in aqueous solution, in the presence of air, to give a short-lived mononuclear superoxo Co(III) complex (I). Then, two molecules of complex I react together, with the loss of oxygen, to yield the dinuclear mu-peroxo Co(III) complex (II); the dimerization follows a second-order rate law with k2 = 200 +/- 50 M-1 s-1 at 25 degrees C. The rate of dimerization is lowered by a factor of 2000 when DNA is present at a molar ratio of [nucleotide]/[Co] higher than 16. These results and studies of circular dichroism and electron paramagnetic resonance spectra of complexes strongly suggest the binding of the superoxo complex to DNA (I') as well as that of the mu-peroxo complex (II'); the binding of 1 molecule of complex II for every 2.9 base pairs in DNA has been determined with an apparent equilibrium constant of 8.4 x 10(4) M-1.  相似文献   

6.
Mechanism of reaction of myeloperoxidase with nitrite   总被引:10,自引:0,他引:10  
Myeloperoxidase (MPO) is a major neutrophil protein and may be involved in the nitration of tyrosine residues observed in a wide range of inflammatory diseases that involve neutrophils and macrophage activation. In order to clarify if nitrite could be a physiological substrate of myeloperoxidase, we investigated the reactions of the ferric enzyme and its redox intermediates, compound I and compound II, with nitrite under pre-steady state conditions by using sequential mixing stopped-flow analysis in the pH range 4-8. At 15 degrees C the rate of formation of the low spin MPO-nitrite complex is (2.5 +/- 0.2) x 10(4) m(-1) s(-1) at pH 7 and (2.2 +/- 0.7) x 10(6) m(-1) s(-1) at pH 5. The dissociation constant of nitrite bound to the native enzyme is 2.3 +/- 0.1 mm at pH 7 and 31.3 +/- 0.5 micrometer at pH 5. Nitrite is oxidized by two one-electron steps in the MPO peroxidase cycle. The second-order rate constant of reduction of compound I to compound II at 15 degrees C is (2.0 +/- 0.2) x 10(6) m(-1) s(-1) at pH 7 and (1.1 +/- 0.2) x 10(7) m(-1) s(-1) at pH 5. The rate constant of reduction of compound II to the ferric native enzyme at 15 degrees C is (5.5 +/- 0.1) x 10(2) m(-1) s(-1) at pH 7 and (8.9 +/- 1.6) x 10(4) m(-1) s(-1) at pH 5. pH dependence studies suggest that both complex formation between the ferric enzyme and nitrite and nitrite oxidation by compounds I and II are controlled by a residue with a pK(a) of (4.3 +/- 0.3). Protonation of this group (which is most likely the distal histidine) is necessary for optimum nitrite binding and oxidation.  相似文献   

7.
Sheng X  Guo X  Lu XM  Lu GY  Shao Y  Liu F  Xu Q 《Bioconjugate chemistry》2008,19(2):490-498
A preorganized cleft dinuclear zinc(II) complex of 2,6-bis(1-methyl-1,4,7-triazacyclonon-1-yl)pyridine 1 as an artificial nuclease was prepared via an improved method. The interactions of 1, 2 [1,4,7-triazacyclononane (TACN)], and their zinc(II) complexes with calf thymus DNA were studied by spectroscopic techniques, including fluorescence and CD spectroscopy. The results indicate that the DNA binding affinities of these compounds are in the following order: Zn(II)2 -1 > Zn(II) -2 > 1 > 2. The binding constants of the Zn (II)2 -1 and Zn(II)-2 complexes are 3.57 x 10(6) and 1.43 x 10(5) M(-1), respectively. Agarose gel electrophoresis was used to assess the plasmid pUC 19 DNA cleavage activities in the presence of the dinuclear Zn (II)2 -1 complex, which exhibits powerful DNA cleavage efficiency. Kinetic data for DNA cleavage promoted by the Zn(II)2 -1 complex under physiological conditions give the observed rate constant ( k obs) of 0.136 h(-1), which shows an 10(7)-fold rate acceleration over uncatalyzed supercoiled DNA. The comparison of the dinuclear Zn(II)2 -1 complex with the mononuclear zinc(II) complex of 1,4,7-triazacyclononane indicates that the DNA cleavage acceleration promoted by the Zn(II)2 -1 complex is due to the efficient cooperative catalysis of the two proximate zinc(II) cation centers. A hydrolytic mechanism of the cleavage process was suggested, and a preliminary study of the antitumor activity was also conducted.  相似文献   

8.
The previously synthesized, terminally blocked heptapeptide Ac-Aib-ATANP-Aib-Aib-ATANP-Aib-Aib-OMe (1a), where ATANP is (S)-2-amino-3-[1-(1,4,7-triazacyclononane)]propanoic acid and Aib is alpha-aminoisobutyric acid, which is soluble in neutral water where it largely adopts a 3(10)-helical conformation, has been studied, as bimetallic complex [metal ions: Cu(II), Ni(II), Zn(II)], for the transphosphorylation catalysis of the RNA-model substrate 2-(hydroxypropyl)-p-nitrophenyl phosphate (HPNP). A detailed analysis was carried out with the Zn(II) dinuclear complex. Comparison with the mononuclear Zn(II) complex with 1,4,7-triazacyclononane (3) points to cooperativity between the two Zn(II) ions in the process catalyzed by 1a-2Zn(II). On the contrary, the dinuclear Zn(II) complex of dipeptide Ac-(ATANP)(2)-OMe (2), lacking any ordered conformation, is less active than 3-Zn(II). The kinetic analysis suggests the following: (a) the peptide is conformationally very robust and does not loose activity up to 50 degrees C; (b) the substrate binds to the peptide-Zn(II) complex, although not all modes of complexation allow us to take advantage of the cooperativity between the two metal centers. The maximum rate acceleration estimated at pH 7 for the fully bound substrate is ca. 200-fold compared with the uncatalyzed process.  相似文献   

9.
The antitumor antibiotic Altromycin H was studied using electronic absorption (UV-Vis.) and circular dichroism (CD) spectroscopy. The dissociation constants of the phenolic groups on C(5) and C(11) were estimated as pK(1)=6.7 and pK(2)=11.8 at 25 degrees C, respectively, and a complete assignment of the CD and UV-Vis. bands is proposed. The interaction of Cu(II) ions with the Altromycin H has been also investigated by UV-Vis., CD and electron paramagnetic resonance (EPR) spectroscopy. A pH depended stepwise complex formation was observed. At pH<4 no copper-Altromycin H interactions were detected. At the 4相似文献   

10.
The [Ru(II)(Hedta)NO(+)] complex is a diamagnetic species crystallizing in a distorted octahedral geometry, with the Ru-N(O) length 1.756(4) A and the RuNO angle 172.3(4) degrees . The complex contains one protonated carboxylate (pK(a)=2.7+/-0.1). The [Ru(II)(Hedta)NO(+)] complex undergoes a nitrosyl-centered one-electron reduction (chemical or electrochemical), with E(NO+/NO)=-0.31 V vs SCE (I=0.2 M, pH 1), yielding [Ru(II)(Hedta)NO](-), which aquates slowly: k(-NO)=2.1+/-0.4x10(-3) s(-1) (pH 1.0, I=0.2 M, CF(3)COOH/NaCF(3)COO, 25 degrees C). At pHs>12, the predominant species, [Ru(II)(edta)NO](-), reacts according to [Ru(II)(edta)NO](-)+2OH(-)-->[Ru(II)(edta)NO(2)](3-), with K(eq)=1.0+/-0.4 x 10(3) M(-2) (I=1.0 M, NaCl; T=25.0+/-0.1 degrees C). The rate-law is first order in each of the reactants for most reaction conditions, with k(OH(-))=4.35+/-0.02 M(-1)s(-1) (25.0 degrees C), assignable mechanistically to the elementary step comprising the attack of one OH(-) on [Ru(II)(edta)NO](-), with subsequent fast deprotonation of the [Ru(II)(edta)NO(2)H](2-) intermediate. The activation parameters were DeltaH(#)=60+/-1 kJ/mol, DeltaS(#)=-31+/-3 J/Kmol, consistent with a nucleophilic addition process between likely charged ions. In the toxicity up-and-down tests performed with Swiss mice, no death was observed in all the doses administered (3-9.08 x 10(-5) mol/kg). The biodistribution tests performed with Wistar male rats showed metal in the liver, kidney, urine and plasma. Eight hours after the injection no metal was detected in the samples. The vasodilator effect of [Ru(II)(edta)NO](-) was studied in aortic rings without endothelium, and was compared with sodium nitroprusside (SNP). The times of maximal effects of [Ru(II)(edta)NO](-) and SNP were 2 h and 12 min, respectively, suggesting that [Ru(II)(edta)NO](-) releases NO slowly to the medium in comparison with SNP.  相似文献   

11.
The focus of this study is to examine volume and enthalpy profiles of ligand binding associated with CO-Fe(II) tetrakis-(4-sulfonato phenyl)-porphyrin (COFe(II)4SP) in aqueous solution. Temperature dependent photothermal beam deflection was employed to probe the overall enthalpy and volume changes associated with CO-photolysis and recombination. The analysis demonstrates that ligand recombination occurs with a pseudo first order rate constant of (2.5+/-0.2)x10(4) s(-1) (at 25 degrees C) with a corresponding volume decrease of 6+/-1 ml/mol. The activation enthalpy (DeltaH(double dagger)) and volume (DeltaV(double dagger)) change for CO recombination (determined from temperature/pressure dependent transient absorption spectroscopy) are found to be 3.9 kcal/mol and 8.2 ml/mol, respectively. These data are consistent with a mechanism in which photolysis yields a five-coordinate high spin (H(2)O)Fe(II)4SP complex that recombines in a single step to form the low spin (CO)(H(2)O)Fe(II)4SP complex. Base elimination, often associated with CO photolysis from hemes, is not observed in this system. The overall volume changes suggest a transition state with significant high spin character. Furthermore, these results demonstrate the utility of coupling photothermal techniques with variable pressure/temperature transient absorption spectroscopy to probe heme reaction dynamics.  相似文献   

12.
The metallochromic indicator 4-(2-pyridylazo)resorcinol (PAR) has been used at pH 7.0 to monitor the mercurial-promoted Zn2+ release from Escherichia coli aspartate transcarbamoylase and Zn2+ uptake by regulatory dimers upon displacement of the mercurial reagent with 2-mercaptoethanol. The release of Zn2+ (as reflected by a yellow to orange color change in PAR solutions) is linked to dissociation of the enzyme since the six Zn2+ bonding domains stabilize catalytic and regulatory chain contacts; the rebinding of Zn2+ produces enzyme assembly and a corresponding decrease in the amount of PAR-Zn2+ complex. Using greater than 10-fold PAR to free Zn2+ at pH 7.0, delta epsilon = 6.6 +/- 0.2 X 10(4) M-1 cm-1 at 500 nm (20 degrees C) for (PAR)2Zn2+ complex formation (beta'2 approximately equal to 10(12) M-1). In kinetic studies at pH 7.0, PAR (10(-4) M) has been used to measure the instantaneous concentration of Zn2+ released from micromolar quantities of protein; second-order k = 2 X 10(7) M-1 s-1 for forming the 1:1 PAR:Zn2+ complex. These properties of PAR-Zn2+ interactions make PAR a generally useful reagent for studying Zn2+ release from proteins.  相似文献   

13.
On the mechanism of the metallo-beta-lactamase from Bacteroides fragilis.   总被引:1,自引:0,他引:1  
Z Wang  W Fast  S J Benkovic 《Biochemistry》1999,38(31):10013-10023
The catalytic mechanism of metallo-beta-lactamase from Bacteroides fragilis, a dinuclear Zn(II)-containing enzyme responsible for multiple antibiotic resistance, has been investigated by using nitrocefin as a substrate. Rapid-scanning and single-wavelength stopped-flow studies revealed the accumulation during turnover of an enzyme-bound intermediate with intense absorbance at 665 nm (epsilon = 30 000 M(-1) cm(-1)). The proposed minimum kinetic mechanism for the B. fragilis metallo-beta-lactamase-catalyzed nitrocefin hydrolysis [Wang, Z., and Benkovic, S. J. (1998) J. Biol. Chem. 273, 22402-22408] was confirmed, and more accurate kinetic parameters were obtained from computer simulations and fitting. The intermediate was shown to be a novel anionic species bound to the enzyme through a Zn-acyl linkage and contains a negatively charged nitrogen leaving group. This is the first time such an intermediate was observed in the catalytic cycle of a Zn(II)-containing hydrolase and is evidence for a unique beta-lactam hydrolysis mechanism in which the amine can leave as an anion; prior protonation is not required. The electrostatic interaction between the negatively charged intermediate and the positively charged dinuclear Zn(II) center of the enzyme is important for stabilization of the intermediate. The catalytic reaction was accelerated in the presence of exogenous nucleophiles or anions, and neither the product nor the enzyme was modified during turnover, indicating that a Zn-bound hydroxide (rather than Asp-103) is the active site nucleophile. On the basis of all the information on hand, a catalytic mechanism of the B. fragilis metallo-beta-lactamase is proposed.  相似文献   

14.
Some new complexes of mefenamic acid with potentially interesting biological activity are described. The complexes of mefenamic acid [Mn(mef)(2)(H(2)O)(2)], 1, [Co(mef)(2)(H(2)O)(2)], 2, [Ni(mef)(2)(H(2)O)(2)], 3, [Cu(mef)(2)(H(2)O)](2), 4 and [Zn(mef)(2)], 5, were prepared by the reaction of mefenamic acid, a potent anti-inflammatory drug with metal salts. Optical and infrared spectral data of these new complexes are reported. Monomeric six-coordinated species were isolated in the solid state for Mn(II), Ni(II) and Co(II), dimeric five-coordinated for Cu(II) and monomeric four-coordinated for Zn(II). In DMF or CHCl(3) solution the coordination number is retained and the coordinated molecules of water are replaced by solvent molecules. The anti-oxidant properties of the complexes were evaluated using the 1,1-diphenyl-2-picrylhydrazyl, DPPH, free radical scavenging assay. The scavenging activities of the complexes were measured and compared with those of the free drug and vitamin C. We have explored their ability to inhibit soybean lipoxygenase, beta-glucuronidase and trypsin- induced proteolysis. The complex [Mn(mef)(2)(H(2)O)(2)] exhibits the highest antioxidant activity and the highest inhibitory effect against the soybean lipogygenase (LOX), properties that are not demonstrated by mefenamic acid. Their inhibitory effects on rat paw edema induced by Carrageenan was studied and compared with those of mefenamic acid. The complex [Zn(mef)(2)] exhibited a strong inhibitory effect at 0.1 mmol/Kg B.W. (81.5 +/- 1.3% inhibition), superior to the inhibition induced by mefenamic acid at the same dose (61.5 +/- 2.3% inhibition). Mefenamic acid and its metal complexes have been evaluated for antiproliferative activity in vitro against the cells of three human cancer cell lines: MCF-7 (human breast cancer cell line), T24 (bladder cancer cell line), A-549 (non-small cell lung carcinoma) and a mouse fibroblast L-929 cell line. The copper(II) complex displays against T24, MCF-7 and L-929 cancer cell lines, IC(50) values in a microM range similar to that of the antitumor drug cis-platin and they are considered for further stages of screening in vitro and/or in vivo as agents with potential antitumor activity.  相似文献   

15.
1. At neutral pH ferricytochrome c is reduced by the superoxide anion radical (O2-), without loss of enzymatic activity, by a second order process in which no intermediates are observed. The yield of ferrocytochrome c (82-104%), as related to the amount of O2- produced, is slightly dependent on the concentration of sodium formate in the matrix solution. 2. The reaction (k1 equals (1.1+/-0.1) - 10(6) M-1 - s-1 at pH 7.2, I equals 4 mM and 21 degrees C) can be inhibited by superoxide dismutase and trace amounts of copper ions. The inhibition by copper ions is removed by EDTA without interference in the O2- reduction reaction. 3. The second-order rate constant for the reaction of O2- with ferricytochrome c depends on the pH of the matrix solution, decreasing rapidly at pH greater than 8. The dependence of the rate constant on the pH can be explained by assuming that only the neutral form of ferricytochrome c reacts with O2- and that the alkaline form of the hemoprotein is unreactive. From studies at pH 8.9, the rate for the transition from the alkaline to the neutral form of ferricytochrome c can be estimated to be 0.3 s-1 (at 21 degrees C and I equals 4 mM). 4. The second-order rate constant for the reaction of O2- with ferricytochrome c is also dependent on the ionic strength of the medium. From a plot of log k1 versus I1/2-(I + alphaI1/2)-1 we determined the effective charge on the ferricytochrome c molecule as +6.3 and the rate constant at I equals 0 as (3.1+/-0.1) - 10(6) M-1 - s-1 (pH 7.1, 21 degrees C). 5. The possibility that singlet oxygen is formed as a product of the reaction of O2- with ferricytochrome c can be ruled out on thermodynamic grounds.  相似文献   

16.
Both cyclooxygenase and peroxidase reactions of prostaglandin H synthase were studied in the presence and absence of diethyldithiocarbamate and glycerol at 4 degrees C in phosphate buffer (pH 8.0). Diethyldithiocarbamate reacts with the high oxidation state intermediates of prostaglandin H synthase; it protects the enzyme from bleaching and loss of activity by its ability to act as a reducing agent. For the reaction of diethyldithiocarbamate with compound I, the second-order rate constant k2,app, was found to fall within the range of 5.8 x 10(6) +/- 0.4 x 10(6) M-1.s-1 less than k2,app less than 1.8 x 10(7) +/- 0.1 x 10(7) M-1.s-1. The reaction of diethyldithiocarbamate with compound II showed saturation behavior suggesting enzyme-substrate complex formation, with kcat = 22 +/- 3 s-1, Km = 67 +/- 10 microM, and the second-order rate constant k3,app = 2.0 x 10(5) +/- 0.2 x 10(5) M-1.s-1. In the presence of both diethyldithiocarbamate and 30% glycerol, the parameters for compound II are kcat = 8.8 +/- 0.5 s-1, Km = 49 +/- 7 microM, and k3,app = 1.03 x 10(5) +/- 0.07 x 10(5) M-1.s-1. The spontaneous decay rate constants of compounds I and II (in the absence of diethyldithiocarbamate) are 83 +/- 5 and 0.52 +/- 0.05 s-1, respectively, in the absence of glycerol; in the presence of 30% glycerol they are 78 +/- 5 and 0.33 +/- 0.02 s-1, respectively. Neither cyclooxygenase activity nor the rate constant for compound I formation using 5-phenyl-4-pentenyl-1-hydroperoxide is altered by the presence of diethyldithiocarbamate.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
The syntheses, characteristics of dinuclear macrocyclic polyamine zinc complexes and their interaction with plasmid DNA are reported. The two cyclen (1,4,7,10-tetraazacyclododecane) moieties are bridged by rigid and flexible linkages. The crystal structures of Zn2C27H43N8O15Cl4 [5c.(ClO4)3.2H2O] and Zn2C30H43N10O13Cl3 [5e.(ClO4)3.H2O] have been determined. The complexes crystallize in the monoclinic space group C2/c and P2(1)/c with the following unit cell parameters: 5c.(ClO4)3.2H2O: a=32.568(4)A, b=14.8593(17)A, c=19.443(2)A, alpha=90.00 degrees , beta=119.435(4) degrees , gamma=90.00 degrees , Dc=1.551 mg/m3, FW=956.71, F(000)=3932; 5e.(ClO4)3.H2O: a=15.807(2)A, b=16.756(2)A, c=16.161(2)A, alpha=90.00 degrees , beta=97.062(4) degrees , gamma=90.00 degrees , Dc=1.546 mg/m3, FW=988.83, F(000)=2032. The distance between the two Zn(II) ions is about 4.0 A. The structures show that two zinc ions can synergistically interact with the substrate DNA. With this novel structural characteristics, the dinuclear macrocyclic polyamine Zn(II) complexes via the synergetic effect between the two zinc ions can catalyze the cleavage of plasmid DNA (pUC18) with unprecedented speed at physiological conditions.  相似文献   

18.
Previously, we reported that intraperitoneal injections of the Zn(II) complex (Zn(alx)(2)) with allixin, which is isolated from dry garlic, with a Zn(O(4)) coordination environment, exhibited high anti-diabetic effects in obesity-linked type 2 diabetic KKA(y) mice. However, this complex exhibited low activity when administered orally. To improve the effect of Zn(alx)(2), we prepared a novel Zn(II) complex with the allixin-derivative bis(1,6-dimethyl-3-hydroxy-5-methoxy-2-pentyl-1,4-dihydropyridine-4- thionato)Zn(II), abbreviated as Zn(II)-thioallixin-N-methyl (Zn(tanm)(2)), having a Zn(S(2)O(2)) coordination environment; this complex has extremely high in vitro insulin-like activity. Because Zn was extensively absorbed from the gastrointestinal tract when Zn(tanm)(2) was orally administered, its anti-diabetic effects were examined in KKA(y) mice. Daily oral administrations of Zn(tanm)(2) for 4 weeks in KKA(y) mice significantly improved hyperglycemia, glucose intolerance, insulin resistance, hyperleptinemia, obesity, and hypertension. Interestingly, Zn(tanm)(2) increased depressed plasma adiponectin levels in the mice. Here, we propose that Zn(tanm)(2) will be an orally active therapeutic for obesity-linked type 2 diabetes and metabolic syndromes.  相似文献   

19.
The mechanisms of S-nitrosothiol transformation into paramagnetic dinitrosyl iron complexes (DNICs) with thiol- or non-thiol ligands or mononitrosyl iron complex (MNICs) with N-methyl-D-glucamine dithiocarbamate catalyzed by iron(II) ions under anaerobic conditions were studied by monitoring EPR or optical features of the complexes and S-nitrosothiols. The kinetic investigations demonstrated the appearance of short-living paramagnetic mononitrosyl-iron complex with L-cysteine prior to the formation of stable dinitrosyl-iron complex with cysteine in the solution of iron(II)-citrate complex (50-100 microM), S-nitrosocysteine (400 microM), and L-cysteine (20 mM) in 100 mM Hepes buffer (pH 7.4). The addition of deoxyhemoglobin (100 microM) did not influence the process, which points to a direct interaction between S-nitrosocysteine and iron(II) ions to yield DNIC. The reaction of DNIC-cysteine formation is first- and second-order in iron and S-nitrosocysteine, respectively. The third-order rate constant is (1.0 +/- 0.2) x 10(5) M(-2) s(-1) (estimated from EPR results) or (2.0 +/- 0.1) x 10(4) M(-2) s(-1) (estimated by optical method). A similar process of DNIC-cysteine formation was observed in a solution of iron(II)-citrate complex, L-cysteine, and NO-proline (200 microM) as a NO* donor. The appearance of a less stable dinitrosyl-iron complex with phosphate was detected when solutions of iron(II)-citrate containing 100 mM phosphate buffer (pH 7.4) were mixed with S-nitrosocysteine or NO-proline. The rapid formation of DNIC with phosphate was followed by its decay. When the concentration of L-cysteine in solutions was reduced from 20 to 1 mM, the life-time of the DNIC-cysteine diminished notably; this was caused by consumption of L-cysteine in the process of DNIC-cysteine formation from S-nitrosocysteine and iron. Thus, L-cysteine is consumed. Formation of DNIC with glutathione was also observed in a solution of glutathione (20 mM), S-nitrosoglutathione (400 microM), and iron(II) complex (800 microM) in 100 mM Hepes buffer (pH 7.4), but the rate of formation was about 10 times slower than the formation of the DNIC-cysteine. The rate of MNIC-MGD formation from iron(II)-MGD complexes and S-nitrosocysteine was first-order in both reactants. The second-order rate constant for this reaction, estimated from EPR measurements, was 30 +/- 5 M(-1) s(-1). Rate constants of MNIC-MGD formation from iron(II)-MGD and the more stable S-nitrosoglutathione and S-nitroso-D,L-penicillamine were equal to 3.0 +/- 0.3 and 0.3 +/- 0.05 M(-1) s(-1), respectively. Thus, the concerted mechanism of DNIC and MNIC formation from S-nitrosothiols and iron(II) ions can be suggested to be predominant.  相似文献   

20.
The hydrolysis of glycylglycine (GylGly), glycyl-L-leucine (GlyLeu), L-leucylglycine (LeuGly) and glycyl-DL-serine (GlySer) promoted by a copper(II)- cis, cis-1,3,5-triaminocyclohexane complex [Cu(II)TACH] was investigated at 70 degrees C and pH 7-10, using HPLC. The observed pseudo-first-order rate constants (k(obs)) and rate enhancing factors (REF) were as follows: 4.1x10(-3 )h(-1)(REF=23) for GylGly, 1.6x10(-3 )h(-1)(REF=21) for GlyLeu, 5.1x10(-3 )h(-1)(REF=64) for LeuGly and 9.2x10(-2 )h(-1)(REF=47) for GlySer [pH 8.1, dipeptide 2 mM, copper(II) 2 mM and TACH 2 mM]. Based on the pH dependence and dipeptide concentration dependence of the initial rates and speciation of the Cu(II)-TACH-dipeptide system at 25 degrees C and I=0.1, the reactions proceed via the formation of a ternary complex [Cu(TACH)(dipeptide)](+) as an intermediate followed by OH(-)-dependent and OH(-)-independent paths to give amino acid(s). GylGly, GlyLeu and LeuGly preferred the OH(-)-dependent path, while GlySer preferred the OH(-)-independent path. The latter can be explained by the intramolecular attack of the amide carbonyl group coordinated with its oxygen atom by the OH group in the serine residue. The X-ray crystal structure of [Cu(TACH)(GlyGly)]BPh(4).MeOH confirmed that GlyGly coordinates to copper(II) ion with its terminal amino N and amide O atoms. The crystal structures of [Cu(TACH)(Gly)]BPh(4) and [Cu(2)(TACH)(2)(OH)(2)](ClO(4))(2).NaClO(4).H(2)O are also reported.  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